FprEN - 1992 1 1 BD
FprEN - 1992 1 1 BD
When comparing the carbonation rates observed after 25 years for the outdoor conditions (XC4 in
Fig. 2.4.4) it is seen that they are significantly lower than the reference specimens stored indoor.
They are in fact much lower than the ratios applied in the model calculations in this report, cf. Table
2.4.3.
Fig. 2.4.4: Carbonation rates (mm/year½) observed for concretes stored indoor for 10, 20 and 25
years. Furthermore, the carbonation rates measured for outdoor exposure for 25 years are given for
the same concretes. The w/c ratios and the 28-days strengths are indicated in the diagram.
Fig. 2.4.5 contains the results for the higher quality concretes. Mixes 6-11 are used for many outdoor
applications where carbonation takes place and chlorides are often present (XD1, XS1). Mixes 12-16
are considered to be high-quality concretes for civil structures with harsh exposure conditions (e.g.
bridges). None of the specimens in Fig. 2.4.5 experienced full depth carbonation during the 25-years
period.
One observation from the diagram is that the carbonation rates calculated at 10, 20 and 25 years are
fairly identical (about 10% variation) for most of the concretes, which confirms the square-root time
law. The observations of outdoor (XC4) exposed specimens for 25 years shows to be significantly
lower than the reference (indoor) specimens. For many of the high-quality concretes, only 1 mm car-
bonation depths were measured after 25 years of outdoor exposure, which is in contrast to say 25
mm carbonation depth measured on the same specimens stored inside.
It is also obvious that the differences between the various concretes’ carbonation resistance become
rather small for w/c ratios around 0.40, despite different strengths, w/c and binder compositions.
The average carbonation rates measured for the best concrete qualities lie from 4.5 to 5.5
mm/year½, which is probably corresponding to resistance class XRC7.
For citations, copying and referencing see Note on Front page! page 157
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Claus V Nielsen, DTI, 22 Feb. 2020
CEN/TC 250/SC2/WG1/TG10
Fig. 2.4.5: Carbonation rates (mm/year½) observed for concretes stored indoor for 10, 20 and 25
years. Furthermore, the carbonation rates measured for outdoor exposure for 25 years are given for
the same concretes. The w/c ratios and the 28-days strengths are indicated in the diagrams for each
concrete.
Helland (2016) converts the measured carbonation depths from the actual w/c ratio of the actual
concrete to a reference w/c = 0.60 by means of a Norwegian conversion factor [Skjoldsvold 1986].
However, such a conversion is not applied in Fig. 2.4.6 as it does not provide any increase in the pre-
cision because of the large scatter experienced on these measurements. The results in Fig. 2.4.6
show a slight tendency towards a higher carbonation rate for exposure class XC3 (sheltered) but the
number of observations is too few to show statistical significance. The coefficient of variations range
between 30 to 40% on the 2000 investigation and approximately 58% on the 2013 investigation.
In the 2013 investigation a mean value of 2.7 mm/year½ is found and in the 2000 investigation the
mean value for concretes containing pure CEM I is 1.6 mm/year½ and 2.1 mm/year½ for the con-
cretes with CEM I and SF.
For citations, copying and referencing see Note on Front page! page 158
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Claus V Nielsen, DTI, 22 Feb. 2020
CEN/TC 250/SC2/WG1/TG10
Fig. 2.4.6: Carbonation rates (mm/year½) measured on cores sampled from different outdoor struc-
tures having w/c ratio around 0.60. The typical age oat the time of sampling was 10 years.
It is expected that the concretes represented in Fig. 2.4.6 are similar in terms of performance with
those depicted in Fig. 2.4.4 even though it is recognised that the cement types are not exactly identi-
cal. If the Norwegian data are used as input for the stochastic modelling, then the cumulative distri-
bution of the initiation phase will look like Fig. 2.4.7. The Norwegian requirements call for minimum
cover thickness of 25 mm for exposure class XC4, which is applied, together with an execution toler-
ance of cdev = 10 mm.
If we consider only the initiation phase it is seen that concrete based on fly ash cement only reaches
tini = 40 years if the safety index β=1.5 is to be fulfilled. For the pure Portland cement (CEM I) the ini-
tiation phase easily reaches several hundred years under the same assumptions and when silica
fume is added tini = 85 years is reached. Moreover, since the Norwegian field observations are based
on the square-root time law, which is generally more conservative than the models applied in the
present document (Table 2.4.3) the model predictions in Fig. 2.4.7 will also be conservative.
However, if the execution tolerance is reduced from cdev = 10 mm to say 5 mm (higher precision on
the building site) then the predictions will be less conservative. This is illustrated by comparing Figs.
2.4.7 and 2.4.8. The initiation phase is reduced significantly when the execution tolerance on the re-
inforcement is improved.
For citations, copying and referencing see Note on Front page! page 159
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Claus V Nielsen, DTI, 22 Feb. 2020
CEN/TC 250/SC2/WG1/TG10
Fig. 2.4.7: Cumulative probabilities for the initiation phase calculated on basis on the measured data
from Fig. 2.4.6 and minimum cover thickness 25 mm. It is assumed that the square-root time law is
applicable (b = 0.5). The red circles indicate the failure probability of 7% (β=1.5) which defines the
limit state.
Fig. 2.4.8: Similar to Fig. 2.4.7. Only difference is the assumed execution tolerance on the rebars.
For citations, copying and referencing see Note on Front page! page 160
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Claus V Nielsen, DTI, 22 Feb. 2020
CEN/TC 250/SC2/WG1/TG10
The figures in Table 2.4.6 for outdoor exposed to rain (XC4) are in line with the Norwegian findings
described previously. However, for the XC3 exposure conditions the numbers seem to be higher
than what is observed in the field.
Table 2.4.6: Carbonation rates for different exposure conditions and concrete qualities according to
Lagerblad (2005). Based on CEM I concrete types. The values included in EN 16757 is shown in paren-
theses.
Carbonation rates, mm/year½
Strength class,
< C15 C15 – C20 C25 – C35 > C35
cylinder MPa
Wet/submerged
2.0 (-) 1.0 (-) 0.75 (-) 0.5 (-)
(XC2)
Buried in ground 3.0 (-) 1.5 (1.1) 1.0 (0.8) 0.75 (0.5)
Outdoor exposed
5.0 (5.5) 2.5 (2.7) 1.5 (1.6) 1.0 (1.1)
to rain (XC4)
Outdoor shel-
10 (11) 6 (6.6) 4 (4.6) 2.5 (2.7)
tered (XC3)
Indoors dry (XC1) 15 (16.5) 9 (9.9) 6 (6.6) 3.5 (3.8)
An extensive Finnish investigation on existing concrete facades and balconies made from precast
concrete elements is reported in Lahdensivu (2012). The investigation links the degradation and the
reinforcement corrosion with investigations of the concrete quality, the cover thicknesses and the
carbonation and chloride content. The results from this investigation have also been utilised in a re-
cent article on the role of carbonation on corrosion [Angst et al. 2020].
Carbonation rates were measured from cored samples by means of the square-root time law. Sev-
eral thousand samples were taken from buildings from the 1960’ies up to the 1990ies. Most of the
samples were unprotected from the rain (XC4) and the carbonation rates are of the same order of
magnitude as described Helland (2016), ranging from 2 to 3 mm/year½ (average) having coefficients
of variation from 40 to 60% and in some cases up to almost 100%. Hence, very large variation in the
results exist.
For citations, copying and referencing see Note on Front page! page 161
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Claus V Nielsen, DTI, 22 Feb. 2020
CEN/TC 250/SC2/WG1/TG10
The various field observations seem to confirm the carbonation rates that is typically applied in mod-
elling. Furthermore, they seem to confirm the validity of the simple square-root time law. However,
it is acknowledged that the observations are mainly related to the Northern European wet and hu-
mid climate conditions.
2.4.5 References
Angst, U., et al.: “Corrosion of steel in carbonated concrete: mechanisms, practical experience and
research priorities – a critical review by RILEM TC 281-CCC”, RILEM Technical Letters, Vol. 5, pp. 85-
100, 2020
Bager, D.: “Aalborg Portland durability project – 25 years judgement”, Nordic Concrete Research
Mini-Seminar on Exposure Sites, 2008.
fib: Model Code for Service Life Design, Bulletin no. 34, International Federation of Structural Con-
crete, Lausanne, Switzerland, 2006.
Greve-Dierfeld, S.v., et al.: ”Understanding the carbonation of concrete with supplementary ce-
mentitious materials: a critical review by RILEM TC 281-CCC”, RILEM Materials and Structures, Vol.
53, No. 136, 2020.
Helland, S.: “Assessment of carbonation ingress in-field as a tool to calibrate code requirements”,
Proceedings fib Symposium, Cape Town, SA, Nov. 2016.
Lagerblad, B. Carbon dioxide uptake during concrete life cycle, State of the art, CBI Report 2, 2005
Lahdensivu, J.: Durability properties and actual deterioration of Finnish concrete facades and balco-
nies, Tampere University of Technology, Report 1028, 2012
Skjolsvold, O.: “Carbonation depths of concrete with and without condensed silica fume”, Proceed-
ings: Fly ash, silica fume and natural pozzolans in concrete, ACI SP-91, pp. 1031-1048, 1986.
For citations, copying and referencing see Note on Front page! page 162
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
The models applied for the carbonation of concrete are different but give similar results. All models
describe the progress of the carbonation front as a time-power law multiplied by a term describing
the effect of the environment (relative humidity). Moreover, the (admissible) duration of the
propagation phase is deduced from the design service life when considering the carbonation front
progress for deriving the cover depth. The following numbers and differences in the model input
parameters are shown in table 2.5.1.
In Table 2.5.1 fenv and ke are different abbreviations for the effect of environmental condition (relative
humidity) on carbonation rate. The time exponent b describes the progress of carbonation depth with
time. The time exponent can be expresse by b=0.5-w. CO2 is the ambient CO2 concentration and tprop
is the propagation phase.
Table 2.5.1 Summary of model input. These numbers are values taken from subsection 2.1-2.4 and for
more details it is referred to section 2.1-2.4
parameter exposure Model 2.1 Model 2.2 Model 2.3 Model 2.4
class and
related
mean RH
fenv, ke XC2≥ 85% 0.30 0.40 0.50 0.5
RH
XC3: 75% RH 0.80 0.75 0.87 0.87
XC3: 65% RH 1 1 - -
XC4: 75% RH 0.70 0.75 0.83 0.80
For citations, copying and referencing see Note on Front page! page 163
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
* In Chapter 2.1 XC2 is calculated with carbonation at RH = 90% (for the case no leaching is considered), for the case leaching
is considered it is referred to chapter 2.1; in chapter 2.2 XC2 is calculated with carbonation at RH = 85%
Figure 2.5.1a) and b) show the fraction by which each result of model reduces the concrete cover in
the exposure classes compared to XRC reference conditions. In Figure 1 a) and b) 1 means no influence
on carbonation depth and 0.5 means the carbonation depth is half of the initial value). All influences
combined are shown in figure 1 a) and b) right. No safety margins nor safety factors have been applied
on the model parameters in the figures 2.5.1.
For calibration of the input values, model 2.1 used data from exposure tests exposed to a characteristic
average humidity area in Switzerland and a dry region in Switzerland. In the chapter 2.2 the model of
[Izquerdo 2002] was applied with data from Spain and fib MC for SLD 2006]. Model 2.3 used data from
experience of site structures in France and Model 2.4 used boundaries derived from the model
calculations of Model 2.1, 2.2 and 2.3.
Figure 2.5.1: Influence of each model parameter on the concrete cover (carbonation depth at initiation
time)
Figure 2.5.2 shows average values of carbonation depths in dependency of average values of
carbonation rates considering the input parameters summarized in table 2.5.1. Note: Relation to the
XRC definition depends on the probabilistic assumptions, as detailed in subsections 1.2 and in chapter
2.6.
For citations, copying and referencing see Note on Front page! page 164
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Figure 2.5.2: comparison of mean carbonation rate and mean carbonation depth at initiation time
The differences in the models, especially the time exponent, explain some differences in the relation
between XRC designation and mean carbonation rates, due to different allocations of average
carbonation rate after 50 year and average carbonation rate after 1 year. For example, chapter 2.3
the mean carbonation rate after 50 year is translated to the mean value of carbonation rate at 1 year
assuming a 0.45 times power law.
For citations, copying and referencing see Note on Front page! page 165
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Table 2.6.1 displays the recommended minimum cover depth values which are selected for the
Eurocode 2 revision. The values depend on the different exposure classes with respect to carbonation
and on the required design service life (50 or 100 years).
In table 2.6.1 the given concrete cover depth refers to cmin,dur. The correlation between the mean value
of concrete cover and cmin,dur is different for each model. This is explained in chapter 1.2. with further
details in chapters 2.1-2.4. The XRC designation is the carbonation rate derived from the 90% quantile
of carbonation depth under XRC reference conditions. The CoV of carbonation rate expected under
the XRC reference condition is taken different in each model and includes effects of the variability of
the material and of the curing and execution. CoV is in the lower range and was used in the range of
CoV = 25-50%). As detailed in Subsections 2.2 and 2.4 each model has used different CoV of placed
concrete and different definitions of cmin,dur.
As an example, for model 2.1 the carbonation rate is normal distributed. As an example, the mean
value of carbonation rate k of a placed concrete in XRC reference conditions after 50 year of exposure
is km = 3.0 mm/year0.5 and assuming carbonation rate is normal distributed and the CoV of the placed
concrete in XRC reference conditions is CoV = 50% (upper value) the XRC designation is XRC5 and
calculated as follows:
Ø 3.0 mm/year0.5 + 1.282*0.5*3.0= 5 mm/year0.5 → km=3.0 = XRC5
Note: the CoV is assumed under constant relative humidity and moisture conditions including the
effect of execution and production
As another example in model 2.3 the XRC designation associated to the characteristic carbonation
depth not exceeded with a 90% probability in the reference situation at 50 years, helps deriving the
mean value of carbonation rate at 1 year assuming a 0.45 times power law for the carbonation front
in the reference conditions and a 25% CoV (lower value). With these assumptions:
- XRC5 corresponds to kµ = 4.6 mm/year0,5
Derivation of cmin,dur to ensure meeting a target reliability index of 1.5 is detailed in Subsection 2.3
To compare the calculated (and rounded) minimum concrete cover depths from sections 2.1-2.4, the
diagrams in Figs. 2.6.1 and 2.6.2 show the recommended minimum cover depths and the different
model approaches taken from sections 2.1-2.4.
The comparison in Figs. 2.6.1 and 2.6.2 are for service lives of 50 and 100 years, respectively. Only XC2
to XC4 are depicted because in general, the durability modelling for XC1 only causes very limited
provisions for the minimum cover depth as seen in Table 2.6.1.
For citations, copying and referencing see Note on Front page! page 166
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Table 2.6.1: Minimum cover depths cmin,dur recommended in prEN 1992-1-1 (D7), clause 6.5.2.2 (Table
6.3(NDP).
Figure 2.6.1: Comparison of calculated minimum concrete cover depths (rounded values) as a
function of the XRC classes for b = 1.5. The dashed lines are the recommended values given in Table
2.6.1.
For citations, copying and referencing see Note on Front page! page 167
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
For citations, copying and referencing see Note on Front page! page 168
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
3.1.1 Introduction
The transport of chloride ions is for simplicity modelled with the ERFC solution to Fick`s second law of
unidirectional diffusion under the assumption of infinity by delivering a total chloride profile at time t.
Permeation, capillary suction in partially saturated concretes may considerably increase the ingress
rate at the surface.
Because some of the chlorides are bound only part of the chlorides are mobile and affect passivation
performance and corrosion kinetic. Part of the bound and unbound chlorides originate from the raw
materials. The type of compound affects binding capacity (CaCl, MgCl > NaCl). Chloride binding capacity
in turn affects the maximum and total chloride content, which is often higher than what can be
expected from equilibrium and higher than during tests where NaCl is the main source of chlorides.
Carbonation or leaching can liberate bound chlorides that can be washed off or redistribute. This can
results in a peak or increases of total amount of chlorides behind the concrete surface and / or
decreases the amount of total chlorides at the concrete surface.
In addition to that, ongoing hydration, and other sources of alteration of the pore structure, dry out
and temperature variation in turn affects the effective chloride diffusion coefficient with time.
The calculations are performed with an apparent diffusion coefficient at time t determined according
to Figs. 1 considering a chloride peak concentration.
10
CEM I + 4% SF w/z=0.35
XRDS reference conditions
XS1
Dapp(t) [10-12 m2/s]
XS3
1 fenv
fenv a
aXRDS
a
0.1
0.1 1 10
time [year]
Figure 3.1.1 apparent diffusion coefficient and a (a) procedure to design with XRDS based on the
apparent diffusion coefficient after 50 years under XRDS reference conditions and (b) apparent
diffusion coefficients determined in [Costa et al. 2006]
For citations, copying and referencing see Note on Front page! page 169
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
The designation of XRDS classes for resistance against corrosion induced by chloride ingress is derived
from the depth of chlorides (XRDS) penetration [mm] (characteristic value 90 % fractile),
corresponding to a reference chlorides concentration (0,6 % by mass of binder cement + type II
additions), assumed to be obtained after 50 years on a concrete exposed to one-sided penetration of
reference seawater (30 g/l NaCl=3% NaCl-solution) at 20 °C. XRDS has the dimension of a diffusion
coefficient [10-13 m²/s].
The semi probabilistic limit state equation is given with Eq. (2).
≥ ∙ , , Eq. (2)
gf partial factor for the material variable under curing, execution, and moisture conditions [-]
xcrit,k Characteristic value of the depth with corrosion inducing chloride content [mm]
cmin is the design value of concrete cover [mm]
The design value of concrete cover is calculated with Eq. (3).
= −∆ Eq. (3)
Dcdev deviation from nominal cover [mm] and simultaneously the partial factor for the geometry
variable [mm] it depends on the execution class EC
cnom characteristic value of concrete cover [mm]
tini,d initiation time [year]
The design value of initiation time is calculated with Eq. (4)
,
, = − , = − ∙
Eq. (4)
,
For citations, copying and referencing see Note on Front page! page 170
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
xcor,k characteristic value of corrosion loss [mm] prior corrosion induced cracking
vcor,k characteristic value of corrosion rate [µm/year]
Table 3.1.1 summarizes possible partial factors and the reliability level expected if (i) the design
equation is used without corrosion phase bini and (ii) if the design equation is used for corrosion
initiation and propagation phase bini+corr.
Table 3.1.1: Selected partial factors for the additional parameter study
D c [mm] gf [-] (i) gcor [-] (i) (ii) bini [-] (i) bini+cor [-] (i) (ii)
acc. to 1.1 (1.0-1.4) 1.0 ≥ 0.5
Tolerance 1.6 (1.2-1.9) 1.0 (< 1.1) ≥ 1.5
class
For the calibration of partial factors the following parameter study has been applied as summarized in
Table 3.1.2.
Table 3.1.2: Parameter study applied to derive partial factors
Figure 3.1.2 shows a (left) summary of sensitivity factors and figure 3.1.2 (right) shows ranges of
required partial factors in dependency of target reliability.
Figure 3.1.2: sensitivity (left. DAfStb623 and additional according to table 2) and ranges of partial
factors gf (right) for b = 1.5
For citations, copying and referencing see Note on Front page! page 171
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
= − Eq. (6)
Because, the designation of XRDS classes for resistance against corrosion induced by chloride ingress
is derived from the depth of chlorides (XRDS) penetration [mm] (characteristic value 90 % fractile),
corresponding to a reference chlorides concentration (0,6 % by mass of binder cement + type II
additions), assumed to be obtained after 50 years on a concrete exposed to one-sided penetration of
reference seawater (30 g/l NaCl=3% NaCl-solution) at 20 °C. XRDS has the dimension of a diffusion
coefficient [10-13 m²/s], the model has to be modified for placed concrete. The following modifications
have been made:
Ø The model is based on the diffusion coefficient under XRDS conditions (i.e. curing class 2)
DXRDS. DXRDS is related to the apparent diffusion coefficient at time tini by the ageing factor a
which depends on the exposure classes and the conditions like time of first exposure to
chlorides, moisture conditions during exposure, possibility of further hydration and maturing.
Ø An additional factor is included, that accounts for e.g. the diffusion in partially saturated
conditions (fenv) different from XRDS definition.
Ø An initial chloride content is included.
Ø The effect of temperature on the apparent diffusion coefficient is considered from
experimental results.
The model used to calculate the chloride content at the depth of reinforcement at initiation time is
shown in Eq. (7-9).
For citations, copying and referencing see Note on Front page! page 172
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
∆
( , )= + ,∆ − ∙ Eq. (7)
∙ ∙ ∙ , ∙
( )= , · 50 ∙ Eq. (9)
For citations, copying and referencing see Note on Front page! page 173
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Table 3.1.3 Selected ranges of 5 % quantiles of concrete cover (cmin), selected mean values of concrete
cover cm and corresponding standard deviations and partial factors Dc
5% quantile of concrete Mean value of Distribution Dc [mm] Standard
cover = cmin [mm] concrete cover cm function deviation s [mm]
[mm]
15 - 65 5 3
10 - 60 20 - 70 Normal 10 6
5 - 55 15 9
For citations, copying and referencing see Note on Front page! page 174
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Figure 3 summarizes influences on the critical chloride threshold level [DuraCrete2000, Angst et al.
2019, Pettersen 1995]. Additionally, the total critical threshold level increases with chloride binding
capacity, c.p. chapter 2.1.1 and in turn with higher initial chloride content. Additionally, the threshold
level is affected by the rebar quality and production process (e.g. prickled, geometry, hot/cold rolled)
[e.g. Bisschop et al. 2017]. Generally, the chloride inducing threshold level (Ccrit) decreases with
decreasing OH- concentration [e.g. Breit 2006] and therefore with increasing degree of carbonation or
Ca-leaching. Other species are likely to affect the passivation performance.
Table 3.1.6 summarizes critical chloride contents determined from literature allocated to specific
exposure classes.
Table 3.1.6 chloride inducing threshold levels (Ccrit in wt.%/cement), the proposed values from fib 2006
and DuraCrete2000 are not distinguished with respect to the exposure class
Expo- Distribution Mean value m Single values from different sources CoV
sure function [wt.%/cement] [%]
class
XD1 -
XD2 Normal, - submerged samples 0.8-2.3 [Bisschop et al. 2016]
XD3 Lognormal, 0.60 [fib2006], 0.68 [Schiegg et al. 2017],
Beta a=0.2 0.48 [DuraCrete 0.4-1.0 [Greve-Dierfeld et al. 2017] 20 -
XS1 and b=2.0 2000] 1.5 - 2.2 [Frederiksen 2000] 30
XS2 [fib 2006] 0.8 - 2.8 [Frederiksen 2000, Sörensen et al. 2018]
XS3 0.3 - 2.0 [Frederiksen 2000, Sörensen et al. 2018]
0.77 [Markeset 2007]
For citations, copying and referencing see Note on Front page! page 175
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
The maximum chloride concentration depends on for example on the distance to the chloride source,
chloride concentration of the ambient solution, the chloride binding capacity (increases with the
cement content and is dependent on the type of binder), the concrete age / exposure time, the time
of investigation (moisture state). Table 3.1.7 summarizes maximum chloride concentrations (acid
soluble concentrations) in structural elements exposed to chlorides in Switzerland and Germany
[Greve-Dierfeld et al. 2017, Schiegg et al. 2017&2018, Bisschop et al. 2016, Greve-Dierfeld et al. 2017,
Kapteina 2011, Müller et al. 2010, Ungricht et al. 2008, fib Bulletin 76, DAfStb issue 623 2016] and in
XS exposed structures summarized in [fib bulletin 76, Müller et al. 2010 and Nürnberger 1996, Greve-
Dierfeld et al. 2014, Markeset et al. 2010, Helland et al. 2010, Wall 2007, Ghods et al. 2005, Fluge et
al. 2001]. Brackish water or in baltic sea lower chloride concentrations are apparent. For XD exposed
structural elements it was distinguished between XD1 and XD3 by the structure. For example, tunnel,
gallery roofs, bridge bottom views where related to XD1 and the lower part of the tunnel walls, bridge
columns, car park floors, retaining walls to XD3. It should be noted, that especially for XD exposed
structural elements the time of inspection is of interest [Kapteina 2011, Ungricht et al. 2008]. XS
exposed structures were allocated into XS3 until a distance to wave crest of 10 m otherwise the
structural element was allocated to XS2 (submerged) respectively XS1 (> 10 m).
Figure 3.1.4 exemplifies chloride profiles determined from XD-exposed structures and figure 3.1.5
exemplifies chloride profiles from XS-exposed structural elements.
For citations, copying and referencing see Note on Front page! page 176
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Figure 3.1.4: Chloride profiles from XD exposed structures and exposure tests
Figure 3.1.5: Chloride profiles from XS exposed structures and exposure tests
At existing structures in Switzerland C0 of 0.04 wt.%/cement are frequently found. The tolerable
chloride content in raw materials gives raise, that the initial chloride content is around 0.1 wt.-%. The
initial chloride content depends on the mix design. For design of new structures, the assumed initial
chloride concentration must be the same as the chloride content class specified. Current EN 206
allow specifications up to 0.4%.
For citations, copying and referencing see Note on Front page! page 177
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
For citations, copying and referencing see Note on Front page! page 178
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Environmental effect fenv, age exponent under site conditions a and apparent diffusion coefficients
Dapp(t) at exposed structures:
As described initially, the age exponent and the environmental effect fenv are correlated and depend
on the degree of saturation and degree of hydration at the first time of chloride exposure and their
development. The concrete starts with a concrete quality expected after curing protocol curing class
2. If the concrete is directly exposed to water contact with chlorides, ongoing hydration will affect
ageing exponent. In addition to that, precipitates, moving depth with maximum chloride content affect
the age exponent binder dependent.
Ø ongoing hydration, dry out and early chloride exposure (XS1, XD1, XD3)
Ø partially saturated concrete and late chloride exposure (XD1, XD3)
Ø Moving maximum chloride concentration (all exposures)
Ø Ongoing hydration under saturated conditions early exposure (XS2, XD2, XS3)
Ø Ongoing hydration under saturated conditions late exposure (XD2)
In the exposure class XS2 one difference between XRDS model and fib model exist, because of the
exposure time prior to corrosion. The XRDS model starts with an apparent diffusion coefficient
expected after curing protocol according to curing class 2. The fib model starts with a diffusion
coefficient of a concrete cured for 28 day. Therefore, the age exponents in the XRDS model is slightly
higher than in the fib model due to the initially lower apparent diffusion coefficient.
Figure 3.1.6 shows apparent diffusion coefficients of existing structures from Bamforth 2004. The
apparent diffusion coefficients are in the range of < 0.05 up to app. 10 · 10-13 m2/s for FA and BFSC
concretes and 5 up to > 50 · 10-13 m2/s for ordinary Portland cement concrete.
Figure 3.1.7 shows collected apparent diffusion coefficients of single structures with specific
combinations of cement + addition.
For citations, copying and referencing see Note on Front page! page 179
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Figure 3.1.7: Apparent diffusion coefficients from existing structures from site investigations and
data analyses provided by members of fib TG8.5/5.6 [Greve-Dierfeld et al. 2014]
Table 3.1.10 factors for fenv derived from literature review and age exponents under different exposure
conditions and according to figures 3.1.6-3.1.7.
Table 3.1.10 factors fenv and a from existing structures and from experimental investigations of chloride
transports under partially saturated conditions [Olsson2019, Zhang2020, Shah2019, Honglei et al.
2020, Savoye et al. 2018, Shah et al. 2019, Kapteina 2011]
Expo- fenv [-] a [-] Comment and source
sure Distri- mean Co Distri- mean CoV
class bution value V bution value [%]
func- [%] functi
tion on
XD1, Const. 1 0 0.6-0.8 XD1 [Kapteina 2011]
XD3, 1 Beta 0.60 CEM I [Polder2012]
XS1, 1 a=0, 0.65 40 CEM II/B-S, III/A [Polder2012]
tests 1 b=1 0.70 CEM I > 50% BFS [Polder2012]
1 0.80 CEM I < 30% FA [Polder2012]
ND 0.2-0.3 10 0.5-0.6 Wet dry cycles [Gang et al. 2015]
0.1-0.3 - Sd = 80% [Zhang et al. 2020]
0.1-0.5 - Sd = 80% [Savoye et al. 2018]
0.6-1.0 0.6-0.4 55-85%RH [Vera et al. 2007]
0.2-0.6 0.15-0.25 Atmospheric [Costa et al. 2001]
XD2, Const. 1 0 0.27 CEM I [Bamforth 2004]
XS2, 1 DRCM 0.30 CEM I [Gehlen 2000]
XS3 1 DRCM 0.45 CEM I + slag [Gehlen 2000]
1 0.62 CEM I + slag [Bamforth 2004]
1 0.70 CEM I + FA [Bamforth 2004]
1 0.46-0.64 CEM I +5-8%SF [Helland et al.]
1 0.61-0.84 CEM I+5-8%SF+46-48%BFS
1 0.4 CEM I [Polder2012]
1 0.45 CEM II/B-S, III/A [Polder2012]
1 0.5 CEM I > 50% BFS [Polder2012]
1 0.7 CEM I < 30% FA [Polder2012]
1 0.2 CEM I XS3 [Markeset et al. 2010]
1 0.38-0.43 CEM I +8-20%SF
1 0.46-0.52 CEM I+10-35%FA
For citations, copying and referencing see Note on Front page! page 180
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Average app. diffusion coefficient at Initiation time Dapp(tini) in dependency of DXRDS,m [10-13 m2/s]
(average Diffusion coefficient under XRDS reference conditions:
Especially in XD1 conditions, where the initiation time is much lower than the design service life and
the age exponent may differ strongly from age exponents expected under unidirectional diffusion in
saturated conditions, the apparent diffusion coefficient at initiation time differs from DXRDS. Equation
(10) repeats the calculation procedure and table 3.1.11 compares apparent diffusion coefficients at
initiation time and apparent diffusion coefficients under XRDS reference conditions.
( )= , · 50 ∙ Eq. (10)
Table 3.1.11: Example for Diffusion coefficient at initiation time in the exposure class XD1 in
dependency of DXRDS,m
DXRDS,m a [-] aXRDS [-]* tini = 40 year tini = 90 year
[10-13 Dapp(tini) [10-13 m2/s] respectively factor for DXRDS,m
m2/s]
0.8 0.3 0.17 0.09
0.5 0.37 0.19
0.7 0.81 0.42
1 0.7 0.3 0.24 0.14
0.5 0.53 0.30
0.7 1.17 0.66
0.6 0.3 0.35 0.22
0.5 0.77 0.48
* Some single values can be found in section 3.1.7
XRDS is the 90% quantile of the apparent diffusion coefficient expected under XRDS conditions
(DXRDS). XRDS is calculated with Eq. (11).
For citations, copying and referencing see Note on Front page! page 181
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
DXRDS is the mean value of XRDS designation in [m2/s]; apparent diffusion coefficient under XRDS
conditions (curing class 2, 20°C, 30g/L NaCl)
XRDS is the 90% quantile of DXRDS for the case of a normal distribution [m2/s]
Table 3.1.13 summarizes DXRDS,m in dependency of CoV within the lower ranges given in table 3.1.12.
XRDS [10-13 0.5 1 1.5 2 3 4 5 6 8 10 15 20 25
m2/s]
CoV DXRDS,m [10-13 m2/s]
40 0.3 0.7 1.0 1.3 2.0 2.6 3.3 4.0 5.3 6.6 9.9 13.2 16.5
60 0.3 0.6 0.8 1.1 1.7 2.3 2.8 3.4 4.5 5.7 8.5 11.3 14.1
100 0.2 0.4 0.7 0.9 1.3 1.8 2.2 2.6 3.5 4.4 6.6 8.8 11.0
For citations, copying and referencing see Note on Front page! page 182
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Dx mm Beta 10 50 0 50 10 50
xcorr mm Constant 50
vcorr mm/year normal 30 120 20 50 70 120
* Variability is included in the apparent diffusion coefficient.
Figure 3.1.8: Minimum concrete cover in dependency of XRDS for b=1.5 (left) tSL = 50 year and (right)
tSL=100 year
For citations, copying and referencing see Note on Front page! page 183
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
3.1.6 Results
Figure 3.1.6 shows the approximate curves to the median between maximum and minimum depth
xcrit for the exposure classes of XD1/XD2a and the exposure classes XD3/XD2b.
Table 3.1.15 summarizes the curves approximated to the calculated minimum cover.
For citations, copying and referencing see Note on Front page! page 184
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Table 3.1.15 summarizes minimum concrete cover versus approximate values of figure 3.1.9
cmin,dur [mm]
XD1/XD2a XD3/XD2b
b=1.5 b=1.5 b=0.5
XRDS 50 years 100 years 50 years 100 years 50 years 100 years
0.5 20 20 20 30 25 20
1 20 25 25 35 25 25
1.5 20 30 30 40 30 30
2 20 30 35 45 30 35
3 25 35 40 55 35 45
4 25 35 45 65 35 50
5 30 40 50 70 40 60
6 30 40 55 80 40 65
8 35 50 60 90 45 65
10 35 50 65 90 45 70
15 40 55 70 100 55 75
3.1.7 Verification
3.1.7.1 Link between XRDS and durability testing standards
XRDS is the apparent diffusion coefficient Dapp(50year) in reference conditions:
Ø At 20°C
Ø Saturation in 30g/l NaCl
Ø Curing class 2
Actual testing conditions (CEN Test Standards) and national testing standards are listed in table
3.1.16.
From Table 3.1.17 it can be seen that the curing and testing conditions are similar in between XRDS
reference conditions and testing conditions. That means the samples and XRDS means water
For citations, copying and referencing see Note on Front page! page 185
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
saturation over the entire exposure time at constant 20°C. It can be assumed, that the degree of
hydration is similar and the temperature and environmental factors fsd and fT (respectively ke) equals
1. Finally, the age exponent is the one expected in a 30% NaCl solution (or in see water conditions
submerged).
3.1.7.3 Corelations
Figure 3.1.10 shows correlations between curing duration and curing temperature on chloride
migration coefficient from data investigated in [Audenaert et al. 2010, Jacobs 2016, Jung et al. 2018,
Gjorv O. 2014].
Figure 3.1.10: Influence of curing duration dc and curing temperature prior to testing on chloride
migration coefficient
Minimum cover calibration and verification for XRDS
For citations, copying and referencing see Note on Front page! page 186
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Figure 3.1.11 shows (a) apparent diffusion coefficients determined in ponding tests and (b) chloride
migration coefficients. All coefficients have been determined at concretes cured for 28-day prior
vacuum saturation. Curing times different from 28-day are given with empty symbols.
Figure 3.1.11: Chloride diffusion coefficients determined under conditions similar to XRDS definition
(but curing time 28day) (left) and chloride migration coefficients (right)
Figure 3.1.12 summarizes age exponents from literature determined at concretes exposed at an age
of 28 day to conditions similar than XRDS conditions.
Figure 3.1.12: age exponents determined under conditions similar than XRDS reference conditions
but exposed at an age of 28 day
The apparent diffusion coefficient of the placed concrete (Curing Class 2) in XRDS reference
conditions at reference time t = 1 year can be calculated with Eq. (12).
.
(t = 0.0767year) = , / Eq. (12)
D0 is the apparent diffusion coefficient at time t=1 year under XRDS reference conditions in
[m2/s]
aXRDS age exponent [-] under XRDS reference conditions
The XRDS classes and the corresponding diffusion coefficients DXRDS, and those at reference time t0
after placing and curing in CC2 are given D0 are given in Table 3.1.19 in dependency of age exponent
under reference conditions and standard deviation.
For citations, copying and referencing see Note on Front page! page 187
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Table 3.1.19 XRC, carbonation rate at initiation time tini, carbonation rate at reference time t0 after
placing and curing in CC2 are given in Table 9 in dependency of age exponent and standard deviation.
XRDS [10-13 CoV aXRDS 0.5 1 1.5 2 3 4 5 6 8 10 20
mm2/s] [%] [-]
MRCM (t=28 0.0 0.0 0.1 0.1 0.1 0.2 0.3 0.3 0.4 0.5 0.7 1.3
day) [10-12 40% 0.3 0.2 0.5 0.7 0.9 1.4 1.8 2.3 2.8 3.7 4.6 9.2
m2/s] for 0.4 0.4 0.9 1.3 1.8 2.6 3.5 4.4 5.3 7.1 8.8 18
fMnss=1 0.5 0.8 1.7 2.5 3.4 5.1 6.8 8.4 10 14 17 34
0.6 1.6 3.2 4.8 6.5 9.7 13 16 19 26 32 65
MRCM (t=28 0.0 0.0 0.1 0.1 0.1 0.2 0.2 0.3 0.3 0.5 0.6 1.1
day) [10-12 60% 0.3 0.2 0.4 0.6 0.8 1.2 1.6 2.0 2.4 3.2 3.9 7.9
m2/s] for 0.4 0.4 0.8 1.1 1.5 2.3 3.0 3.8 4.5 6.0 7.5 15
fMnss=1 0.5 0.7 1.4 2.2 2.9 4.3 5.8 7.2 8.7 12 14 29
0.6 1.4 2.8 4.1 5.5 8.3 11 14 17 22 28 55
MRCM (t=28 0.0 0.0 0.0 0.1 0.1 0.1 0.2 0.2 0.3 0.4 0.4 0.9
day) [10-12 100% 0.3 0.2 0.3 0.5 0.6 0.9 1.2 1.5 1.8 2.4 3.1 6.1
m2/s] for 0.4 0.3 0.6 0.9 1.2 1.8 2.3 2.9 3.5 4.7 5.9 12
fMnss=1 0.5 0.6 1.1 1.7 2.2 3.4 4.5 5.6 6.7 9.0 11 22
0.6 1.1 2.1 3.2 4.3 6.4 8.6 11 13 17 21 43
3.1.8 References
Audenaert A, Yuan Q, De Schutter G (2010) On the time dependency of the chloride migration
coefficient in concrete. Construction and Building Materials 24(3): 396-402.
Sang-Hwa Jung, Hwa-Sung Ryu, Subbiah Karthick, Seung-Jun Kwon (2018) Time and Crack Effect on
Chloride Diffusion for Concrete with Fly Ash. International Journal of Concrete Structures and
Materials 12, Article number: 14.
Kapteina G (2011) Modell zur Beschreibung des Eindringens von Chloriden in Beton von
Verkehrsbauwerken. Dissertation Technische Universität München.
DBV 2015 Concrete Cover and Reinforcement. Securing of Concrete Cover at Design, Production and
Installation of Reinforcement and Concrete according to Eurocode 2
JCSS (2000), JCSS probabilistic model code Part 1: Basis of design. Joint Committee on Structural Safety.
JCSS (1999), JCSS probabilistic model code Part 3: Resistance models. Joint committee on structural
safety.
DAfStb Issue 622 (2016) Durability design rules for XC-exposed structural elements (in German) Greve-
Dierfeld (2016).
Kenel A, Jacobs F (2018) Measurement and evaluation of the cover depths. Final Report Astra AGB
2014/007 (in German).
For citations, copying and referencing see Note on Front page! page 188
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
DuraCrete 2000 Statistical quantification of the variables in the limit state function. BriteEuram final
Report
Alonso C, Andrade C, Farina J, Lopez F, Merino P, Novoa R (1995): Galvanic corrosion of steel in
concrete. Materials Science Forum 1992-1995:899-906
Contectvet (2000) A validated users manual.
Tuutti, K. 'Service life of structures with regard to corrosion of embedded steel' in 'Performance of
concrete in marine environment', ACI SP-65 (1980) pp.223-236.
Hunkeler F, Mühlan B, Ungricht H (2006) Risk to spalling due to reinfocement corrosion (Risiko von
Betonabplatzungen infolge Bewehrungskorrosion). Report No. 603 (in German)
Nürnberger U (1996) Corrosion and corrosion protection in the building industry (Korrosion und
Korrosionsschutz im Bauwesen) ISBN-13 : 978-3000085536
Y. Schiegg, F. Hunkeler, H. Ungricht, D. Keller, Measures to increase durability – Results
to the field study in the Naxberg tunnel, VSS report no. 683 (2017)
Harald S. Müller, Edgar Bohner, Christian Fischer, Joško Ožbolt, Christoph Gehlen, Kai Osterminski,
Peter Schießl, Stefanie von Greve-Dierfeld (2010) Dauerhaftigkeitsbemessung von
Stahlbetonbauteilen auf Bewehrungskorrosion; Teil 2: Dauerhaftigkeitsbemessung. Rissbildung infolge
Bewehrungskorrosion; Verbundfestigkeit von korrodierter Bewehrung in Beton; Voll-probabilistische
Modellierung von Bewehrungskorrosion: Ein Beitrag zur Dauerhaftigkeitsbemessung;
Lebensdauerbemessung-Anwendung und Validierung. DAfStb Issue 602
Tuutti, K. 'Service life of structures with regard to corrosion of embedded steel' in 'Performance of
concrete in marine environment', ACI SP-65 (1980) pp.223-236.
K. Pettersson. Chloride threshold value and the corrosion rate in reinforced concrete. In: Proc. of
the Nordic Seminar. Lund, 1995, 257-266
Angst U, Elsener B, Larsen C, Vennesland Ø (2009) Critical chloride content in reinforced concrete —
A review. Cement and Concrete Research 39(12):1122-1138.
Angst U, Geiker M, Alonso M, Polder R, Isgor B, Elsener B, Wong H, Michel A, Hornbostel K, Gehlen C,
Francois R, Sanchez M, Criado M, Sörensen H, Hansson C, Pillai R, Mundra S, Gulikers J, Raupach M,
Pacheco J, Sagues A (2019) The effect of steel concrete interface on chloride-induced corrosion
initiation in concrete: a critical review by RILEM TC 262-SCI.
Breit W (2006) Critical corrosion inducing chloride content – state of the art and new investigation
results. Beton 145-168.
Boschmann Käthler C, Angst U, Aguilar AM, Elsener B (2019) A novel approach to systematically
collect critical chloride contents in concrete in an open access data base. Data in brief 27:104675.
For citations, copying and referencing see Note on Front page! page 189
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Bischop J, Schiegg Y (2016) Effect of rebar and cement type on the critical chloride content of
cementitious materials. EuroCorr 2016.
Frederiksen, J. M. (2000). On the need for more precise threshold values for chloride initiated
corrosion. Materials and corrosion, 60 (8), 597-601.
Markeset G, Vennesland Ø (2008) Critical chloride content in reinforced concrete. COIN workshop, 5-
6 June 2008, Trondheim, Norway
Sörensen HE, Poulsen SL, Jönsson U (2012) Chloride threshold values from concrete blocks exposed
at rödbyhavn marine field exposure site.
Poulsen SL, Sörensen HE (2012) Chloride threshold values – state of the art. Danish Expert Centre for
Infrastructure Constructions.
Ungricht H, Hunkeler F (2008) Massnahmen gegen chloridinduzierte Korrosion und zur Erhöhung der
Dauerhaftigkeit. Final Report ASTRA 628.
Bisschop J, Schiegg Y, Hunkeler F (2016) Modelling the corrosion initiation of reinforced concrete
exposed to deicing salts. Final Report Astra 676.
Olsson N (2018) Experimental studies of ion transport in cementitious materials under partially
saturated conditions. Dissertation Universite Paris-Est and Lund University
Geng J, Li, LY, Easterbrook D (2016) Effect of carbonation on release of bound chlorides in chloride
contaminated concrete. Magazine of concrete Research 68(7):353-363
Gang X, Yun-Pan L, Yi-Biao S, Ke X (2014) Chloride ion transport mechanism in concrete due to
wetting and drying cycles. Fib Structural Concrete DOI: 10.1002/suco.201400035.
Li K, Zhang Y, Wang S, Zeng J (2018) Impact of carbonation on the chloride diffusivity in concrete:
experiment, analysis and application. Materials and Structures 51-164
For citations, copying and referencing see Note on Front page! page 190
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Wang Y, Nanukuttan S, Bai Y, Basheer PAM (2017) Influence of combined carbonation and chloride
ingress regimes on rate of ingress and redistribution of chlorides in concrete. Construction and
Building Materials 140:173-183.
Liu J, Qiu Q, Chen X, Nigxu H, He Y, Ma Y (2017) Understanding the interacted mechanism between
carbonation and chloride aerosol attack in ordinary Portland cement concrete. Cement and Concrete
Research 95:217-225.
Touil B, Ghomari F, Bezzar A, Khelidj A, Bonnet S (2017) Effect of temperature on chloride diffusion in
saturated concrete.
Care S (2008) Effect of temperature on porosity and on chloride diffusion in cement pastes.
Construction and Building Materials 22(7):1560-1573.
M.N. Balakrishna, Fouad Mohamad, Robert Evans, M.M. Rahman (2020) Chloride Diffusion
coefficient in partially saturated conditioned concrete cubes. International Research Journal of
Engineering and Technology (IRJET) 07(04):4732-4739.
Honglei C, Zuquan J, Tiejun Z, Benzhen W, Zhe L, Jian L (2020) Capillary suction induced water
absorption and chloride transport in non-saturated concrete: The influence of relative humidity,
mineral admixtures and sulfate ions. Construction and Building Materials 236:117581.
https://doi.org/10.1016/j.conbuildmat.2019.117581
Y Zhang, Z Yang, G Ye (2020) Dependency of unsaturated chloride diffusion on the pore structure in
cementitious materials. Cement and Concrete Research 127:106919
SHA Shah, R Liu (2019) Penetration of chloride ion in cement concrete structure exposed to drying
and wetting cycles Journal of Mechanical and Civil Engineering 16(2): 13-21
Fraj AB, Bonnet S, Leklou N, Khelidj A (2019) Investigating the early-age diffusion of chloride ions in
haredening slag-blended mortars on the light of their hydration progress. Construction and Building
Materials 225:485-495
Savoye S, Mace N, Lefecre S, Spir G, Robinet JC (2018) Mobility of chloride through cement-based
materials under partially saturated conditions. Applied Geochemistry 96:78-86.
For citations, copying and referencing see Note on Front page! page 191
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Q Yuan, C shi, G De Schutter, K Audenaert (2009) Effect of temperature on transport of chloride ions
in concrete. Proceeding of second conference of Concrete Repair, Rehbilitation and Retrofitting ,
Dresden, Germany.
Gang X, Yun-Pan L, Yi-Biao S, Ke X (2015) Chloride ion transport mechanism in concrete due to
wetting and drying cycles. Structural Concrete 2:289-295
For citations, copying and referencing see Note on Front page! page 192
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
CHAPTER 3.2
3.2.1 Objective
3.2.2 Chloride induced corrosion
3.2.2.1 Model of the initiation period in marine environments
3.2.2.1.1 Input Parameters of the chloride model and their statistical
characterization
3.2.2.1.1.1 Surface chloride concentration and their CoV
3.2.2.1.1.2 Critical chloride content (Ccr) and its CoV
3.2.2.1.1.3. Ageing factor n and its CoV
3.2.2.1.1.4 Skin zone (x)
3.2.2.1.1.5 Chloride velocity VCl (t) and its CoV
3.2.2.1.1.6 Concrete cover and its CoV
3.2.2.1.1.7 Summary of input parameters of initiation of corrosion
due to chloride ingress
3.2.2.2 Model of the propagation period
3.2.3.2.1 Input Parameters of the propagation model and their
statistical characterization
3.2.2.3 Service life model
3.2.3 Formulation of Limit State Function
3.2.4.1 Sensitivity factors
3.2.4 Design values
3.2.4.1 Design values for the propagation period
3.2.4.2 Cover depths for chloride induced corrosion
3.2.5 References
For citations, copying and referencing see Note on Front page! page 193
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
3.2.1 Objective
The objective of present document consists in establishing the cover depths for 50 and 100
years that fulfil the definition of the Exposure Resistance Classes (ERC) given in chapter 1 in the
Introduction. For that objective the principles given in the Probabilistic Model Code of the JCSS
and the carbonation model of the fib MC2010 have been used.
For achieving that objective, the steps followed are:
Description of the Time-explicit mathematical model used for calculating the service
life to fulfil the definition of ERC
Phases of the model and selected input parameters
Statistical characterization of the input parameters.
Formulation of the Limit state function (LSF). Reliability level of compliance of the LSF.
In present document the reliability factor =1.5 has been adopted.
Probabilistic calculations of the cover depths complying with the ranking of ERC
defined in the chapter 1 of Introduction.
Where:
C0 is the initial chloride concentration of chloride in concrete in %
CS is the concentration at the surface (a fitted value not a real one)
erf is the error function
Dapp (t) is the apparent diffusion coefficient for chlorides at time t, which usually is
estimated with Eq. 3.2.2:
𝑡0 𝑛
𝐷𝑎𝑝 (𝑡) = 𝐷0 ∙ 𝑡
Eq. 3.2.2
For citations, copying and referencing see Note on Front page! page 194
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Probabilistic evaluation of all the input parameters in Eq. 3.2.1 is complex, since a total number
of 5 variables has to be calibrated in a posterior analysis. The equation is then simplified as was
made that of the carbonation model by embodying several parameters in the Velocity of
chloride ingress, VCl. The rearranged equation supposes the following mathematical change of
variables:
Being the Ccr the critical chloride content (in %) it can be used to define the variable :
𝐶𝑐𝑟 −𝐶0
𝜉= Eq. 3.2.3
𝐶𝑠 −𝐶0
and then,
For calculation, all scatter is merged into Vcl and n in order to make calibration easier and
feasible.
For exposure case XS1, surface concentration is dependent on many parameters (seashore
distance, height of exposure, wind direction, wave height, etc.). In figure 3.2.1 are shown the
data used for the calculation of the scatter and due to it a simplified ranking approximation
was made (Izquierdo, D. Andrade, C., 2011).
For citations, copying and referencing see Note on Front page! page 195
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
1
Data
0,9
Cs [% by mass of concrete ]
For citations, copying and referencing see Note on Front page! page 196
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
These data enabled to propose in all cases a CoV = 20% (upper boundary of recorded values)
for being adopted in the calculations.
Table 3.2.2 Values of aging factors used in the calculations and references
Cem type Source nµ nµ
(XS2/XS3) (XS1)
CEM I ((fib), 2015), (Izquierdo, D. 0.45 0.60
Andrade, C., 2011), (Polder, R.B.
Rooij, M.R., 2005)
CEM II/A-V (Izquierdo, D. Andrade, C., 2011), 0.80 0.60
(Polder, R.B. Rooij, M.R., 2005)
CEM III/B ((fib), 2015), (Polder, R.B. Rooij, 0.50 0.70
M.R., 2005)
CEM II/A-D ((fib), 2015) 0.40 0.65
2
1.8 y = 1.6878x-0.657
1.6 R² = 0.7926
Coefficient of Variation
1.4
1.2
1
0.8
0.6
0.4
0.2
0
0 5 10 15 20 25
Apparent Vcl [mm/year^0.5]
Figure 3.2.2 Average and CoV for apparent chloride ingress rate
For citations, copying and referencing see Note on Front page! page 197
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Chloride ingress is dependent of cement type, exposure conditions and concrete quality and
then in a parallel manner than in the case of the carbonation rate, the VCl value has been found
to depend on its average values, as shown in figure 3.2.2. It is a hyperbolic function whose
formula will be used in the calculations for the CoV.
−0.657
𝐶𝑜𝑉𝑉𝐶𝑙 = 1.6878 ∙ 𝑉𝐶𝑙 Eq.3.2.6
The same CoV = 30% than in chapter 2.2 for carbonation is adopted in present calculations
((fib), 2015) (Izquierdo, D. Andrade, C., 2011) (Izquierdo. D, 2001).
Table 3.2.3 Adopted Coefficient of variation of the concrete cover thickness
Type of execution Dist. Type Bias CoV
in situ – normal conditions Log-normal 1.0 30%
Precast – dedicated quality control Normal 1.0 10%
𝑃
𝑡𝑝 = 𝑉𝑐𝑜𝑟𝑟 Eq.3.2.6
𝑐𝑜𝑟𝑟
where tp is the corrosion propagation time in years, Pcorr (µm) is the accumulated corrosion or
attack penetration after a certain period of time and Vcorr (µm/year) is the annually averaged
For citations, copying and referencing see Note on Front page! page 198
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
corrosion rate.
Sensitivity factors and target – reliability values (Izquierdo, 2019) have been calculated for the
input parameters of the model. The results obtained are the following:
a) For the case of chloride induced corrosion (seawater source), the sensitivity factors
are shown in figure 3.2.3 can be deduced that, on the resistance side Cover and
ageing factors are leading values, whereas chloride ingress rate (VCl) is the leading
variable on the action side.
For citations, copying and referencing see Note on Front page! page 199
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Figure 3.2.3 Sensitivity factors for Chloride induced corrosion (max. 50µm loss of rebar section)
For calculation purposes the parameters given in Table 3.2 6 are adopted
Table 3.2.6 Sensitivity factors of the input service life parameters in the case of
carbonation
Variable Name Type
Cover C 0.40 Resistance
Chloride Ingress rate VCl -0.80 Action
Corrosion rate VCorr ~0 Action
Ageing factor n 0.60 Resistance
In the same manner than in the case of carbonation, the summatory > 1, that implies that
the values are slightly conservative. If a further refinement would be required reported values
for could be divided by =1 in order to normalize the values. However, for this application
and in order to follow EN1990 procedure, no normalization to 1 was adopted.
Another conclusion from this sensitivity analysis is that the most sensitive parameter are the
chloride ingress rate and the aging factor. Then corrosion rate in this case is not predominant
as the values are very high resulting in relatively short propagation periods not impacting
significantly in the total service life except as will be justified in next paragraphs.
For citations, copying and referencing see Note on Front page! page 200
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Since the additive definition of service life: initiation + propagation period and given the fact
that concrete cover is only affecting the first one, in order to determine the required cover for
each exposure class and concrete property will obtained by subtraction of propagation period
from the total required service life.
𝑡𝑑𝑒𝑝 (𝑐𝑜𝑣𝑒𝑟) = 𝑆𝑒𝑟𝑣𝑖𝑐𝑒 𝐿𝑖𝑓𝑒 − 𝑡𝑝𝑟𝑜 Eq. 3.2.8
Therefore, design values for several reliability levels are obtained considering a log-normal
distribution, 500 µm as maximum pitting attack giving:
50
𝑡𝑝𝑟𝑜,𝑑 = Eq. 3.2.10
𝑉𝑐𝑜𝑟𝑟,𝑑
Where = -0.30 is adopted for XS cases. Calculated values for corrosion rate and propagation
period until Pcorr = 500 µm (pitting and end of service life) were given in Table 3.2.5. In view of
the short design propagation periods, no propagation has been discounted from the initiation
in the calculation of service life.
The cover depth values are given in terms of cmin,dur (where 10 mm for tolerance is subtracted
from the design value of concrete cover). Table 3.2.7 shows the calculated minimum cover
depth values (in mm) for 50 and 100 years for each ERC (from 0.5 to 7) and exposure classes
XS1 to XS3). They should be rounded to the closest value ranked every 5 mm. The rounded
values are shown in Table 3.2.8.
Table 3.2.7 Calculated minimum values (10 mm were subtracted from the nominal cover used
in the calculations) for 50 and 100 years of service life)
Calculated mínimum cover depths
XS1 XS2 XS3
=1.50 50 years 100 years 50 years 100 years beta=1.50 100 years
XRD 0.5 16.0 18.1 19.0 24.7 32.0 37.4
XRS 1 22.0 24.8 30.0 36.7 42.0 51.0
XRS 2 30.0 34.1 45.0 54.5 56.0 69.5
XRS 3 35.0 41.0 56.0 68.7 68.0 83.4
XRS 4 40.0 46.8 66.0 81.0 78.0 94.9
For citations, copying and referencing see Note on Front page! page 201
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
3.2.5 References
Andrade, C., Alonso, M.C., y Gonzalez, J.A., (1990) “An initial effort to use the corrosion
rate measurements for estimating rebar durability”, in Corrosion Rates of Steel in
Concrete, ASTM STP 1065, eds. N.S. Berke, V. Chacker, y D. Whiting, Philadelphia, USA,
pp.29-37
Andrade, C. (2014). Probabilistic Treatment of the Reinforcement Corrosion. 2013 W.R.
Whitney Award Lecture: Corrosion vol. 70, No. 6.
Andrade C. (2017) -Reliability analysis of corrosion onset: initiation limit state- Journal of
Structural Integrity and Maintenance- Nov Vol. 2, no. 4, 200–208.
https://doi.org/10.1080/24705314.2017.1388693
DAfStb Heft 601 Durability design of reinforce concrete structures for reinforcement
corrosion – Part 1: System relevant parameter of reinforcement corrosion.
(Dauerhaftigkeitsbemessung von Stahlbetonbauteilen auf Bewehrungskorrosion – Teil 1:
Systemparameter der Bewehrungskorrosion) Authors: M. Beck, A. Burkert, A. Faulhaber,
C. Gehlen, J. Harnisch, B. Isecke, J. Lehmann, K. Osterminski, M. Raupach, P. Schießl, W.
Tian, J. Warkus.
DAfStb Heft 602 Durability design of reinfoced concrete structures for reinfocement
corrosion– Part 2: Durability design (Dauerhaftigkeitsbemessung von Stahlbetonbauteilen
auf Bewehrungskorrosion: Teil 2: Dauerhaftigkeitsbemessung) Authors: E. Bohner, C.
For citations, copying and referencing see Note on Front page! page 202
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
For citations, copying and referencing see Note on Front page! page 203
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
François Toutlemonde
Contribution to Part 3:
The class designation for resistance against corrosion induced by chloride ingress (XRDS) is derived
from the depth of chlorides penetration in mm (characteristic value 90 % fractile), corresponding to a
reference chlorides concentration (0,6 % by mass of cement + type II additions), assumed to be
obtained after 50 years on a concrete exposed to one-sided penetration of reference seawater (30 g/l
NaCl) at 20°C. Being based on a depth in mm, it can be related for calibration to objective measures of
chlorides profiles without further model interpretation. Yet for an easier interpretation in relation to
a material property, easier implementation of intermediate classes, and avoidance of numerical
confusion with another property, it has been decided that XRDS should have the dimension of a
diffusion coefficient (10-13 m²/s).
The dimension of a diffusion coefficient is the square of a length divided by a time. The length given in
the definition is the depth of chlorides penetration in mm (characteristic value 90 % fractile), noted dCl,
corresponding to the reference 0,6% chlorides concentration (by mass of cement + type II additions),
assumed to be obtained after 50 years in the reference conditions. The time t50 given in the definition
is 50 years. Thus, XRDS is to be proportional to dCl²/t50 and the only arbitrary choice is the coefficient
of proportionality.
The reason of having chosen a coefficient equal to ¼, which leads to XRDS = dCl²/631 since t50 is
expressed in seconds and dCl in mm, so that XRDS is in 10-13 m²/s, can be justified as detailed hereafter.
The mathematical description generally used as a reference for describing the distribution of the
chloride content in the cover concrete due to chloride diffusion, following e.g. fib Bulletin 34, fib Model
Code 2010 or ISO 16204, derives from the solution of Fick’s second law. The corresponding equation
reads:
C(x,t) = Ci + (Cs – Ci) [erfc {(x – Δx) / 2 (Dapp(t). t)1/2}]
In this equation, x represents the considered depth within the cover concrete and t the duration of
exposure to the chlorides ingress, erfc is the complementary Gauss error function, Dapp(t) is the
apparent chloride diffusion coefficient, Ci is the initial chloride content. In saturated conditions such as
referred to in the XRDS definition, the “convection depth” Δx (which is sometimes also interpreted as
a surface layer where carbonation can take place and disturb chlorides binding) is null. Cs is classically
For citations, copying and referencing see Note on Front page! page 204
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
designated as the “surface concentration”, although this peak chloride concentration is not directly
obtained as if the concrete porosity was filled by the surrounding seawater.
If we assume that the profile corresponding to the 90% fractile of chloride content follows the same
mathematical form, then we can write for t = t50:
0,6 % = Ci + (Cs – Ci) [erfc {dCl / 2 (Dapp(t50). t50)1/2}]
Having defined XRDS = ¼ dCl²/ t50 leads to 0,6 % = Ci + (Cs – Ci) [erfc {XRDS/Dapp(t50)}1/2]
We can then interpret XRDS as the apparent diffusion coefficient at 50 years (characteristic value,
namely corresponding to the 90% fractile profile), which reads XRDS/Dapp(t50) = 1, under certain
combinations of Cs and Ci values.
The reference combination corresponds to chloride concentrations (expressed as a weight ratio to the
cement + type II additions content) equal to Ci = 0,05 % (realistic situation for concrete mixes meeting
the minimum chloride class requirement) and Cs = 3,55 % (which is in line with typical values obtained
from field investigations, e.g. on structures exposed to Atlantic water).
Alternative examples of combinations compatible with interpretation of the XRDS designation as
representative of the apparent diffusion coefficient include:
Ci = 0 (only the propagation of chlorides from outside is considered) and Cs = 3,81 % (by mass of cement
+ type II additions) which is the solution of 0,6 % / Cs = erfc(1) = 0,157.
Ci = 0,1 % (limiting value for chloride class 0,1) and Cs = 3,28 % (by mass of cement + type II additions)
which is the solution of 0,5 % / (Cs – 0,1) = erfc(1) = 0,157.
Ci = 0,2 % (limiting value for chloride class 0,2) and Cs = 2,74 % (by mass of cement + type II additions)
which is the solution of 0,4 % / (Cs – 0,2) = erfc(1) = 0,157.
Moreover, if a surface concentration Cs = 3,3 % is assumed and Ci = 0 is considered (to focus on the
external chloride diffusion only), XRDS/Dapp(t50) is obtained as the solution of the following equation:
0,6 % / 3,3 % = erfc({XRDS/Dapp(t50)}1/2) = 0,1818, thus {XRDS/Dapp(t50)}1/2 = 0,945 leading to
XRDS/Dapp(t50) = 0,893 or Dapp(t50) / XRDS = 1,12
If a surface concentration Cs = 3,9 % is assumed and Ci = 0,1 % is considered, XRDS/Dapp(t50) is obtained
as the solution of the following equation: 0,5 % / 3,8 % = erfc({XRDS/Dapp(t50)}1/2) = 0,1316, thus
{XRDS/Dapp(t50)}1/2 = 1,065 leading to XRDS/Dapp(t50) = 1,134 or Dapp(t50) / XRDS = 0,882
In sum, XRDS = 1/4 dCl²/t50 (which reads XRDS = dCl²/631 with the proper units) is a direct
correspondence between the exposure resistance class designation and the chloride depth. It
corresponds to the characteristic value of the apparent chloride diffusion coefficient at 50 years under
a realistic reference set of assumptions, and in any case leads to XRDS values of the same order of
magnitude as Dapp(t50) for other realistic sets of assumptions. As such, it includes the progressive ageing
in the diffusion process over the 50 years-expected service life.
Presently used mixes generally correspond to depths in chloride penetration under the reference
conditions of XRDS definition, ranging from typically 15 mm (in case of very high-performance
concrete) to up to about 80 mm (in fact for concrete types to be used only in XD1 / XS1 environment),
depending on the severity of XD/XS exposure and on the design service life. This leads to XRDS between
0,5 and 10 which is deemed convenient for class designation and corresponds to recommended values
in Table 6.4. If relevant in the place of use, chloride diffusion in non-saturated conditions for XS1/XD1
might make sense to be considered, it would then turn out possible to include additional classes
corresponding to XRDS15, XRDS20, XRDS25 and XRDS30 corresponding to dCl values equal to 97,3;
112,4; 125,6 and 137,6 mm respectively – as detailed hereafter.
For citations, copying and referencing see Note on Front page! page 205
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
In the approach detailed hereafter, where computations are deterministic while input values include
characteristic values and margins, the limit state function considered is the random variable G(X):
G(X) = tini + tprop - tSL
With tSL = design service life (50 or 100 years)
tini = time until reinforcement depassivation / initiation of corrosion
tprop = time for (acceptable) propagation of corrosion
The design service life is conventional (fixed in reference to EN 1990) and used as deterministic in the
calibration process described hereafter.
The time for propagation of corrosion depends on availability of water and oxygen, which is deemed
to be described by the (environmental) exposure classes. It may also depend on the cracking of the
cover zone, which is limited depending on the (environmental) exposure class (SLS verification
according to EN 1992-1-1, section 9). In a first approach however, it has not been considered as
depending on the concrete itself. Thus, variance of the G(X) variable is obtained as the quadratic sum
of variances of tini and tprop variables because the sources of deviation are independent. In the following
t*prop will be used as “safe” estimate of acceptable tprop and the fractile estimation of tini– (tSL – t*prop)
will be considered for quantifying the reliability index (β) in relation to its target value.
In the case of chlorides ingress, pitting corrosion is the main harmful corrosion mechanism and its
quantitative development with time is not subject of well-established knowledge. Some authors have
attempted to build an equivalence between pitting depth development with time and an equivalent
loss of rebars diameter, so that tprop can be defined from laboratory measurements of corrosion rate.
This approach may be questioned due to the non-constant corrosion rate when chlorides reach the
reinforcement. Therefore, the following times for corrosion propagation t*prop have been considered
from experts’ opinion as characteristic values for the acceptable duration of reinforcement corrosion
propagation under XD and XS classes. These values are somewhat larger than what would derive from
established corrosion current intensity measurements since a time of corrosion “initiation”
corresponding to low-current is included. They however turn out safe with respect to estimated
propagation times for realistic concrete mixes studied in the French national PerfDuB project (El Farissi,
2020).
Under XS3 exposure (tidal and splash zone) the acceptable time for corrosion propagation is almost
negligible due to daily supply of alternatively chlorides, water and oxygen. This time is higher in
submerged conditions (XS2) due to reduced oxygen availability, or at a higher distance from seawater
droplets (XS1) due to reduced chlorides concentration. Moreover, under XS1 conditions, the cover
concrete may not be constantly saturated, which significantly slows down the diffusion process.
Average lower corrosion propagation rates have been considered for XD3 (resp. XD1) due to the
seasonal supply of de-icing salts, as compared to XS3 (resp. XS1). In the case of XD2, often associated
to pools / industrial waters and liquids, a safe low value has been considered due to difficult inspection
For citations, copying and referencing see Note on Front page! page 206
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
The model used for describing the distribution of the chloride content in the cover concrete derives
from the simplified description of the Fick’s second law solution in a saturated porous material. It is
assumed to represent a stabilized profile after some years, following e.g. fib Bulletin 34. The
corresponding equation reads:
In this equation, x represents the considered depth within the cover concrete and t the duration of
exposure to the chlorides ingress, erfc is the complementary Gauss error function, Dapp(t) is the
apparent chloride diffusion coefficient, Ci is the initial chloride content.
Under XS1/XD1 and possibly also XS3/XD3 exposure classes, the chlorides ingress may take place in a
non-saturated material due to concomitant exposure to rainfall and drying. Therefore, the validity of
the equation is limited to x higher than the “convection depth” Δx (which is sometimes also interpreted
as a surface layer where carbonation can take place and disturb chlorides binding). Cs is then
interpreted not as the classically designated “surface concentration”, but as the peak chlorides
concentration which is observed for x = Δx. This description has been agreed in fib MC 2010 and used
e.g. by Rahimi (2020). From field observations, the following values have been assumed for Δx.
Δx (mm)
Design service life (years) XS1 XS2 XS3 XD1 XD2 XD3
50 5 0 8 5 0 8
100 5 0 10 5 0 10
In saturated conditions, the peak / surface concentration Cs can be computed as the product of the
chloride concentration of the surrounding liquid (e.g. seawater in XS2), supposed to be filling the whole
porosity of the material accessible to water, multiplied by an “enrichment factor” which accounts for
evidence of a higher concentration especially in high performance concrete (El Farissi, 2020). This
enrichment factor typically ranges from 2 to 3,5 and decreases with concrete porosity. Since a fix value
has been considered for Δx, this assumption is still safe even for very high-performance concrete
exhibiting an apparent very high enrichment factor, although a moderate modulation has been
considered for this enrichment factor and for concrete porosity depending on the XRDS class, so that
the peak Cs values variations are limited to less than 10% with a typical content of 3,2% of cement +
type II additions.
Due to uncertainty related to XD2 in terms of surrounding chlorides concentrations, in a first approach
the same Cs and Ci chlorides contents have been assumed as in the reference equation corresponding
to XS2. In tidal or splash zones (XS3), where the convection zone is submitted to drying and wetting
For citations, copying and referencing see Note on Front page! page 207
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
cycles, the peak of the profile, which is shifted at the edge of this zone, corresponds to an “enriched”
chlorides content typically equal to 1,2 times the value taken for the XS2 condition. Although fib
Bulletin 76 would advise not to consider it, this choice has been based on typical average field data
and is intended to compensate the deviations between effective chloride profiles, as measured, and
representation of the tail of the curve corresponding to an equivalent diffusion process.
In the case of XD3 exposure, with the typical situation of parts of civil structures exposed to deicing
salts in winter such as in Alpine / mountain regions of Central Europe, the cycles associated to deicing
salts are non-permanent, and the salts concentration should be lower than in XS3, provided water
runoff is correctly ensured at the surface of the member. It should however be noted from recent
chlorides content measurements on a 40 years-old Austrian retaining wall (Binder et al., 2019), that
modelling the chloride profile as stationary is far from observations: location of the peak of the
chlorides content varies from 1 cm-deep in winter to roughly 3 cm-deep in summer, with peaks of
about 10 kg/m3 (related to the concrete) in winter and 6 kg/m3 in summer. These concentrations are
to be compared to sea water (21 g/l for chloride ions) which at the concrete surface (porosity estimated
as 18%) would lead to 4 kg/m3. For the sake of simplicity, it has been considered a peak chloride
concentration in XD3 equal to 1,1 the one corresponding to XS2, and it has also been considered that
the same ratio relating the peak chlorides concentrations applies to x, the depth corresponding to the
reference chloride content. This is not mathematically exact but the order of magnitude is quite
correct.
Finally, in XD1 and XS1 conditions, the salt concentration at the member surface has been taken as half
of the one under XS2 conditions, which accounts for e.g. dilution by rainwater. For the sake of
simplicity, it has also been considered that the same ratio applies to x, the depth corresponding to the
reference chloride content, as the ratio associated to the chloride concentration. This is not
mathematically exact but the order of magnitude is quite correct.
This influence of the exposure conditions on the boundary conditions of the diffusion model is thus
summarized as follows:
Noticeably enough, determination of the chloride profiles from real structures exhibits a high scatter
in the peak concentration, even in the same exposure class, which may depend on local variations of
exposure to rainfall, local humidity and wind exposure, local concrete porosity etc. It has not been
found realistic to use statistical parameters and characteristic values from these experimental
determinations of chloride profiles, which mainly cover XS3 situations and could be controversial.
Taking account of these variations has been addressed by taking conservative values for the ratios XS3
vs. XS2, XS1 vs. XS2, XD3 vs. XS2, and XD1 vs. XS2 peak concentrations, and for the enrichment factor
leading to Cs.
For citations, copying and referencing see Note on Front page! page 208
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
The designation of ERCs for resistance against corrosion induced by chloride ingress (XRDS) is derived
from the depth of chlorides penetration in mm (characteristic value 90 % fractile), corresponding to a
reference chlorides concentration (0,6 % by mass of cement + type II additions), assumed to be
obtained after 50 years on a concrete exposed to one-sided penetration of reference seawater (30 g/l
NaCl) at 20°C. Having the dimension of a diffusion coefficient (10-13 m²/s), it is calculated as the
rounded value of 1/631th of the square of this depth of chlorides penetration in mm, which
corresponds to verifying x (depth) / [2 (XRDS. t (50 years))1/2] = 1
Under these reference conditions, Δx = 0 thus Dapp(t) can be related to XRDS, Ci, the reference chloride
concentration (0,6 % by mass of cement + type II additions), and Cs. However, as indicated previously,
this relation is not unique. It is thus necessary to precise assumptions on porosity and binder content
that have been made in order to express Dapp(t) in relation to XRDS, for further application of the
model.
Presently used mixes in XD/XS environment correspond to depths of chloride penetration under the
reference conditions associated to XRDS definition, ranging from typically 15 mm (in case of very high-
performance concrete) to up to about 80 mm (in fact for concrete types to be used only in XD1 / XS1
environment), depending on the severity of XD/XS exposure and on the design service life. Therefore,
in order to establish the recommended provisions, computations have been done for XRDS classes
ranging from XRDS0,3 to XRDS11. Since resistance to chlorides ingress is positively correlated to
concrete compactness (namely, chlorides diffusivity is positively correlated to porosity), and due to
present deemed-to-satisfy provisions for concrete mixes employed in XD/ XS environments, it has been
considered that the cement + type II additions content vary from typically 280 to 450 kg/m3, and the
water to binder ratio also typically varies from 0,5 to below 0,3 for the range of concrete types
considered, leading to porosity ranging from 12 to 15 %. The resulting correlated material parameter
sets have been considered. Interestingly, the surface chlorides concentration in the reference XS2
conditions slightly varies from 3 to 3,3 % of the binder content, which is considered a safe estimate,
since values representative of the minimum binder contents according to present standards have been
accounted for. Namely, this corresponds to typical peak chlorides contents of 12 to 13 kg/m3 in
submerged concrete, which can be considered a reasonable design value.
Being fixed the depth corresponding to XRDS definition and the reference chloride concentration, the
reference 50 years duration of the diffusion process, and Cs, Dapp can be derived with an upper bound
value considering that the initial chlorides concentration is null, although this is never strictly true (the
diffusion process is maximized). Inversion of the erfc function thus leads to the following Dapp values,
which can be interpreted as a characteristic value (90 % fractile) of the apparent diffusion coefficient
of the concrete at 50 years (namely, chlorides have diffused more slowly or until a lower depth in 90%
of the cases). It can be noted that based on this set of assumptions, Dapp/XRDS varies between 1,11
and 1,22. This ratio exhibits low variability and is close to the 1,0-value referred to in the XRDS
definition, the reason for this being the set of assumptions corresponding to Cs, Ci and the influence
of the porosity and cement + type II addition content considered for the Dapp determination.
For citations, copying and referencing see Note on Front page! page 209
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
In the XRDS definition, the reference chlorides concentration (0,6 % by mass of cement + type II
additions) has been selected as possibly related to the threshold for initiation of the rebars corrosion,
noted Cth, although this concept is highly debated and its quantitative determination quite variable,
including expressing this threshold as a ratio with respect to the binder content. With this definition,
discussions have arisen concerning the interpretation of the chlorides content threshold, the role of
the initial chlorides content, distinction between free and total chlorides content, etc. In a general
approach for calibration it is not possible to comply with all cases. Following (Angst et al., 2009) and
(Cao et al., 2019) a chloride content equal to 0,6 % by mass of cement + type II additions has been
deemed as a relatively low value for representing the threshold triggering the corrosion propagation,
noting the fact that the definition has been associated to total chlorides content.
Moreover, based on these data, it has been noted that this threshold tends to be lower for concretes
with a high porosity (high water to binder ratio) and higher for high performance concrete. It has also
been noted that the threshold would be lower in situations of salty water available (XS2, XD2, XS3,
XD3) and higher in non-constantly saturated situations representative of XD1 /XS1 (Angst et al., 2009).
Erring on the safe side, it has been considered that for the range of XRDS classes of concrete under
consideration, the thresholds would range from 0,5 to 0,7 % by mass of binder in XD1 / XS1 classes,
depending on the concrete porosity, and from 0,4 to 0,6 % by mass of binder in XS2, XD2, XS3, XD3
environments, also depending on the concrete porosity. The following chloride thresholds Cth were
thus considered as corresponding to the end of initiation period. Knowing Dapp and Cs in the reference
For citations, copying and referencing see Note on Front page! page 210
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
conditions, the depth xth corresponding to a probability of 90% not having reached this chloride
content threshold can be calculated by inverting erfc, as an intermediate indicator.
The equation describing the chloride profile is to be applied in the different exposure conditions in
order to determine the depth x corresponding to reaching the chloride threshold Cth at tini = tSL – t*prop:
In this application, Cs, Δx and t*prop are varied depending on the exposure class and design service life
as discussed before. While 20°C has been considered as a reference temperature in the XRDS
definition, it has been deemed that a more realistic calibration should consider an average 15°C
temperature. Based on e.g. (Jones et al., 1995) and further works, the resulting diffusion coefficient
has been multiplied by 0,75, leading to corresponding chloride depths multiplied by 0,866.
In the approach adopted, only chlorides transferred from the environment have been considered,
while the concrete mixes have been assumed to meet current requirements in terms of initial chlorides
content. In fact it is meant that the chloride class requirements are fulfilled and that aggregate shall
For citations, copying and referencing see Note on Front page! page 211
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
meet the standards requirements imposing to be washed, in case of marine origin. Only considering
chlorides brought by external transfer would formally correspond to Ci = 0 in the diffusion equation.
The underlying assumption is that the chlorides initially present in concrete (which are limited by the
chloride class) are mainly bound and do not contribute to the corrosion risk associated to the Cth
content, which has been taken from the lower bound estimate of the corrosion threshold ranges
expressed in terms of total chloride content.
As an equivalence to considering the limit sate function associated to time, the reliability associated to
the durability limit state (target Beta index) has been evaluated as the probability of the specified
minimum cover for durability cmin,dur to be exceeded by the progress of the chlorides ingress above the
considered threshold. Namely, this relies on the definition that cmin,dur = cnom - Δc,dev is a limiting
minimum value in the sense of the execution standard complemented by the provisions valid in the
place of use, so that the probability of having reinforcement at a distance to the surface lower than
cmin,dur is negligible. With this assumption, estimating the probability of reaching the durability limit
state is done without considering the real distribution of rebars location1 around cnom. Subsequently,
the target reliability is thus assumed to be reached when the probability of having a chloride depth
(depth xCl where the chlorides concentration exceed the corrosion threshold) in excess of cmin,dur at
time tSL - tprop is lower than the target reliability.
With a target β equal to 1,3 and an assumed normal distribution this could read:
Deriving xCl from the equation above and adopting it as the recommended cover depth means that this
depth is overpassed by the threshold value of the chlorides profile with a 10 % probability. In fact, the
cover concrete in the considered structure belongs to the ERC considered, namely given the “regular”
curing process (that shall be described in EN 13670) 90 % of it would have a chloride content lower
than the reference value in the reference environment. The distribution thus includes scatter due to
production and execution. The t*prop has been taken as a safe (low) estimate. And, as detailed before,
the other assumptions taken to represent the climate influence are assumed to introduce no unsafe
bias, although they can hardly be reliably attached to a precise fractile.
1
Noticeably enough, no consideration has been given to the “real” location of the reinforcement,
accounting of cnom instead of cmin,dur: especially, no assumption has been deemed as having to be
included concerning the margin for deviation in the rebars placement, except related to the definition
of cmin,dur = cnom – Δc,dev being the lower limiting value for rebars location, because the statistical
distribution of the margin in excess of cmin,dur is highly variable depending on the construction
conditions and production control, which would make a statistical modelling uncertain.
For citations, copying and referencing see Note on Front page! page 212
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
In non-saturated conditions however with probable concomitant carbonation (XS1 and XD1),
uncertainties in the chlorides transfer process motivate taking a higher margin. This margin should
range from 5 to 10 mm depending on the concrete quality, an empirical formula for this extra-depth
in mm has been taken as 0,04 times the square of the concrete porosity (%). Calibration of the
reliability enhancement associated to this extra cover could however deserve additional checking.
When applying the model describing the chloride concentration at the end of the initiation time,
namely tSL – t*prop, the apparent diffusion coefficient considered should be affected by the ageing
factor. Namely, it is empirically admitted that the chlorides profile follows the time and space
dependency according to :
[erfc {(x – Δx) / 2 (Dapp(t). t)1/2}],
with Dapp = D0 (t0/t)a
with the ageing factor α typically ranging from 0,3 to 0,7 depending on the binder type.
Assuming α = 0, namely considering the assumption of a “stabilized profile” without ageing, would
have resulted in extrapolating the definition as follows: for the times to initiation tSL – t*prop considered,
namely from 35 to 98 years, the depth x(tSL – t*prop) corresponding to the reference chloride content
evolves as the square root of time with respect to the reference x(50 years). This has been considered
as excessively severe, although effective evidence of ageing after about 30 years of age appears as
limited. Therefore, an α = 0,2 value has been considered for derivation of cmin,dur values in the different
exposure conditions and design service lives, which corresponds to a variation of x – Δx as a 0,4 time-
power law when extrapolating from the XRDS definition at 50 years to determine the xCl value.
As an illustration of the whole computational sequence for the determination of raw cmin,dur values, the
following examples are given:
A selection of obtained recommended cover thickness raw values is given hereafter and has been used
together with full-probabilistic modelling approaches to draft recommended provisions at the
European level.
For citations, copying and referencing see Note on Front page! page 213
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
For practical implementation of recommended cover depths values that should be multiple of 5 mm,
a safe rounding process has been adopted, as follows:
- for recommended depths ending by 7 to 1, value rounded to the next number ending by 0.
- for recommended depths ending by 2 to 6, value rounded to the next number ending by 5.
Moreover, cmin,dur values higher than 80 mm have been considered as oversensitive to model
assumptions, with limited practical feedback. They would have led to unsatisfactory provisions for
thick unreinforced cover concrete. Therefore, it has been decided to indicate “N.A.” (not applicable) in
the corresponding cells.
In previous calculations, the diffusion coefficient has been taken identical in any of the exposure
classes, corresponding to chlorides diffusion in a saturated porous medium. This assumption is highly
questionable for XD1 and XS1 conditions, especially when interpretation of the limit between these
classes and XS3/XD3 classes (typically resulting from national interpretation in guidance for selecting
exposure classes) clarifies that splash of salty water droplets is excluded in the former ones. Additional
calculations have been done, considering a diffusion coefficient multiplied by 0,377 = 0,856, namely
corresponding to diffusion within concrete pores at a probable 85 % relative humidity (close to the
equilibrium associated to self-desiccation), following models detailed in (Buchwald, 2000) and (Mai-
For citations, copying and referencing see Note on Front page! page 214
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Nhu, 2013). Additional classes corresponding to higher apparent diffusion coefficients appear as
relevant in the continuity of presently accepted cover depths and mixes.
It has also been noticed that for many industrial facilities corresponding to XD2 classification, the
chlorides concentration of the brine is much lower than seawater. An alternative computation has thus
been performed assuming Cs/2.
Taking these assumptions and the rounding process into consideration, and ensuring that XD2
provisions keep stricter than XD1 (light grey cells), leads to the following table of recommended cover
depths. Light blue cells correspond to different provisions between XD1 and XS1, and between XD3
and XS3, respectively. Figures in red correspond to presently recommended cmin,dur values as from Table
4.4N of Eurocode 2 (structural classes S4 for 50 years-service life and S6 for 100 years-service life,
respectively).
For citations, copying and referencing see Note on Front page! page 215
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
- Chlorides ingress model: the Fick’s second law associated to an ideal diffusion model is known to be
simplistic and reasonably valid only for the inner tail of the chlorides profile. The present approach
includes an enrichment factor and a convection depth, which is in line with up-to-date state-of the art.
However, calibration of the associated input is debatable and high variability of field data e.g. for peak
concentration is of clear evidence. Moreover, interaction of chlorides with the concrete depending on
the binder type will significantly affect the effective critical chloride content for corrosion, which can
hardly be considered at the design step when the concrete mix is not known.
- Since the propagation period is relatively short, the sensitivity to the assumption associated to the
critical pitting corrosion depth of the durability limit state is rather low, which limits a source of
divergence in appreciation of the models.
- Climatic conditions: chloride diffusion appears to be significantly affected by the temperature. Local
appreciation of this influence both for XS ad XD classes should be re-considered in National Annexes.
Moreover, appropriate consideration of this effect for environments associated to XD exposure classes
may be more difficult due to higher seasonal variations directly linked to chlorides supply.
- The deterministic computations in the present subsection only rely on a characteristic value of the
expected chloride ingress depth for estimating the material performance, which is consistent with the
XRDS definition. Conversely, the probabilistic models need to use an estimated mean value of the
performance, identified at early age and derived from the ERC definition, and an estimate of the ageing
factor, and have to extrapolate back to a low quantile value, which makes the computation rather
sensitive to assumptions associated to the scatter in material properties and boundary conditions. This
scatter may be real for site concrete, it may also be an artefact of durability parameters measurement.
In view of further comparison between the present approach and full-probabilistic models, the
following assumptions have been made: the coefficients of variation (CoV) associated to the chloride
depth corresponding to the XRDS class designation range from 70% to 100%, their value increases for
higher quality concretes. A direct determination of this coefficient has been associated to the
deterministically computed Dapp(t50) which is used for cover depth derivation. This CoV however
includes uncertainty in the diffusion coefficient associated to the concrete mix, variations in the peak
concentration (or enrichment factor) and scatter and uncertainty in the ageing factor determination.
If a characteristic value of the migration coefficient is used for concrete performance-based
specification, it should be derived from the mean value of the Dapp(t50) associated to the XRDS class,
considering discrete approximate values of the ageing factor, and include a coefficient of variation
associated to the migration test, typically 30%. A selection of corresponding possible provisions is given
hereafter, for the XRDS classes corresponding to the most frequently used combination of exposure /
cover depths.
Dapp(50) Dapp(50)
dCl charac µ M_RCM (90) µ M_RCM (90) specif = 1,6µ
-13
(mm) 10 10-13 10-12 m²/s 10-12 m²/s
m²/s m²/s
Alpha Alpha Alpha Alpha Alpha Alpha Alpha Alpha
class = 0,3 = 0,4 = 0,5 = 0,6 = 0,3 = 0,4 = 0,5 = 0,6
XRDS1 25,1 1,213 0,55 0,27 0,46 0,78 1,32 0,4 0,7 1,2 2,1
XRDS1,5 30,8 1,811 0,831 0,41 0,69 1,18 2 0,7 1,1 1,9 3,2
XRDS2 35,5 2,394 1,112 0,55 0,93 1,57 2,67 0,9 1,5 2,5 4,3
XRDS2,5 39,7 2,954 1,388 0,68 1,16 1,96 3,33 1,1 1,9 3,1 5,3
XRDS3 43,5 3,484 1,657 0,81 1,38 2,34 3,98 1,3 2,2 3,7 6,4
XRDS10 79,4 11,174 5,85 2,87 4,87 8,27 14,05 4,6 7,8 13,2 22,5
XRDS15 97,3 17,195 9,062 4,44 7,54 12,82 21,77 7,1 12,1 20,5 34,8
XRDS30 137,6 34,39 18,125 8,88 15,09 25,63 43,54 14,2 24,1 41 69,7
For citations, copying and referencing see Note on Front page! page 216
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
A similar table could be established for t0 = 28 days, in this case M_RCM(t0) directly refers to the result
obtained from EN 12390-18 testing standard. Yet this option could be less reliable due to possibly
overestimated ageing factors of blended cements (Mai-Nhu et al. 2015; Linger et al. 2017; Denis et al.
2018).
For further use in the comparison with other modelling approaches the correspondence between XRDS
classes and the mean apparent diffusion coefficient is given extensively.
With all above assumptions, it turns out difficult to provide a general straightforward determination
of the reliability index associated to meeting the recommended minimum cover depth. Importance of
scatter in the inputs and sensitivity of the models lead to hardly demonstrating achievement of a target
reliability higher than 1,5. It seems however rather satisfactory to have kept in the determination of
XRDS and XRC provisions this consistent and similar minimum target reliability whatever the type of
exposure and corrosion mechanism, namely for XD, XS and XC environments, since this expresses
equality in the durability objective for similar constructions, whatever their location.
For citations, copying and referencing see Note on Front page! page 217
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
References
Angst U., Elsener B., Larsen C., Vennesland Ø., “Critical chloride content in reinforced concrete – A
review”, Cement and Concrete Research 39 (2009), 1122-1138.
Binder F., Burtscher S., Strauss A., “Prediction of chloride profiles and discussion of time variant
alterations”, IABSE Symposium 2019 Guimarães, Portugal, March 25-27, pp. 1250-1260.
Buchwald A. (2000) “Determination of the ion diffusion coefficient in moisture and salt loaded
masonry materials by impedance spectroscopy”, in 3rd int. Symp., Vienna, pp. 475-482.
Cao Y., Gehlen C., Angst U., Wang L., Wang Z., Yao Y., “Critical chloride content in reinforced
concrete – An updated review considering Chinese experience”, Cement and Concrete Research
117 (2019), 58-68.
Denis P.E., Linger L., Aït Alaïwa A., Ben Attaya S., Hoarau G., Caro J., Cussigh F. (2018), Use of
resistivity as a concrete quality routine control tool. Outcomes of records gathered during 2-year
for the New Coastal Road on Reunion Island (France) offshore viaduct, fib Congress, Melbourne,
Australia.
El Farissi A. (2020) Prédiction de la durée d’utilisation des ouvrages en béton armé par une
approche performantielle dans le cas de la corrosion induite par la carbonatation ou l’attaque des
ions chlorure, Ph.D. Thesis, univ. La Rochelle, France
fib Bulletin No. 34: Model Code for Service Life Design - Model Code”, Schiessl P. et al., Technical
report, 2006.
fib Bulletin No. 76: Benchmarking of Deemed-to-Satisfy provisions in Standards, Gehlen C. et al.,
2015.
Jones M.R., Dhir R.K., Gill J.P. (1995) “Concrete surface treatment: Effect of exposure temperature
on chloride diffusion resistance”, Cement and Concrete Research, 25, 197-208.
Linger L., Mai-Nhu J., Rougeau P., Torrenti J.M., Outteryck C., Denis P.E., Magné S., Dupuy P. (2017),
Implementation of a global durability approach in close cooperation between Owner/Engineer,
Designers and Contractor’s Joint-ventures for the New Coastal Road on Reunion Island (France),
Proceedings of the 2017 fib Symposium, Maastricht, The Netherlands, pp 2601-2612.
Linger L., collected durability indicators during production control of concrete for a large viaduct,
PerfDuB projet, personal communication, 2020.
Mai-Nhu, J., Rougeau, P., Linger, L., Denis, P.E., Magné, S. (2015) The New Coastal Road on Reunion
Island (France): Application of durability model to a real case, fib symposium, Copenhagen.
For citations, copying and referencing see Note on Front page! page 218
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
• The rate of chloride ingress is defined by the apparent diffusion coefficient Dapp [10-12 m2/s]
which is defined at the age 50 years, taking into account the ageing factor.
• The severity of exposure is defined by the chloride surface concentration Cs [%w-c.m.].
• The critical chloride concentration for the initiation of corrosion Ccr [%w-c.m.].
• Initial chloride content Ci [%w-c.m.] of the concrete mix.
• Cover thickness c to the reinforcement [mm].
These parameters are treated as stochastic variables, defined by distribution functions, mean values
and standard deviations, and Monte-Carlo simulations have been applied to obtain the distribution
of the time for the critical chloride concentration to reach the reinforcement.
The calculations herein are based on the error function solution [fib 2006]. The time-dependent in-
gress of chlorides follows the following equation:
𝑥
𝐶(𝑥, 𝑡) = 𝐶𝑠 − (𝐶𝑠 − 𝐶𝑖 ) ∙ erf { } (1)
2√𝐷𝑎𝑝𝑝 (𝑡)∙𝑡
Where C denotes chloride concentration in weight-% of the cementitious material [%w-c.m.]. The
parameters are defined as:
For citations, copying and referencing see Note on Front page! page 219
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Claus V Nielsen, DTI, Feb. 2020
CEN/TC 250/SC2/WG1/TG10
The field observations have shown that the diffusion coefficient decreases with age and the follow-
ing ageing effect model is used:
𝑡 𝛼
𝐷𝑎𝑝𝑝 = 𝑘𝑒 𝑘𝑡 𝐷0 ( 𝑡0 ) (2)
Where = ageing factor, D0 = the reference diffusion coefficient, typically determined experimen-
tally at t0 = 28 days. The two factors ke and kt are taking the temperature influence into account and
the transfer between the field observations and the laboratory results. In the present calculations
both ke and kt are unity.
Another simplification that is applied herein, is the depth x that is used to describe that the peak
concentration of chlorides is located a few millimetres below the exposed surface rather than at x =
0. In the present calculations x is omitted [Nielsen 2019].
When the chloride content in the depth of the reinforcement bars (cover thickness c) reaches the
critical threshold value Ccr for corrosion initiation the initiation period tini is found by inserting (1) into
the following equation:
Propagation phase
When the initiation phase is done, and the chlorides have reached the critical level at the dept of the
reinforcement, it may start to corrode during the so-called propagation phase. The definition of the
design service life is a sum of the initiation and the propagation phases (tSL = tini + tprop ).
In the present calculations the propagation phase has been included in a simplified manner:
• For the less severe exposure conditions XS1 and XD1 a constant propagation phase of tprop =
10 years has been applied.
• For the other exposure conditions where the concrete is directly exposed to chlorides from
deicing agents and/or seawater the propagation period is not taken into account. Thus, tprop
= 0 for all other exposure conditions than XD1 and XS1.
These assumptions are in line with the those given in the previous models.
• Reference diffusion coefficient D0: Normal distributed with mean value and CoV = 20%.
• Ageing factor : Beta distributed between 0 and 1 with mean value and CoV = 20 to 30%.
• Surface concentration Cs: Log-normal distributed with mean value and CoV = 45%.
• Critical threshold level Ccr: Beta distributed between 0.2 and 2 with mean value 0.6 %w-c.m.
and CoV = 25%.
For citations, copying and referencing see Note on Front page! page 220
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Claus V Nielsen, DTI, Feb. 2020
CEN/TC 250/SC2/WG1/TG10
• Cover thickness c: Normal distributed with mean value equal to the nominal cover and
standard deviation = cdev/1.64. Hence, if the tolerance is cdev = 10 mm the standard devia-
tion is 6 mm whis is applied in the examples given herein.
All other parameters are constants. In Nielsen (2019) a large number of combinations of the varia-
bles, corresponding to various binder types, ageing factors, etc., have been simulated by means of
Monte-Carlo and the variation of the ingress rate of chlorides after a certain number of years has
been quantified and collected in a single nomogram as shown the following examples.
Fig. 3.4.1: Diagram illustrating the ageing effect in Eq. (2) up to the age of 50 years.
Note that the unit µm2/s is equal to 10-12 m2/s. the red circles indicate the range of values that are
applied in the next section examples.
For citations, copying and referencing see Note on Front page! page 221
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Claus V Nielsen, DTI, Feb. 2020
CEN/TC 250/SC2/WG1/TG10
Field observations show that for concretes with CEM I and water to binder ratios around 0.40, the
apparent diffusion coefficient would expectedly end up in class III or IV in Fig. 3.4.1. For fly ash ce-
ments the result would be class II and for slag cements we would typically end up in class I or better
[Nielsen, 2019]. Hence, the classes indicated in Fig. 3.4.1 illustrates the actual range of performance
which is often observed in the literature for concrete mixes exposed to chlorides from deicing agents
and seawater [fib 2006] [Helland et al. 2010].
Regarding the ageing effect it can be debated whether it shall be applied for the full duration of the
service life (say up to 100 years or more) or if it should be cut-off after a certain period. It is out of
the scope to go deeply into this discussion and there is hardly any long-term evidence to support it.
However, there has been reported observations from slag cements up to 60 years [Bamforth 2004]
but others seem to suggest a cut-off after 25 to 30 years [Nielsen 2019] which would be a conserva-
tive approach. The different models presented in sections 3.1-3.3 have applied different approaches,
regarding the ageing effect, which will result in differences in the minimum cover thicknesses, c.f.
section 3.5 and 3.6.
In the present calculations the ageing effect beyond 50 years has been neglected which means that
the lines of Fig. 3.4.1 will become horizontal after 50 years. This is considered to be a conservative
approach.
The Dapp(50 years) values that are applied herein are based on actual observations from field expo-
sure sites [Nielsen 2019]. Hence, they are obtained from temperature conditions that are about 8 to
10 ⁰C (annual average). If the correct modification of the diffusion coefficient according to the fib
Model code [fib 2006] were to be applied the reference diffusion coefficient (at 20 ⁰C reference tem-
perature) would be about twice as high as the observed field values. However, this is far from the
realistic experiences from the laboratory. Therefore, it has been chosen to use kt = ke = 1 in (2) which
again is a conservative assumption. More details of the background for this discussion are found in
Nielsen (2019).
The exposure classes XD2 and XS2 are not included in the calculation examples provided in the next
section. The reason for this is that XD1/XS1 typically represents the less severe chloride exposure
and XD3/XS3 represents very severe exposure conditions and XD2/XS2 is found somewhere in be-
tween.
For citations, copying and referencing see Note on Front page! page 222
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Claus V Nielsen, DTI, Feb. 2020
CEN/TC 250/SC2/WG1/TG10
3.4.4 Relationship between the XRDS classes and the diffusion coefficient
The exposure resistance classes against chlorides are denoted XRDSxx, where the designation xx is
the 90%-quantile of the diffusion coefficient under standardised exposure conditions for 50 years
[10-13 m2/s]. This has been explained in detail in the previous sections and it is not repeated here.
The apparent diffusion coefficient Dapp applied in the previous section (Eqs. (1) and (2)) is the mean
value obtained after 50 years. It relates to the XRDS classes and Fig. 3.4.2 shows the dependency
taken from section 3.1 and 3.2, which correspond approximately to the following simple relation-
ship:
This is applied in the examples in the next section. For each of the Dapp values, depicted in Fig. 3.4.2
(red circles), that represent a wide range of realistic concrete qualities to be expected for chloride
exposure, the minimum cover thicknesses are calculated by means of the simplified method outlined
above and in Nielsen (2019).
For citations, copying and referencing see Note on Front page! page 223
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Claus V Nielsen, DTI, Feb. 2020
CEN/TC 250/SC2/WG1/TG10
Fig. 3.4.3: XD1/XS1 minimum cover thickness for design service life 50 years (tini = 40 years + tprop = 10
years) with a probability of failure approximately 10% which corresponds to β ≈ 1.3. Diagram taken
from Nielsen (2019). Red lines and circles represent the input values used in the examples herein.
For citations, copying and referencing see Note on Front page! page 224
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Claus V Nielsen, DTI, Feb. 2020
CEN/TC 250/SC2/WG1/TG10
Fig. 3.4.4 show the similar picture for service life 100 years. Like it was described earlier, the Dapp(50
years) is maintained for the 100 years predictions, which corresponds to the assumption that the
ageing effect is negligible beyond 50 years.
For both Figs. 3.4.3 and 3.4.4 the execution tolerance is cdev = 10 mm and the initial chloride con-
centration is Ci = 0.05 %w-c.m.
Fig. 3.4.4: XD1/XS1 minimum cover thickness for design service life 100 years (tini = 90 years + tprop =
10 years) with a probability of failure approximately 10% which corresponds to β ≈ 1.3. Diagram
taken from Nielsen (2019). Red lines and circles represent the input values used in the examples
herein.
For citations, copying and referencing see Note on Front page! page 225
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Claus V Nielsen, DTI, Feb. 2020
CEN/TC 250/SC2/WG1/TG10
The exposure resistance class XRDS10 has been excluded since it results in excessive cover thickness
and instead XRSD2 has been inserted in Fig. 3.4.5 and XRSD4 has been inserted in Fig. 3.4.6.
Fig. 3.4.5: XD3/XS3 minimum cover thickness for design service life 50 years (tini = 50 years + tprop = 0
years) with a probability of failure approximately 10% which corresponds to β ≈ 1.3. Diagram taken
from Nielsen (2019). Red lines and circles represent the input values used in the examples herein.
For citations, copying and referencing see Note on Front page! page 226
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Claus V Nielsen, DTI, Feb. 2020
CEN/TC 250/SC2/WG1/TG10
Fig. 3.4.6: XD3/XS3 minimum cover thickness for design service life 100 years (tini = 100 years + tprop =
0 years) with a probability of failure approximately 10% which corresponds to β ≈ 1.3. Diagram taken
from Nielsen (2019). Red lines and circles represent the input values used in the examples herein.
As it is seen in the next sections the results for 50 years of service life scenario are corresponding
well with the other models in sections 3.1-3.3. However, when it comes to 100 years of service life
the differences become significant and the model presented herein produces significantly higher val-
ues of the minimum cover thickness than what is seen from the other models. It is believed that
these discrepancies are related firstly to the ageing effect approach taken for the apparent diffusion
coefficient beyond the 50 years of service life. Secondly, the different model approaches contain dif-
ferent assumptions related to the effect of ambient temperature on the apparent diffusion coeffi-
cient.
For citations, copying and referencing see Note on Front page! page 227
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Claus V Nielsen, DTI, Feb. 2020
CEN/TC 250/SC2/WG1/TG10
Table 3.4.2: Minimum cover thickness [mm] from the examples described above.
Exposure
XD1/XS1 XD2/XS2 XD3/XS3
classes
Not ap-
Service life 50 years 100 years 50 years 100 years
plied
XRSD1 25 40 - 40 65
XRDS2 - - - 50 75
XRDS3 35 60 - 55 85
XRDS4 - - - - 100
XRDS6 45 70 - 80 -
XRDS10 55 90 - - -
3.4.8 References
Bamforth, P.B.: Enhancing Reinforced Concrete Durability, Guidance on Selecting Measures for Mini-
mising the Risk of Corrosion of Reinforcement in Concrete, TR61, The Concrete Society, Camberley,
UK, 2004.
fib: Model Code for Service Life Design, Bulletin no. 34, International Federation of Structural Con-
crete, Lausanne, Switzerland, 2006.
Helland, S., Aarstein, R. & Maage, M.: “In-field performance of North Sea offshore platforms with re-
gard to chloride resistance”, Structural concrete, Vol. 11, No. 1, pp. 15-24, International Federation
of Structural Concrete, Lausanne, Switzerland, 2010.
Nielsen, C.V.: “Service Life Modeling of Chloride Ingress”, ACI Materials Journal, Sept. 2019, Vol. 116,
pp. 17-30, 2019.
For citations, copying and referencing see Note on Front page! page 228
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
The models applied for the chloride induced corrosion are similar in all approaches for describing the
incubation and initiation period where the chloride ions progress towards the reinforcement. The
models are based on the erfc solution to Fick`s second law of diffusion.
The XRDS designation represents the apparent diffusion coefficient as a characteristic value (not
exceeded in 90% of the cases) with the units [10-13 m2/s] and refers to the depth where an average
concentration Cth equal to 0.6 wt.% chloride/binder-cement+type II addition is reached after 50 year in
30g/L NaCl of external solution at 20°C by unidirectional diffusion. The XRDS designation refers to a
concrete of at least execution class 2. Under actual exposure conditions temperature and moisture
condition, the chloride load or the chloride content at reinforcement that may lead to corrosion
initiation may differ from the XRDS-reference condition and XRDS-reference depth.
Table 3.5.1 summarises the parameters used in the erfc solution models. Furthermore, table 3.5.1
allocates the defined XRDS reference conditions to model parameters and model factors. The model
factors equal 1 if the condition in the exposure class is like the XRDS reference condition. The model
parameter has the same average value as the parameter of the reference condition if the exposure
condition equals (in average) the reference condition.
Table 3.5.1 summary of mean values of model parameter and transfer parameter in comparison with
XRDS-reference conditions
parameter XRDS-reference Exposure Model Model Model Model
condition class 3.1 3.2 3.3 3.42)
XS1 0.65 0.6 0.2 0.3-0.6
Time exponent Apparent XD1 0.50 - 0.2 (to 50 y)
a=n diffusion XS2 0.3-0.6 0.8 0.2 -
coefficient XD2 0.40 - 0.2
(Dapp at t=50 year) XS3 0.3 - 0.6 0.8 0.2 0.3-0.6
XD3 0.40 - 0.2 (to 50y)
XS1 0.56 1
XD1 0.70 -
Factor for the 20°C: XS2 0.56 1 0.75 1
effect of (ke=fT=1) XD2 0.60 -
temperature [-] XS3 0.56 1
k e = fT XD3 0.70 -
Factor for the XS1 1 1 (1)1)
effect of Fully saturated XD1 0.4 -
transport in condition XS2 1 1 1
partially satu- (fenv = 1) XD2 0.8 - 1 1
rated conditions XS3 1 4
fenv [-] XD3 0.5 -
For citations, copying and referencing see Note on Front page! page 229
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Each factor and each average model parameter, that differs from the XRDS reference condition affects
the calculated depth with the specific chloride content. For example, if the temperature is lower than
in the temperature in XRDS reference conditions, the depth allocated to XRDS reference condition is
reduced in dependency of the factors proposed by each model in table 3.5.1. Another example is the
critical chloride content. If the critical chloride content is higher than the reference `fictive` threshold
given for XRDS reference depth, than the XRDS depth is reduced following the erfc-solution model.
For citations, copying and referencing see Note on Front page! page 230
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Fig. 3.5.1 average effect of each model parameter on depth with average chloride content of 0.6
Figure 3.5.2 shows the calculated depth with critical corrosion induction chloride content in each
exposure class for average apparent diffusion coefficients after 50 years of exposure to XRDS reference
conditions.
For calibration of the input values, model 3.1 used data from exposure tests and existing structures in
Switzerland and literature (i.e. from Denmark, Sweden, France, Denmark). In the chapter 3.2 the model
of [Izquerdo 2002] was applied with data from Spain and fib MC for SLD 2006]. Model 3.3 used data
from experience of site structures in France and Norway. Model 3.4 used boundaries derived from the
model calculations of Model 3.1, 3.2 and 3.3, from data from Denmark and fib MC for SLD 2006.
For citations, copying and referencing see Note on Front page! page 231
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
For citations, copying and referencing see Note on Front page! page 232
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Figure 3.5.2: Average depth with critical chloride content xcrit according to each model in
dependency of average apparent diffusion coefficient determined after 50 years in XRDS reference
conditions
The differences in the models, especially the critical chloride content, chloride surface concentration
and initial chloride content explain some of the differences. Others are explained by the different
environmental conditions. Furthermore, age exponents different from XRDS definition i.e. under dry
conditions (XD1, XS1) explain the differences in XD1/XS1 conditions.
For citations, copying and referencing see Note on Front page! page 233
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Table 3.6.1 displays the recommended minimum cover depth values in dependency of XRDS classes
which have been selected for the Eurocode 2 revision. The values depend on the different exposure
classes and on the required design service life (50 or 100 years).
In table 3.6.1 the given concrete cover depth refers to cmin,dur. The correlation between the mean value
of concrete cover and cmin,dur is different for each model. This is explained in chapter 1.2. with further
details in chapters 3.1-3.4. Examples are given in the following.
The XRDS designation has the units of a diffusion coefficient in 10-13 m2/s derived from the 90% quantile
of depth with 0.6 wt.% chloride/binder-cement+type two addition after 50 year in XRDS reference
conditions. The CoV of apparent diffusion coefficient under XRDS reference conditions (respectively,
the CoV of the depth after 50 years under XRDS reference conditions) includes the following effects:
But excludes variability of environmental conditions and the variability of critical chloride content.
CoV was used with CoV = 40-100 %. The lower the CoV used to calculate the XRDS values based on the
mean values from figure 3.5.1 the more conservative is the XRDS class designation but the lower the
CoV used to calculate the mean value from XRDS value, the less conservative is the concrete cover.
As an example, if the apparent diffusion coefficient (respectively, the depth with chloride content of
0.6 wt.%/binder after 50 years of exposure in XRDS reference condition) is normally distributed and if
the mean value is Dm(XRDS)= 3.97 10-13 m2/s and assuming the CoV of the placed concrete in XRDS
reference conditions is CoV = 40% (lower value of the range) the XRDS designation is consequently
XRDS 6 calculated as follows:
Ø 3.97 10-13 m2/s + 1.282*0.4*3.97 = 6 10-13 m2/s → Dm(XRDS)= 3.97 10-13 m2/s = XRDS 6
Subsections 3.1-3.4 give details on the CoV and calculation procedure used to determine the 90%
quantile of the apparent diffusion coefficient, in dependency of the selected distribution function.
Table 3.6.1: Minimum cover thickness recommended in prEN 1992-1-1 (D7), clause 6.5.2.2 (Table
6.3(NDP).
For citations, copying and referencing see Note on Front page! page 234
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Accounting for the different possible interpretations in the target reliability it was agreed to select for
the Eurocode revision the average of recommended values from Figs 3.6.1, which are displayed in
Table 3.6.1. For values in excess of 80 mm deviations between models was considered excessive and
corresponding minimum cover depths values were decided not to be recommended (indication NA –
“not applicable”).
The XRDS classes are assumed safe for structural elements exposed to Atlantic-sea / Mediterranean
sea water. Because de-icing salt exposure is assumed similar or less severe, but highly variable through
Europe the provisions for XS and XD have been merged.
Figure 3.6.2: Comparison of calculated minimum concrete covers (rounded values) as a function of the
XRDS classes for b = 1.5. The dashed lines are the recommended values given in Table 3.6.2. Only
XD1/XS1 and XD3/XS3 exposure classes are shown as examples. Note that the x-axes differ
For citations, copying and referencing see Note on Front page! page 235
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
02.03.2021
For citations, copying and referencing see Note on Front page! page 236
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Section 6.5.2.2 Background to FprEN 1992-1-1:2023
Introduction
This document has been prepared to explain the increase of minimum cover for durability for prestressing
tendons, as an addition to the minimum cover for durability recommended for reinforcing steel. The text in this
document is proposed to be implemented into the background document to clause 6.5 Concrete Cover.
References [1] Ganz H.R. et al., “Polymer-duct systems for internal bonded post-tensioning”, fib Bulletin 75, Lau-
sanne, 2014.
28.04.201
For citations, copying and referencing see Note on Front page! page 238
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Introduction
This document has been prepared in order to validate the proposed EC2-provisions to increase the minimum
cover by cmin considering the increased uncertainty and variability of concrete cast directly against soil.
Figure 1: Images of a concrete core extracted from a DW and impregnated with fluorescent epoxy. Areas with
high porosity under normal light (left) and ultraviolet light (right) appear green
Examples of soil inclusion are given in Figure 2-4 showing veins/voids with lighter colour and higher porosity in
the cover zone, i.e. interior "defects" not visible from the outside. For concrete a light discoloration of the
cement paste is an undisputable indicator of reduced concrete quality enabling aggressive ions to move much
more rapidly towards the reinforcement than unaffected concrete does.
C. Edvardsen 28.04.2021
For citations, copying and referencing see Note on Front page! page 239
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Figure 2: Inferior concrete quality at extracted Figure 3: Polished section of a core from DW
core from SP due internal soil inclusions (not showing veins/voids (light colour) in the cover
necessarily always visible at the concrete zone filled with trapped support fluid/bentonite
surface) mixture
[1] M.M.R. Boutz, B.J. Admiraal, R. van Berkel, P.R.B: van der Werff, “X10 Concept to improve quality
References
and durability of diaphragm walls", DFI-EFFC International Conference on Deep Foundations and
Ground Improvement, Rome, Italy, 2018.
[2] M. Czopowska-Lewandowicz, “Concrete properties in diaphragm wall embedded in non-cohesive
soils". MATEC Web of Conferences 174, 01007 (2018) ECCE 2018
[3] I.H. Wong, “Experience with waterproofness of basements constructed of concrete diaphragm walls
in Singapore”, Tunnelling and Unterground Space Technology, 12, 4, p. 491-495, 1997.
[4] Undisclosed major infrastructure project in Europe. Confidential.
C. Edvardsen 28.04.2021
For citations, copying and referencing see Note on Front page! page 240
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Berlin, 15.01.2021
For citations, copying and referencing see Note on Front page! page 241
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Introduction
This document has been prepared in order to explain the proposed EC2-provisions for the nominal cover of
bored piles and diaphragm walls according to concrete design with EC2 and execution requirements accord-
ing to EN 1536/EN 1538.
6.5.1(3) Problem
Nominal cover for bored piles
and diaphragm walls The geotechnical execution standards EN 1536 [1] and EN 1538 [2] contain design and execution rules for
concrete members according to Figure B.6.5.1 and B.6.5.2:
Figure B.6.5.1: Bored pile with circular and non Figure B.6.5.2: Diaphragm wall – Geometry of a panel
circular cross-sections (barrettes) [1] [2]
The definition of cover in EN 1536 [1] and EN 1538 [2] is:
Cover: Distance between the outside of the reinforcement cage and the nearest concrete surface
([1] 3.39 and [2] 3.13)
The requirements to concrete cover in the geotechnical standards are summarized in Table B.6.5.1.
The requirement of 7.7.1 in [1] and [2] seems clear. The design rules of EC2 for minimum cover for durability
and bond apply.
The requirements of 7.7.2 in [1] and [2] are not so clear. The meaning of cover in relation to execution as
minimum or nominal cover is open.
The NOTE to 7.7.2 should give clarity about the nominal cover. But there is an inconsistency: It is stated, that
the minimum cover of EC2 or the minimum cover in relation to execution is the nominal cover to be used
for bored piles or diaphragm walls. This is confusing and inconsistent.
From the EC2-designers view it is clear, that a minimum cover has to be increased by an allowance of
deviation considering the execution tolerances to specify the nominal cover using in design documents.
If the cover in relation to execution according to 7.7.2 in [1] and [2] is really the nominal cover including exe-
cution tolerances, it has to be clarified by the responsible standardisation committees of CEN/TC 288. An
improvement of the NOTE to 7.7.2 in next issues of [1] und [2] is necessary.
Table B.6.5.1: Requirements to concrete cover for bored piles and diaphragm walls in geotechnical standards
EN 1536 [1] 7.7 Minimum and nominal cover EN 1538 [2]: 7.7 …
7.7.1 The minimum cover in relation to environmental conditions and to adhesion shall comply with EN 1992 (all parts).
7.7.2 The minimum cover in relation to execution shall not be less than: 7.7.2 The minimum cover in relation to
– 75 mm for barrettes; execution shall not be less than 75 mm.
– 60 mm for piles with D > 0,6 m; or
– 50 mm for piles with D ≤ 0,6 m,
– unless otherwise agreed.
NOTE The minimum cover in relation to execution (in order to ensure that the concrete flows freely) is specified by
reference to target values and not by reference to execution tolerances. So the nominal cover is the greater of the
minimum cover in relation to environmental condition and to adhesion and the minimum cover in relation to execution.
7.7.3 The minimum cover in relation to execution should be increased to –
75 mm where:
– piles penetrate soft soil and are constructed without a casing;
– submerged placement of concrete with 32 mm max. aggregate is used;
– silica fume is used as cement replacement;
– reinforcement is installed subsequent to concrete placement; or
– the bore hole walls have uneven surfaces.
7.7.4 The minimum concrete cover in relation to execution may be re- –
duced to 40 mm to the external face of a permanent casing or lining,
where used.
F. Fingerloos 15.01.2021
For citations, copying and referencing see Note on Front page! page 242
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Due to the misunderstanding NOTE to 7.7.2 in [1] und [2] the clarifying clause 6.5.1 (3) is implemented in
FprEN 1992-1-1:2023:
(3) For bored piles and for diaphragm walls the nominal cover values for durability and bond of this Euro-
code apply. The cover values according EN 1536 and EN 1538 considering the type and deviations of
execution should be checked additionally. The largest value according to this Eurocode and EN 1536 or EN
1538 applies.
So it is clear, that the nominal cover of EC2 including the allowance of deviation cdev according to 6.5.3 is to
consider in eyery case (see extract of Table 6.7(NDP). Only if the cover of execution in EN 1536 or EN 1538
is greater, then these values apply. The designer shall to check both values.
Table 6.7(NDP) – Allowance for deviation cdev in EC2
Case cdev
8 Concrete cast against prepared ground (including uneven blinding layer) 40 mm a)
9 Concrete cast directly against unprepared soil 75 mm a)
a)
These allowances for deviation cdev apply for the nominal cover according to this Eurocode also for
bored piles and for diaphragm walls unless a National Annex gives other special values.
With two examples for bored piles the design of cover may be illustrated (from CEN-TC250-SC2-
Example 1
WG1_N1050).
Example 2
F. Fingerloos 15.01.2021
For citations, copying and referencing see Note on Front page! page 243
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Background to
6.5.3 Allowance in design for deviation
06.05.2021
For citations, copying and referencing see Note on Front page! page 244
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Introduction
This document has been prepared in order to explain the proposed design values for the allowance in deviation.
6.5.3 Allowance in design The values for the allowance in deviation cdev are concentrated in one Table 6.7(NDP). The designer has to
for deviation choose an appropriate deviation value for nominal cover cnom under consideration of the expected execution
quality and the necessary reliability to reach the minimum cover cmin in the finished member.
EN 13670 [1] contains geometrical tolerances for execution in clause 10. The tolerance classes are consid-
ered as:
Tolerance class 1: normal tolerances,
Tolerance class 2: lower tolerances intended to be used with adjusted material factors in EN 1992-1-1,
Annex A (in prEN 1992-1-1-D7: Table A.1a(NDP), case (a)).
10 mm
In general: for execution in To- value from
lerance class 1 according to EN 1992-1-1:
1 EN 13670 2004:
Standard cases in buildings by 4.4.1.3(1)P,
usual execution quality. accepted by
22 NSBs [2]
5 mm
Where fabrication is subjected value from
to a quality assurance system, EN 1992-1-1:
3 in which the systematic monito- 2004:
ring includes measurements of 4.4.1.3(3)
the cover 10 mm …
5 mm (4.3N)
0 mm
Where it can be assured that an
value from
accurate measurement device is
EN 1992-1-1:
used for systematic monitoring
4 2004:
and non conforming members
4.4.1.3(3)
are rejected (e.g. precast ele-
5 mm … 0 mm
ments)
(4.4N)
For concrete members in expo- 5 mm
sure class XC1, where the risk (but not in
5 of corrosion is insignificant combination See line 2
New proposal considering, that a with cmin,b for
shortfall of cmin,dur is without risk. bond)
For concrete cast against sur-
faces with exposed aggregate
(e.g. interfaces)
New proposal considering
- the lower corrosion risk inside a
concrete member without direct 5 mm
6
exposion due to environmental
condition,
- the lower bond failure risk inside
a concrete member due to redu-
ced cracking or spalling of « in-
ternal » cover.
Draft 2 06.05.2021
For citations, copying and referencing see Note on Front page! page 245
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
10 mm +
For concrete cast against une- dimension of
venness due to formwork or ex- unevenness
7
cavation sheeting (e.g. ribbed fi- similar to
nishes or architectural textures) EN 1992-1-1:
2004: 4.4.3(4)
40 mm a)
(increased mi-
nimum cover
EN 1992-1-1:
Concrete cast against prepared 2004: 4.4.3(4),
8 ground (including uneven blin- accepted by
ding layer) 20 NSBs [2],
here adopted
as increased
cdev on safe
side)
75 mm a)
(increased mi-
nimum cover
EN 1992-1-1:
2004: 4.4.3(4),
Concrete cast directly against
9 accepted by
unprepared soil
20 NSBs [2],
here adopted
as increased
cdev on safe
side)
a)
These allowances for deviation cdev apply for
the nominal cover according to this Eurocode
also for bored piles and for diaphragm walls
unless a National Annex gives other special va-
8/
lues.
9
See also BD to 6.5.1 (3).
Depending on real execution methods and national
experience adjusted values may be given.
Post-installed reinforcing bars
5 mm or
Ensuring the minimum concrete
according to
cover cmin for post-installed reinfor-
a project
cing bars in all directions
specification
- according to Table 11.2 for cmin,b
(reduced value
depending on the drilling
because the
10 method and the execution
real direction
procedure,
and length of
- according to Tables 6.3(NDP)
the drilling
and 6.4(NDP) or Q.3(NDP) for
hole can be
cmin,dur,
controlled and
- or according to Table P.2(NDP)
checked)
for cmin,dur.
Draft 2 06.05.2021
For citations, copying and referencing see Note on Front page! page 246
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Geometrical imperfections
23-12-2022
For citations, copying and referencing see Note on Front page! page 247
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.2.1 Background to FprEN 1992-1-1
Introduction
This document has been prepared in order to validate the proposed EC2-provisions for geometric imperfections
that should be considered in the structural analysis to determine the internal forces in bracing structures and
vertical load bearing elements.
When imperfections are considered a distinction can be made between local and global imperfections. Local
imperfections are imperfections that will influence the internal forces in one element only. An example of a local
imperfection is a deviation from an assumed perfect straightness of a column. Global imperfections on the other
hand are imperfections that will influence the internal forces in the bracing elements of a structure like shear
walls and floor diaphragms.
In nowadays construction methods, hybrid structures are used more often. For instance for structures with
braced steel columns and concrete shear walls it is of importance that rules for global imperfections are
harmonized over the several Eurocodes. This to prevent that structural engineers have to make a choice
between the global imperfections which are prescribed in the several material related Eurocodes. For the future
it is worthwhile to consider these global imperfections in the Eurocode for loading, for instance in EN 1992-1-1.
Hereafter the relations between the several Eurocodes with respect to the global imperfections is discussed
based on the first generation of Eurocodes. When they are available a similar comparison will should made for
the second generation of the Eurocodes. Additional a relation between the geometrical imperfection in EN 1992-
1-1[2005] and the allowable deviations according to EN 13670 [2009] is discussed.
Steel structures
EN 1993-1-1 [2011] – Eurocode 3: Design of steel structures – Part 1-1: General rules and rules for buildings
Geometric imperfections are considered in clause 5.3. The global imperfections are described in 5.3.2 and 5.3.3
as follows:
Timber Structures
EN 1995-1-1 [2005] – Eurocode 5 – Design of timber structures – Part 1-1: General – Common rules and rules
for buildings
Imperfections are described in 5.4.4 which deals with the influence of imperfections on the internal forces in
plane frames and arches.
Masonry structures
EN 1996-1-1 [2013] – Eurocode 6 – Design of masonry structures – Part 1-1: General rules for reinforced and
unreinforced masonry structures
Global imperfections are described in paragraph 5.3
Comparison
From these summaries from the several Eurocodes, table C7.2.1-1 is derived in which the inclinations that
should be taken into account in the design and a possible reduction of this inclination related to the height of the
structure are described.
When both the inclination and the reduction for the height of the structure are considered, it can be concluded
that there is no significant difference between the several Eurocodes.
Both the EC2 and EC3 give an additional reduction that takes account for the number of elements, m. This
reduction is also identical.
Only EC2 and EC3 describe the effect of imperfections on floor diaphragm. In FprEN 1992-1-1 it is done in a
similar way.
Conclusion
From the comparison it shows that the global imperfections that influence the internal forces in the bracing
system are kind of similar in the 4 considered Eurocodes. In Eurocode 2 and Eurocode 3, additional also the
influence of imperfections on the internal forces in the diaphragms is described.
However some definitions of variables such as m and upper and lower limitations of these variables can be
different between the several Eurocodes. For that reason the definition of m is updated in FprEN 1992-1-1 to: ‘is
the number of load bearing members in one section that bear a significant part of the vertical load and that, due
to their inclination, contribute to the effect considered’ in the general description. Is is even described more
precise for the several situation to consider, such as a single columns, a bracing structure or a floor diaphragm.
In this way the definition of m is more in line with EC3, however, EC3 states that it should be about 50% of the
average load. This is more precise, but it is a discontinuous criterion and if a column for instance carries 48%, it
is questionable whether it should not be included. Besides, this percentage will change with every load
combination.
Due to the fact that structures are often not created from one material only for the ease of use it is of importance
that the global imperfections described in the Eurocodes are as identical as possible. In the revision process of
the Eurocodes this requires coordination. This coordination task can be simplified when global imperfections will
be described together with other load(effects) in Eurocode 1. l
Inclination
The inclination of the structure considered in the design, should be compared to the total height of the structure.
There it should be noted that the execution requirements are valid for each bracing element (shear wall) while in
the design the effect of the inclination can be reduced when several walls are available. Hereafter in figure
C7.2.1-5 the rules for deviations during construction and imperfections for design are compared, for the situation
where only one wall is considered.
Execution rule:
= the smaller of 50 mm/h and /(200n)
where:
n is number of stories (assumed n = h/3m)
Design rule
= h 1/200
where
h = 2/h
2/3
1
0.006
0.005
0.004
0.003
design
execution
0.002
0.001
0
0 10 20 30 40 50
total height [m]
Figure C7.2.1-5: Comparison of inclination q due to execution and design rules [n = h/(3
metres)]
Execution rule:
e = 3,75 mm + h/300 18,75 mm
Which, in figure C7.2.1-6, is compared to the initial eccentricity in the design.
Design rule:
e = h 1/400 l0
where
h = 2/h
2/3 and 1
For this comparison its assumed that the effective length l0 equals the height of the wall h.
25
eccentricity at mid height [mm]
20
15
design
10
execution
0
0 1 2 3 4 5 6
single wall height [m]
Conclusion
From figure C7.2.1-5 it can be concluded that the inclination of shear walls, considered in the design is large
enough. It even can be stated that specially for structures with a significant total height, the design rule in EN
1992-1-1 [2005] can be considered to be conservative. This among other reasons has led to the introduction of
7.2.1.1(3) in which, as an alternative method it is described that geometrical imperfections that should be taken
into account in the design also may be derived from specific execution specifications. In that situations the
geometrical imperfections to be considered should be at least 1,2 times the specified allowable deviations. This
is to account for deviations other than those that can be measured, such as due to differential shrinkage and
differential temperatures. This has also led to the lowering of the lower limit of variable n form 2/3 to 0,4.
From figure C7.2.1-6 it should be concluded that the initial eccentricities for single walls or columns considered
in the design is small in relation to the allowable eccentricities and horizontal deviations. This should be a
relevant item when the revision of EN 13670 is considered.
References
EN 1992-1-1 [2005] – Eurocode 2: Design of concrete structures – Part 1-1: General rules and rules for buildings
EN 1993-1-1 [2011] – Eurocode 3: Design of steel structures – Part 1-1: General rules and rules for buildings
EN 1995-1-1 [2005] – Eurocode 5 – Design of timber structures – Part 1-1: General – Common rules and rules for buildings
EN 1996-1-1 [2013] – Eurocode 6 – Design of masonry structures – Part 1-1: General rules for reinforced and unreinforced masonry structures
EN 13670 [2009] – Execution of concrete structures
28-03-2023
For citations, copying and referencing see Note on Front page! page 256
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (3) Background to FprEN 1992-1-1:2023
Introduction
This document has been prepared in order to validate the EC2-provisions for the rules to apply redistribution of
moments resulting from a linear elastic analysis without verification of the rotation capacity, see 7.3.2(3). Rules
for a further redistribution of moments, where verification of the rotation capacity is required, are addressed in
7.3.2(5) and are considered in the background document that has been written for that clause.
The proposal is based on the EN 1992-1-1:2004 and NEN 6720 (VBC 1995) [1] the latest Dutch code for the
design of concrete structures.
k1 + k2 kx (eq 5.10a)
where k1 and k2 are NDP’s of which the recommended value are given as:
k1 = 0,44
0,0014
k2 = 1,25 0,6
cu2
Hence:
0,44
0,44 + 1,25 kx or kx = 0,8 - 0,35
1,25
δ k x ( δ)
0.7 0.208
0.4 0.8 0.288
0.9 0.368
0.3
kx( δ) 1 0.448
0.2
0.1
0
0.7 0.75 0.8 0.85 0.9 0.95 1
δ
Figure C7.3.2(3)‐1: Relation between and kx according 5.10a
This relation is valid for concrete strengths up to 50 MPa. For higher strengths equation 5.10b should be used.
k3 + k4 kx (eq 5.10b)
where k3 and k4 are NDP’s of which the recommended value are given as:
k3 = 0,54
0,0014
k4 = k2 = 1,25 0,6
cu2
The relation between en kx for several concrete strengths is shown in figure C7.3.2(3)-2
For citations, copying and referencing see Note on Front page! page 257
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (3) Background to FprEN 1992-1-1:2023
Figure C7.3.2(3)‐2: Relation between and kx according 5.10b for fck = 50, 60 and 80 MPa
In these relations a variation of the yield strength of the reinforcement, which may have a more significant
influence on the ductility than the ultimate strain of the concrete, is not considered. Furthermore, the use of
prestressing steel is not considered. Therefore, a proposal for a change is given hereafter.
Background
General
Rotation capacity is required to allow for redistribution. Rotation capacity is created by the yielding of
reinforcement. Yielding of reinforcement is also desired in cases of the determination of the moment resistance,
even without redistribution.
Besides the considered redistribution of the effect of the considered external forces, there is also a need to
make the none considered redistribution, caused by the effects of shrinkages, deformation due to temperature
effects and differential settlement of supports, possible.
xu
d
As shown from figure C7.3.2(3)-3, yielding of the tension reinforcement will be assured when the depth of the
compression zone is limited:
xu cu2 cu2 E s
= kx =
d fyk cu2 E s fyk
cu2
Es
To ensure yielding of the reinforcement a margin can be applied to the strain that will occur in the
reinforcement before the concrete will fail. This margin is required because the considered material properties
are not the actual properties and it is those actual properties that will determine whether yielding of steel or
crushing of concrete will be the decisive criteria. Besides that the assumption that plane sections remain plane
is not always a valid assumption, hence:
For citations, copying and referencing see Note on Front page! page 258
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (3) Background to FprEN 1992-1-1:2023
cu2 E s
kx
cu2 E s fyk
This is valid for reinforcing steel. In structures in which prestressing steel or a combination of prestressing steel
and reinforcing steel is applied it is also desirable that yielding of steel occurs before the concrete will fail in
compression. Prestressing steel will start to behave non-linearly at a stress level of fpd. However in prestressing
steel already the working stress p,m is present. So non-linear behaviour will start in case of a prestressing steel
stress increase equal to fpd - p,m.
Where:
f is the steel stress increase where nonlinear behaviour occur, taken as a weighted average of the
reinforcing steel and the prestressing steel
(fpd p,m )Ap fyd As
=
Ap As
For situations without considered redistribution where cu2 equals 0,0035 and Es = 2,0ꞏ105 MPa this results in:
500
kx.max =
500 f
For fyk = 500 MPa the value of kx.max = 0,50, which is significantly greater than kx.max according equation 5.10
where kx.max = 0,45.
Beside this, a linear relation is assumed between the redistribution ratio and the allowable depth of the
compression zone:
500 f
kx.max = - (1 - ) = -
500 f 500 f
f f xu
kx.max = - = +
cu2Es cu2Es d
f f
Figure C7.3.2(3)-4: Relation between and kx according proposal for fck = 50, 60 and 80 MPa, fyk = 500 MPa
and = 1,4
For citations, copying and referencing see Note on Front page! page 259
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (3) Background to FprEN 1992-1-1:2023
1 xu
= +
E d
1 cu s
f
0,7 where Class B and Class C reinforcing steel or prestressing steel is used (see Annex C)
0,8 where Class A reinforcing steel is used (see Annex C)
where:
is the ratio of the redistributed moment to the elastic bending moment
It should be noted that in FprEN 1992-1-1:2023 the ultimate strain cu is not varied for the different concrete
grades due to which the influence of the different concrete grades is lost. However, as can be seen in figure
C7.3.2(3)-4, the influence of the different values for cu is limited.
[1] NEN, TGB 1990 Regulations for concrete - Structural requirements and calculation methods, NEN
References
6720, Delft 1995.
.
For citations, copying and referencing see Note on Front page! page 260
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
For citations, copying and referencing see Note on Front page! page 261
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
Introduction
This document provides the background to paragraph 7.3.2 (5) (Rotation Capacity).
Problem statement
Significant rotation capacity and thereby structural ductility can be achieved by providing the structure with
ductile sections. In structures that can be assimilated to continuous beams, plastic rotation occurs mostly in
the zone around the negative bending moment. The area from the support to the point of zero moment can
be isolated and assimilated to a simply supported span subjected to a point load (the reaction) and the
distributed loads which have the opposite direction. The problem can further be simplified (and tested) by
disregarding the effect of the distributed loads. In this way the problem can be modelled as a simply
supported beam with a span equal to the distance between the points of contraflexure and a single
concentrated load where the slenderness corresponds to roughly half of the span of the continuous beam
subjected to a point load (see Figure C7.3.2(5)-2).
For citations, copying and referencing see Note on Front page! page 262
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
The plastic rotation capacity is the rotation due to the excess curvature with respect to the curvature
corresponding to the yielding moment that occurs between the section at which the yielding moment (Myd)
is reached and the section at which the ultimate moment (MRd) is reached (see Figure C7.3.2(5)-2). Due to
shear cracking, it can be assumed that the moment curve is shifted horizontally by a given magnitude. In
Figure C7.3.2(5)-2, the magnitude of the shift is assumed to be one effective height (d). Reality, of course
can be more complex.
The bottom part of Figure C7.3.2(5)-2 shows a graphical representation of the plastic rotation.
Figure C7.3.2(5)‐2 Top: Bending moment (shifted due to shear cracking near the point of application
of the load) and Bottom: curvatures and graphical representation of plastic rotation
Note that in Figure C7.3.2(5)-2 the curvatures are mean curvatures, since curvatures will, have maximum
values at the cracks and will reduce between cracks due to tension stiffening effects. This will result in a
reduction of the rotation capacity with respect to the assumption that all sections are fully cracked. So,
tension stiffening is an unfavourable effect when dealing with rotation capacity. A more detailed
representation of the curvature law is given in Figure C7.3.2(5)-3.
For citations, copying and referencing see Note on Front page! page 263
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
Figure C7.3.2(5)‐3 A more detailed representation of the distribution of curvatures in the plastic
rotation zone, taken from reference[C07-6] (Eligehausen, 1987)
1 1
l r u , m r y , m lcf 1 1 1
pl , simpl 2 cf aMy 2 aMy 2d
2 2 2 r y,m r u , m r y , m
1 1
(1)
lcf r u , m r y , m 1 1
2 aMy 2d
2 2 r r y , m
u ,m
lcf 1 1
aMy 2d
2 r u , m r y , m
For citations, copying and referencing see Note on Front page! page 264
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
lcf
M Rd PRd
4 M yd 2aMy a M yd lcf
My (2)
PRd M lcf M Rd 2
M yd aMy Rd
2
Myd and MRd can be determined assuming linear behaviour of concrete in compression for Myd and a
rectangular stress block for MRd, as shown in Figure C7.3.2(5)-5, x should be larger than xu, since in most
cases the increase in strain at the reinforcement will be much larger than the increase of the strain at the
compression fibre. At the limit they will be the same for a section that reaches, at the same time, yielding
and maximum concrete strain. Of course, it is possible that when the yielding moment is reached the
behaviour of concrete is not linear anymore. This will occur for sections with large steel quantities. In these
cases, however, the rotation capacity will be small, and plastic analysis would not be recommended. For
this reason, the assumption is maintained in the derivation that follows, without further limitation than that
the rotation capacity cannot be negative.
For citations, copying and referencing see Note on Front page! page 265
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
With the assumptions described in Figure C7.3.2(5)-5, the yielding and the ultimate bending moments can
be derived, and from these expressions, aMy can be determined as shown in Eq. (3):
1 x 1 x
M yd As f yd d 1 1
3d f yd 3 d lcf
aMy (3)
x f s , ef xu 2
M ud As f s , ef d 1 0.4 u 1 0.4 d
d
The mean curvature at yielding (1/r)y,m can be derived from Eq. (4)
1 yd
TS My
r y,m x
d 1
d (4)
M
TS My 1 0.6 cr 0.4
M
y
In Eq. (4) TSMy is the tension stiffening factor for SLS and Mcr is the cracking moment. For tension stiffening
the upper bound is taken (factor of 0.6) because this is the conservative assumption.
The mean curvature at ULS (1/r)u,m can be derived from Eq. (5):
1 ud
TSMu min ; cu
. (5)
r u , m d xu xu
The two limits of the above expression depend on whether the failure of the section is due to failure of steel
or failure of concrete respectively. In this case the effect of tension stiffening is very high, because when the
second branch of the constitutive law of steel is reached a small increase in steel stress will produce very
large strains. This means that the strain at the crack can be very large, while the strain between cracks can
be quite small. To model Tension Stiffening effects, the Tension Chord Model (Marti, 1998) [C07-7] will be
adopted. This model is relatively simple. It considers a constant bond stress whose value depends on the
concrete strength and on whether the steel has yielded or not (see Eq. (6)).
2
0.30 f 3
cm if s f yd
b 2
(6)
0.60 f 3 if s f yd
cm
The tension stiffening factor TSMu represents, approximately, the ratio between the mean strain in steel
between the section of the crack and the section between cracks, and the maximum strain at the crack (see
Eq. (7)).
For citations, copying and referencing see Note on Front page! page 266
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
srm
y dy
s
(7)
TS Mu 0
ud , ef
For the determination of TSMu, two cases can be distinguished, as shown in Figure C7.3.2(5)-7, where s1 is
the distance between the crack and the point where fyd may be found as the stress in the rebar. The first
case is where srm/2 is smaller than s1 and second case is the oposite. Hereafter s1 is described as a function
of srm and . Where is the ratio between the difference in the force in the bar between the ultimate force
and the yielding force end the force that can be transmitted by bond between the section of the crack and
the section between cracck with the lowr bond stress.
srm
s1
2
( f s , ef f yd ) AØ 2 ( f s ,ef f yd )Ø
Ø 0,3 f cm 2/3
srm 0, 6 srm f cm 2/3
:
- In the first case where the difference between the ultimate force and the yielding force of the bar
is larger than the force that can be transmitted by bond between the section of the crack and the
section between cracks with the lower bond stress (>1.00), then the variation in force will be
linear and the minimum and maximum strains and TSMu factor can be determined as shown in Eq.
(8):
1
min su ,ef yd
ud , ef
ud ,ef min
mean (8)
2
mean 1 yd
TS Mu 1 1
ud ,ef 2 ud ,ef
- If, on the contrary, the difference between the ultimate force and the yielding force in the bar is
smaller than the force that can be transmitted by bond between the section of the crack and the
section between cracks with the lower bond stress (<1.00), then the variation in force will be bi-
linear and the minimum and maximum strains and TSMu factor can be determined as follows.
First, the force in the bar at the section in the middle between two cracks can be determined from
Eq. (9):
2 2
s
Fmin A f s , ef 0.3 f cm3 s1 0.6 f cm3 rm ,cal s1
2
A f f s s ,ef yd
2
srm , cal
A f s ,ef A f s ,ef f yd 0.6 f cm3 1
2 (9)
2
A f s ,ef A f s ,ef f yd 1 1
2
A f s ,ef A f s ,ef f yd 1
Since the bond stress that can be transferred between the section of the crack and the section
between cracks is larger than the maximum force in the bar minus the force at yielding, the stress
at the section between cracks will be within the elastic range and the minimum strain can therefore
be derived as shown in Eq. (10):
For citations, copying and referencing see Note on Front page! page 267
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
yd min min A
f yd Fmin , 2
A f s ,ef A f s , ef f yd 1
A
2
f s ,ef f s , ef f yd 1
min yd
f yd (10)
2 2
f s ,ef 2 f yd 1
yd
f yd
f s ,ef 1 2
yd 2 1 1
f
yd
ud , ef yd s1 yd min s1
s , mean 1
2 srm 2 s rm
2 2
ud ,ef yd yd min 1 1
1 ud ,ef yd min
2 2 2 2 2
1 f s , ef 1 2 1
ud ,ef yd yd 2 1 1
2 2 f 2
yd
1 f s , ef 1 1 1 1
ud ,ef yd yd 2 1 1
2 2 f yd 2
1 f s ,ef 1 1 1
ud ,ef yd 2 1 1 (11)
2 2 f yd 2
1 1 f s , ef 1 1
ud ,ef yd 1 1
2 2
2 f yd 2
1
1
2 2
1 f s , ef 1 1
ud ,ef yd 2
2 2 f yd 2
f s ,ef 1
ud ,ef yd 1 1 2
2
2 f yd
The tension stiffening factor referred to the steel strain can then be obtained from Eq. (12):
mean yd f s , ef 1
TS Mu , s 1 1 2 (12)
ud ,ef 2 ud ,ef 2
f yd
For citations, copying and referencing see Note on Front page! page 268
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
Figure C7.3.2(5)‐7 Distribution of strains between cracks in the zone of maximum strain and
correspondence with the stress-strain diagram of steel, force variation in the bar and bond stresses
As a simplification it is assumed that the tension stiffening factor for strains can be assimilated to the tension
stiffening factor relating to curvatures. This is because it is not easy to determine how the compressive strain
will vary from the section of the crack to the section between cracks, since there may be slip at the section
between cracks. This proposal is conservative since it maximizes the tension stiffening factor and is also
not very unrealistic since generally the strain in the reinforcement will be much larger than the strain in
concrete (see Eq. (13)).
d
By replacing the above formulations into Eq. (1), the final formulation is derived as follows:
1 yd
pl , simpl k1, pl lcf 2d TS My (14)
u , m
r d x
where:
1 x
1 f yd 1
k1, pl 1 3d
2 f s , ef xu
1 0.4 d
M
TSMy 1 0.6 cr 0.4
M
y
1 ud
TS Mu min ; cu .
r u , m d xu xu
For citations, copying and referencing see Note on Front page! page 269
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
1 yd
1 1 if 1.00
2 ud , ef
TS Mu
yd f s , ef 1
2 1 1 2 if 1.00
ud , ef 2 f yd
A f s ,ef f yd f s , ef f yd
2
2
srm ,cal
f 3
cm 2 f cm3 srm ,cal
2
pl , simpl
k k l 2kd d 1
y 1, pl cf yd
TS My
u , m
r d x
(15)
1 ud
TS Mu min ; cu
r u , m d x x
u u
where:
ky is a shape factor for the variation of mean curvature between the section of yielding to the section of
rupture;
kd is a factor to quantify the shift of the bending moment law due to diagonal shear cracking;
is a factor to account for the experimentally demonstrated fact that the 0.35% limit to concrete strain
is an extremely conservative value for most design situations, as it will be shown below. This factor
is key to explain the size effect observed in rotation capacity tests ([C07-5], [C07-8], [C07-9], [C07-
11], [C07-12]);
is a model uncertainty factor to guarantee an adequate level of reliability of the formulation.
Table C7.3.2(5)-1 provides a general overview of the range of variables which are encompassed by these
tests. Note that the reinforcement ranges, especially for the Corley tests are very high, which would provide
high strains in the compressive fibres. Also, note that the value of the ratio between the ultimate steel stress
and the yield stress is very high for European Standards, as factor k is limited to a maximum of 1.35 for
class C steel, according to EN 1992-1-1:2004, and, also, to FprEN 1992-1-1:2023.
Table C7.3.2(5)‐1 – Range of parameters covered in the tests by Mattock and Corley
# tests Range Range Range Range Range
d [mm] xu/d k w
Figure C7.3.2(5)-8 and -9 show the results pertaining to the maximum measured compression strain in the
tests of Mattock and Corley as a function of the effective depth. The tests are 3-point bending tests. The
compressive strain was measured using a Linear Differential Transformer Gage located at mid-span. The
measuring length of the gage is not reported, but from the picture included in the publication it is estimated
to be at least 20 cm long. It is certain that measuring the compressive strain is not an easy task and that,
for instance, the tests by Bigaj [C07-9], show a significant variation of the compressive strain along the
For citations, copying and referencing see Note on Front page! page 270
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
compressed fibre. Nonetheless, the measured values are so much above the 0.35% limit that it must be
conceded that this limit is just a very conservative lower bound. In fact, the conclusion by Mattock is as
follows:
Figure C7.3.2(5)-8 and -9 also clearly show that the maximum compressive strain is a function of the
effective depth and that this function has a hyperbolic shape.
0.060
0.050
cu,exp
0.040
0.030
0.020
0.0035
0.010
0.000
0 100 200 300 400 500 600
d[mm]
0.060
0.050
cu,exp
0.040
0.030
0.020
0.0035
0.010
0.000
0 100 200 300 400 500 600 700 800 900
d[mm]
For citations, copying and referencing see Note on Front page! page 271
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
d ref
cu cu , d , k ,0 k ,1 k ,2 w cu , max (16)
w
d
0.060
0.050
cu,exp
0.040
y = 0.6657x + 0.0112
0.030 R² = 0.0452
0.020
0.0035
0.010
0.000
0.000 0.005 0.010 0.015 0.020 0.025
w
0.070
0.060
0.040
0.030
0.020
0.0035
0.010
0.000
0.000 0.005 0.010 0.015 0.020 0.025 0.030 0.035 0.040
w
For citations, copying and referencing see Note on Front page! page 272
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
Table C7.3.2(5)‐2 Range of parameters covered in the tests by Bigaj, Bosco et al. and Calvi et al.
# tests Range Range Range Range Range
d [mm] xu/d k w
The resulting calibration parameters are shown in Eq. (17). ky can the taken as nil because the values of
factor k1,pl are very small and the effect of the first term of the initial parenthesis on the rotation capacity is
negligible. This simplifies the model.
ky 0
kd 0.65
d ref
cu , d , 0.002 0.0015 3 w cu ,max (17)
w
d
d ref 900 mm
cu ,max 0.015
With these factors, the formulation to estimate the rotation capacity can be written as shown in Eq. (18).
1.3d ud cu , d , w yd
pl , simpl TS Mu min ; TS My
d xu
xu d x
(18)
1.35
cu , d , 0.002 3 w 0.015
w
d mm
Figure C7.3.2(5)-12 to -15 show that the calibrated model can reproduce the effects of the main variables
quite well.
In the tests by Bigaj et al. the size effect and the effect of reinforcement ratio (expressed here in terms of
xu/d) are clearly captured. For the 90 mm beam the effect of reinforcement ratio and the effect of size are
offset by the low ultimate strain of the steel used.
For citations, copying and referencing see Note on Front page! page 273
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
Tests by Bigaj
80
60 d=180 mm
50
[mrad] 40 d=450 mm
d=90 mm
30
20
10 su=3.6%
Others su9.2‐9.4%
0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35
xu/d
Exp. Th.
Figure C7.3.2(5)‐12 Comparison of the test results and predictions obtained from the calibrated model
for the tests of Bigaj and Walraven
Exp. h=600 mm
60
50
Exp. h=600 mm
[mrad]
40 Exp. h=600 mm
30
20 Exp. h=400 mm
10 Exp. h=200 mm
0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45
xu/d
Exp. h=200 mm Exp. h=400 mm Exp. h=600 mm Theoretical
Figure C7.3.2(5)‐13 Comparison of the test results and predictions obtained from the calibrated model
for the tests of Bosco et al. for steel type B500H
Referring to Figure C7.3.2(5)-13 and to the elements with 600 mm of effective height, the effect of xu/d on
the rotation capacity is clearly captured. For very low reinforcement ratios, the rotation capacity increases
with the increase of , up to a maximum when the capacity starts decreasing. For xu/d equal to approximately
0.27 and 0.40, the size effect is clear with the 400 mm depth providing a larger rotation capacity than the
600 mm effective depth and the 200 mm effective depth providing a larger rotation capacity than the 400
mm depth. The same analysis can be done for Figure C7.3.2(5)-14.
For citations, copying and referencing see Note on Front page! page 274
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
Exp. h=200 mm
60
50
[mrad]
40
Exp. h=400 mm
30 Exp. h=400 mm
Exp. h=200 mm
Exp. h=400 mm Exp. h=600 mm
20
10 h=600 mm
Exp.
Exp. h=600 mm Exp. h=400 mm
0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45
xu/d
Exp. h=200 mm Exp. h=400 mm Exp. h=600 mm Theoretical
Figure C7.3.2(5)‐14 Comparison of the test results and predictions obtained from the calibrated model
for the tests of Bosco et al. for steel type B500N
Tests by Calvi et al. (Only high bond tests and single load tests)
140
Higher reinforcement ratios lead to smaller
crack spacing and smaller TS effects. This is why
120 the curves show a growing tendency with xu/d su=7.99%
for small reinforcement ratios
100
[mrad]
80
60
su=3.45%
40
20
0
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14
xu/d
Exp. Theoretical
Figure C7.3.2(5)‐15. Comparison of the test results and predictions obtained from the calibrated
model for the tests of Calvi et al
The tests by Calvi et al, shown in Figure C7.3.2(5)-15, are centred on the effect of the ultimate strain of steel
for elements with very low reinforcement ratio. The model correctly predicts a higher rotation capacity for
members built with the steel with the larger ultimate strain. Because the reinforcement ratios are very low,
the rotation capacity increases with increasing xu/d.
Figure C7.3.2(5)-16 shows a comparison between experimental and estimated rotation capacity values. The
slope of the correlation line is close to 1.00, if slightly on the conservative side.
Tests by Mattock and Corley have not been used for the calibration of the model because they show a much
larger scatter. Nonetheless, these results fall mainly on the conservative side, especially regarding the tests
by Mattock (see Figure C7.3.2(5)-17 and -18). This is most likely due to the limitation of the maximum
compressive strain.
For citations, copying and referencing see Note on Front page! page 275
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
Predicted vs Experimental
(Bigaj, Bosco, Calvi)
y = 1.1044x
R² = 0.9568
140
120
100
exp [mrad]
80
60
40
20
0
0 20 40 60 80 100 120 140
model [mrad]
Figure C7.3.2(5)‐16 Comparison between measured and estimated rotation capacity for the tests of
the calibration set.
6.00
5.00
exp /model
4.00
3.00
2.00
1.00
0.00
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
xu/d
Figure C7.3.2(5)‐17 Ratio of experimental over estimated rotation capacity as a function of xu/d for
the tests of Mattock
For citations, copying and referencing see Note on Front page! page 276
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
18
16
14
exp /model
12
10
0
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
xu/d
Figure C7.3.2(5)‐18 Ratio of experimental over estimated rotation capacity as a function of xu/d for
the tests of Corley
e0.803.8CoV
3.0 (19)
Table C7.3.2(5)‐3– Statistics of the ratio between experimental and estimated values
expth
1.058
= 40.3%
CoV = 38.1%
= 3.01
Using this value, the formulation has been compared with all the experimental results in Figure C7.3.2(5)-
19. Only three estimated values are unconservative (exp/model=0.58, 0.76 and 0.96). Of course, these
values would increase once characteristic values of material properties are considered, and material partial
factors are applied. This is an indication that the value of 3,0 for the model uncertainty factor is acceptable.
For citations, copying and referencing see Note on Front page! page 277
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
18
16
14
exp /model
12
10
0
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
xu/d
Figure C7.3.2(5)‐19 Comparison of experimental results and the estimated value of the rotation capacity
divided by the model uncertainty factor
To illustrate the method two application examples are presented below, one in which failure occurs by
rupture of the steel, and one in which failure occurs by crushing of concrete.
62 1
w 2 1.414 103
4 200 200 (20)
1.35
cu , d , 0.002 3 1.414 103 0.00995
w
365
For ULS analysis a rectangular stress block may be assumed. Whether the sectional failure is governed by
steel or concrete can be determined by looking at the force in concrete and steel for the plane representing
simultaneous failure for both materials. If for this plane the force in concrete is larger than the force in steel,
then failure will be by steel, otherwise the failure will be by concrete.
cu , d , 9.95
xu ,lim w
d 0.365 0.070 m
cu , d , su
w
9.95 41.8
(21)
Fc,lim 0.8 f cm bxu ,lim 0.8 30900 0.2 0.07 346.08 kN
Fsu As f su 4 1.13 64.1 289.73 Fc,lim
Failure occurs by steel, so the depth of the neutral axis can be determined from equilibrium of axial forces:
For citations, copying and referencing see Note on Front page! page 278
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
ud ,ef ud
f su ,ef f su
(22)
Fsu 289.73
xu 0.058 m
0.8 f cm b 0.8 30900 0.2
The tension stiffening factor for ULS can then be determined by applying Eq. (14):
f s , ef f yd A
641.0 595.60 103 1.13 104
5.13
1.047 1
srm ,cal 2
0.088 2
4.90
0.3 f 3 0.012 0.3 30.9 103
3
2
cm
2 (23)
1 yd 1 2.98
TS M u 1 1 1 1 0.556
2 ud , ef 2 1.047 41.8
and, likewise, the mean ultimate curvature can be determined as follows:
1 su 0.0418
TS M u 0.556 0.0757 m -1 (24)
r u , m d xu 0.365 0.058
The depth of the neutral axis for the yielding moment is determined using the classical State II equations:
200000 200000
e 1
6.71
29811
9500 30.9 3
4 1.13
6.19 103 (25)
20 36.5
x 2 2
e 1 1 3
6.71 6.19 10 1 1 0.250
d e 6.71 6.19 10 3
The tension stiffening factor for the yield moment can then be calculated as:
f ctm bh 2
M cr
TS M y 1 0.6 1 0.6 6
M 1 x
yd
2 Ecm d x x b d 3
yd 2
2970 0.20 0.402
(26)
1 0.6 6
0.25 0.365
0.365
1 3
29811 10 0.00298 0.25 0.365 0.2
3 2
2 0.365 0.25 0.365
15.84
1 0.6 0.895
90.408
Finally, the rotation capacity follows from:
1.3d 1 yd
Rd TS My
model r u , m d x
1 0.895 0.00298 (27)
1.3 0.365 0.0757 0.0104 rad
3.00 0.365 0.25 0.365
0.0066
0.0313
For citations, copying and referencing see Note on Front page! page 279
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
600 mm and reinforced with 912 mm bars. The effective height is 565 mm. The calculated mean crack
spacing is 84 mm. The reinforcement is Class B500H with the following experimental characteristics:
- yield strength of 587.3 MPa
- strain at yielding: 0.00294
- maximum tensile strength of 672.2 MPa
- strain at maximum strength: 0.07
The transversal reinforcement is formed by 6 mm stirrups spaced at 150 mm, the reported concrete
compressive strength is 30.9 MPa and the reported tensile strength, 2.97 MPa.
The estimated maximum compressive strain is determined from Eq. (28):
62 1
w 2 1.257 103
4 150 300 (28)
1.35
cu , d , 0.002 3 1.257 103 0.00816
w
565
For ULS analysis a rectangular stress block may be assumed. Whether the sectional failure is governed by
steel or concrete can be determined by looking at the force in concrete and steel for the plane representing
simultaneous failure for both materials. If for this plane the force in concrete is larger than the force in steel,
then failure will be by steel, otherwise the failure will be by concrete.
cu , d , 8.16
xu ,lim w
d 0.565 0.059 m
cu , d , su
w
8.16 70
(29)
Fc,lim 0.8 f cm bxu ,lim 0.8 30900 0.3 0.059 437.54 kN
Fsu As f su 9 1.13 67.22 683.63 Fc,lim
Failure occurs by concrete in this case. Due to steel hardening, the depth of the neutral axis has to be
determined iteratively by modifying its value until equilibrium of axial forces is fulfilled:
Fc 0.8 f cm bxu
cu , d ,
ud ,ef w
d xu
xu
(30)
f su f y
f su , ef f y
ud y
ud , ef y
Fs As f su ,ef Fc
The tension stiffening factor for ULS can then be determined from:
f s , ef f yd As
639.34 587.3 103 1.13 104
5.880
1.257 1
2 2
srm 0.084 4.677
0.3 f cm
3
0.012 0.3 30.9 10
3 3
2 2 (32)
1 yd 1 2.94
TS M u 1 1 1 1 0.619
2 ud ,ef 2 1.257 70
For citations, copying and referencing see Note on Front page! page 280
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
1 cu 0.00816
TS M u 0.62 0.0576 m -1
r u , m xu 0.0878
The depth of the neutral axis for the yielding moment is determined using the classical State II equations:
200000 200000
e 1
6.71
29811
9500 30.9 3
9 1.13
6.00 103 (34)
30 56.5
x 2 2
e 1 1 3
6.71 6.00 10 1 1 0.246
3
d e 6.71 6.00 10
The tension stiffening factor for the yield moment can then be calculated as:
f ctm bh 2
M
TS M y 1 0.6 cr 1 0.6 6
M 1 x
yd
2 Ecm d x x b d 3
yd 2
2970 0.30 0.602 (35)
1 0.6 6
0.246
1
1 3
29811 10 0.00294 0.246 0.565 0.3
3 2
2 1 0.246
53.76
1 0.6 0.895
309.21
(36)
Finally, the rotation capacity follows from:
1.3d 1 yd
Rd TS My
model r u , m d x
1
(1)
0.895 0.00294
1.3 0.565 0.0576 0.0126 rad
3.00 0.246 0.565
0.565
0.735
0.0514
0.0377
For citations, copying and referencing see Note on Front page! page 281
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.3.2 (5) Background to FprEN 1992-1-1:2023
References [C07-5] Bosco, C., Debernardi, P.G. (1993). Influence of some basic parameters on the plastic rotation
of reinforced concrete elements. En R. E. (Ed.), Ductility - Reinforcement. Progress Report of
Task Group 2.2 'Ductility Requirements for Structural Concrete - Reinforcement'. Bulletin
d'Information nº 218 (págs. 25-44). Lausanne: CEB.
[C07-6] Eligehausen, R. L. (1987). Rotation capacity of plastic hinges and allowable degree of moment
redistribution. Struttgart: University of Stuttgarrt.
[C07-7] Marti, P. Á. (1998). Tension Chord Model for Structural Concrete. Structural Engineering
International,, 8(4), 287-298. doi:10.2749/10168669878048887
[C07-8] Bigaj, A., Walraven, J. (2002). Size Effects in Plastic Hinges of Reinforced Concrete Members.
Heron 47(1):53 – 75.
[C07-9] Bigaj, A. (1999).Structural Dependence of Rotational Capacity of Plastic Hinges in RC Beams
and Slabs. PhD Thesis. TNO.Delft University Press. 244 p.
[C07-10] Calvi, J.M., Cantù, E., Macchi, G., Magenes, G., (1993). Rotation capacity of reinforced
concrete slabs as a function of steel properties. Ductility - Reinforcement. Progress Report of
Task Group 2.2 'Ductility Requirements for Structural Concrete - Reinforcement'. Bulletin
d'Information nº 218 (págs. 25-44). Lausanne: CEB.
[C07-11] Mattock, A.H (1965). Rotational capacity of hinging regions in reinforced concrete beams.
Proceedings of the International Symposium, Flexural Mechanics of reinforced concrete.
American Society of Civil Engineers. P 143-181.
[C07-12] Corley, W.G. (1966). Rotational capacity of reinforced concrete beams. Journal of the
Structural Division. Proceedings of the American Society of Civil Engineers. Paper 4939, Vol.
92, ST5, p. 121-146.
For citations, copying and referencing see Note on Front page! page 282
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
28.03.2023
For citations, copying and referencing see Note on Front page! page 283
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.4.3 Background to FprEN 1992-1-1:2023
Introduction
This document provides background information to paragraph 7.4.3.2 Simplified methods based on nominal
curvature and second order linear analysis and annex O Simplified approaches for second order effects.
For the second order linear analysis an effective stiffness should be chosen. This stiffness can conservatively
7.4.3.2
be derived from the point where yielding in the reinforcement will occur. For an isolated member this point
Simplified methods based can be derived from moment-curvature diagram in a governing section of the member. When more global
on nominal curvature effects should be considered in a case where plastic behaviour can occur in different sections in the structure,
it is established that the effective stiffness should be derived from the displacement that is found prior to the
occurrence of the last plastic hinge. These principles are shown in figure C7.4.3.2-1.
Figure C7.4.3.2‐1: Determination of effective stiffness from moment-curvature and force displacement
relations
It is allowed to use less conservative methods to determine the effective stiffness. However, in that case,
adequate justification of the stiffness used is required.
For citations, copying and referencing see Note on Front page! page 284
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
7.4.3 Background to FprEN 1992-1-1:2023
1. Walraven, J., Westerberg, B. (2008), e.o. Eurocode 2 – Commentary. European Concrete Platform
References
ASBL.
2. Hellesland, J. (2012). Evaluation of effective length formulas and applications in system instability
analysis. Engineering Structures, 45(12). 405-420.
3. Hellesland, J. Challamel, N. , Casandjian, C., Lanos, C. (2013) Reinforced Concrete beams, column
and frames – Section and slender member analysis, ISTE Ltd, London, and John Wiley & Sons,
Hoboken, N.J.
4. Westerberg, B. (2004) Second order effects in slender concrete structures – Background to the rules in
EC2, TRITA-BKN Report 77, Betongbyggnad, KTH, Stockholm
For citations, copying and referencing see Note on Front page! page 285
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
FprEN 1992-1-1:2023
For citations, copying and referencing see Note on Front page! page 286
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.1.4 Background to FprEN 1992-1-1:2023
Introduction
This document contains the explanation of the changes in the provisions related to strength increase of concrete
associated to transverse compressive stresses as well as the new provisions for designing confinement
reinforcement.
Conclusions
‐ The failure criterion should account for the composition of the mixtures, namely in terms of maximum
aggregate size.
‐ The failure criterion should be based on tests where the pore pressure was relieved.
‐ The current EN 1992-1-1:2004 failure criterion is not safe for mortar mixtures at higher transverse
pressures (test results (a)).
The proposed failure criterion for FprEN 1992-1-1:2023, which is derived from fib Model Code 2010, but with
some modifications to account for the effect of aggregate size, agrees reasonably well with test results and
captures the effect of maximum aggregate size.
Test results
In the following comparison of test results with failure criteria, test results from the literature have been selected
based on the following criteria:
‐ Test conducted on cylinders (disregarding hollow cylinders and cubes) subjected to active transverse
pressure (e.g. oil pressure).
‐ Cylinders with height to diameter ratio of approximately 2.
‐ Test results where the composition of the mixture is known.
The following aspects have not been directly addressed in the document (the parameters are however
evaluated in the comparison):
‐ Differences in the compressive strength of the mixtures.
‐ Dependency of specimen size (Li and Ansari [4] concluded that a slight size effect on failure strength
for HSC exists).
2
60
(b) ddg 16 Dmax 40 mm for normal weight concrete with fck > 60 MPa
f
ck
Figure C8.1.2: Comparison to test results - 4.75 mm < Dmax < 16 mm, All tests conducted with Hoek Cell
Setup
o – Specimens of size 55x110 mm (Hoek Cell)
Figure C8.1.3: Comparison to test results - 4.75 mm < Dmax < 16 mm, All tests conducted with Test setup
with pore pressure outlet
x - Specimens of size 150x300 mm
* - Specimens of size 100x200 mm
◊ - Specimens of size 76x152 mm
Figure C8.1.4: Comparison to test results from literature, mortar with Dmax ≤ 4.75 mm
x - Specimens of size 150x300 mm
* - Specimens of size 100x200 mm
The provisions in FprEN 1992-1-1:2023 are adapted from fib MC2010 and are consistent with the confinement
effectiveness factors in EN 1998-1:2004 (see background in [20]). These provisions are based on the following
assumptions:
(i) A stress in the confinement reinforcement corresponding to its yield strength fyd can be activated at
ULS.
(ii) Only the concrete area enclosed within the confinement reinforcement (“confinement core” defined by
the centreline of the reinforcement and approximated by replacing the parts of circles by corners) is
affected by the beneficial effect of confinement. The concrete cover outside the confinement core is
not affected by confinement, but it may be accounted for with its strength fcd .
(iii) The effect on the concrete core of circular confinement reinforcement (single circular hoops or
continuous spiral reinforcement) can be assumed as uniformly distributed, so that the full confinement
core may be considered (see Figure C8.1.5a which shows the cross section at the level of the
confinement reinforcement). In case of stirrups with square, rectangular or polygonal shapes (and
additional single-leg links), the transfer of forces between the confinement reinforcement and the
confined concrete occurs mainly at the bends (and at the anchorages of single-leg links). Along
straight segments between bends (and anchorages), the concrete within the confinement core
defined above is not fully confined. As a simplification, parabolic limits with 45° angles between fully
confined and non-confined concrete can be assumed as shown in Figures C8.1.5b and c, so that the
area to be deduced from the confinement core is:
1 2
bi (C8.1.1)
6
where bi are the distances between bends (defined by their related corners) or anchorages of the
confinement reinforcement, see b1 and b2 in Figure C8.1.5c.
(iv) As shown in Figure C8.1.6, between confinement reinforcements, the outer regions of the
confinement core are less affected by the confinement effect than at the level of the confinement
reinforcements. This effect can be accounted for in a simplified manner by assuming again a
parabolic limit between fully confined core and the non-confined concrete with 45° slopes (see limits
between grey and white areas in Figure C8.1.6).
In prEN 1992-1-1:2023, the assumption (i) is considered in Formulae (8.11-14) whereas the effects (ii) and (iii)
are covered with coefficient kconf,b and effect (iv) with coefficient kconf,s in Formula (8.15):
where
fcd,c is the design value of the concrete compressive strength accounting for confinement which may be
assumed to be constant in the whole compression zone (coefficient kconf,b allows smearing the effect
of confinement)
fcd is the compressive strength increase due to triaxial action described in 8.1.4(2)
kconf,b and kconf,s are the coefficient described above
Figure C8.1.5: Influence of the shape of the confinement on the effectively confined concrete area for (a)
circular confinement reinforcements (b) squares and rectangular with simple confinement
reinforcement and (c) with additional confinement reinforcement fixing the straight segments
of the main confinement reinforcement (case with an intermediate one-leg link)
Figure C8.1.6: Influence of the spacing s of the confinement on the effectively confined concrete area for (a)
fully compressed columns and (b) compression zones
With respect to the previous assumptions and the resulting model, the following comments can be made:
(e) The assumption (i) has been validated in [21]. As shown in Figure C8.1.7c, the comparison with tests
results shows fine agreement also for high yield strengths of the confinement reinforcement (the limit
of applicability of FprEN 1992-1-1:2023 is fyd = 610 MPa).
(f) Despite the fact that in some cases, spalling of the reinforcement cover can be observed before the
maximum resistance is attained (see for instance Figure 3h in [21]), the assumption (ii) is justified.
The comparison with the database described in [21] shows again fine agreement (see Figure C8.1.7b,
where good results are obtained also for small values of ratio Acs/Ag which refer to small columns with
relatively thick reinforcement covers). The simplification (ii) is also made possible with the introduction
of the strength reduction factor cc in 5.1.6 (see related background document and more detailed
explanation in [21]).
(g) The approximation with respect to the confinement core area in (ii) (rounded parts of the perimeter at
bends of the confinement reinforcement considered as corners) is not on the safe side, but it is
compensated by the approximation with respect to the definition of bi in (iii) (bi measured up to the
nominal corners as shown in Figure C8.1.5c) which is on the safe side. Here again, the comparison to
tests described in [21] (see Figure C8.1.7) shows that these simplifications are reasonable (similar
results for circular, square and rectangular columns in terms of experimental-to-calculated resistance
ratio).
(h) As shown in Figure C8.1.8a, the coefficient kconf,b , which accounts for the effects (ii) and (iii), strongly
depends on:
• The shape of the confinement reinforcement (circular reinforcements being significantly
more efficient than simple square or rectangular ones)
• The ratio bcs / b , which is associated to the thickness of the reinforcement cover and the
size of the column (in small columns, the reinforcement cover plays a major role)
(i) As depicted in Figure C8.1.8b, the coefficient kconf,s in Formula (8.8), which accounts for effect (iv),
shows a significant influence of the ratio s / bcs (spacing of confinement reinforcement / width of the
confinement core).
Figure C8.1.7: Comparison between tests and model for confinement reinforcement in FprEN 1992-1-
1:2023, influence of experimental-to-calculated ratio as a function of (a) the ratio of the
confinement reinforcement; (b) the ratio between the confinement core area Acs and the total
column area Ag, and (c) the yield strength of the confinement reinforcement (results of 264
centrically loaded columns, adapted from [21])
a)
1.20
large columns
1.10
1.00 small non‐exposed
0.90 columns
0.80
kconf,b
0.70 small exposed columns
0.60
0.50
0.40
0.30
0.20
0.10
0.00
0.40 0.50 0.60 0.70 0.80 0.90 1.00 1.10 1.20 1.30 1.40
bcs / b
b)
1.20
1.10
1.00
0.90
0.80
kconf,s
0.70
0.60
0.50
0.40
0.30
0.20
0.10
0.00
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00 1.10
s / bcs
Figure C8.1.8: Influences of the main geometrical parameters of circular and square columns on (a)
coefficient kconf,b and (b) coefficient kconf,s
The implication of the described model for practical cases are shown in Figure C8.1.9 for a square column
under axial force and bending. To allow for comparisons, the longitudinal reinforcement is kept constant.
Following cases are considered:
(a) fck = 40 MPa, simple confinement renforcement, parabola-rectangle
stress distribution 2
1 bcs
kconf ,b
(b) fck = 80 MPa, simple confinement renforcement, parabola-rectangle 3 b
stress distribution
(c) fck = 40 MPa, multiple confinement renforcement , parabola-rectangle
stress distribution 2
7 bcs
kconf ,b
(d) fck = 80 MPa, multiple confinement renforcement , parabola-rectangle 9 b
stress distribution
(e) – (h) same parameters as (a) – (d), but with stress-block distribution
(a) 0.600
0.40 x 0.40 m2,
Parabola-rectangle fck = 40 MPa,
0.500 4x3 20 B500
0.400
MRd [MNm]
0.300
FprEN 1992-1-1:2023
0.200 phi 8 @300
phi 8 @150
phi 10 @150
0.100 phi 12 @150
phi 14 @150
cc = 0.85
EN 1992-1-1:2004 a_cc0.85
cc = 1.00
EN 1992-1-1:2004 a1.00
0.000
0 -1 -2 -3 -4 -5 -6 -7 -8 -9 -10
NRd [MN]
(b) 0.600
0.40 x 0.40 m2,
Parabola-rectangle fck = 80 MPa,
0.500 4x3 20 B500
0.400
MRd [MNm]
0.300
FprEN 1992-1-1:2023
0.200 phi 8 @300
phi 8 @150
phi 10 @150
0.100 phi 12 @150
phi 14 @150
cc = 0.85
EN 1992-1-1:2004 a_cc0.85
cc = 1.00
EN 1992-1-1:2004 a1.00
0.000
0 -1 -2 -3 -4 -5 -6 -7 -8 -9 -10
NRd [MN]
(c) 0.600
0.40 x 0.40 m2,
Parabola-rectangle fck = 40 MPa,
0.500 4x3 20 B500
0.400
MRd [MNm]
0.300
FprEN 1992-1-1:2023
0.200 phi 8 @300
phi 8 @150
phi 10 @150
0.100 phi 12 @150
phi 14 @150
cc = 0.85
EN 1992-1-1:2004 a_cc0.85
cc = 1.00
EN 1992-1-1:2004 a1.00
0.000
0 -1 -2 -3 -4 -5 -6 -7 -8 -9 -10
NRd [MN]
(d) 0.600
0.40 x 0.40 m2,
Parabola-rectangle fck = 80 MPa,
0.500 4x3 20 B500
0.400
MRd [MNm]
0.300
FprEN 1992-1-1:2023
0.200 phi 8 @300
phi 8 @150
phi 10 @150
0.100 phi 12 @150
phi 14 @150
cc = 0.85
EN 1992-1-1:2004 a_cc0.85
cc = 1.00
EN 1992-1-1:2004 a1.00
0.000
0 -1 -2 -3 -4 -5 -6 -7 -8 -9 -10
NRd [MN]
(e) 0.600
0.40 x 0.40 m2,
Parabola-rectangle fck = 40 MPa,
0.500 4x3 20 B500
0.400
MRd [MNm]
0.300
FprEN 1992-1-1:2023
0.200 phi 8 @300
phi 8 @150
phi 10 @150
0.100 phi 12 @150
phi 14 @150
cc = 0.85
EN 1992-1-1:2004 a_cc0.85
cc = 1.00
EN 1992-1-1:2004 a1.00
0.000
0 -1 -2 -3 -4 -5 -6 -7 -8 -9 -10
NRd [MN]
(f) 0.600
0.40 x 0.40 m2,
Stress-block fck = 80 MPa,
0.500 4x3 20 B500
0.400
MRd [MNm]
0.300
FprEN 1992-1-1:2023
0.200 phi 8 @300
phi 8 @150
phi 10 @150
0.100 phi 12 @150
phi 14 @150
EN 1992-1-1:2004 cc = 0.85
a0.85
EN 1992-1-1:2004 cc = 1.00
a1.00
0.000
0 -1 -2 -3 -4 -5 -6 -7 -8 -9 -10
NRd [MN]
(g) 0.600
0.40 x 0.40 m2,
Stress-block fck = 40 MPa,
0.500 4x3 20 B500
0.400
MRd [MNm]
0.300
FprEN 1992-1-1:2023
0.200 phi 8 @300
phi 8 @150
phi 10 @150
0.100 phi 12 @150
phi 14 @150
EN 1992-1-1:2004 cc = 0.85
a0.85
EN 1992-1-1:2004 cc = 1.00
a1.00
0.000
0 -1 -2 -3 -4 -5 -6 -7 -8 -9 -10
NRd [MN]
(h) 0.600
0.40 x 0.40 m2,
Stress-block fck = 80 MPa,
0.500 4x3 20 B500
0.400
MRd [MNm]
0.300
FprEN 1992-1-1:2023
0.200 phi 8 @300
phi 8 @150
phi 10 @150
0.100 phi 12 @150
phi 14 @150
EN 1992-1-1:2004 cc = 0.85
a0.85
EN 1992-1-1:2004 cc = 1.00
a1.00
0.000
0 -1 -2 -3 -4 -5 -6 -7 -8 -9 -10
NRd [MN]
Figure C8.1.9: Influences of the main parameters on the NRd - MRd resistances of a typical square column,
cases (a)-(h) described above, DLower = 24 mm
[1] B. Halldórsson, “Experimental study of strength and fracture of mortar under complex stress states,
References
Master Thesis, Technical University of Denmark,” Lyngby, 2016.
[2] F. Vind and F. Jensen, “Failure criteria for concrete and mortar subjected to triaxial stress states,
Bachelor Thesis, Technical University of Denmark,” Lyngby, 2017.
[3] E. Hoek and J. A. Franklin, “Simple triaxial cell for field or laboratory testing of rock,” Trans. Inst. Min.
Metall., vol. 77, pp. A22-26, 1968.
[4] Q. Li and F. Ansari, “High-strength concrete in triaxial compression by different sizes of specimens,”
ACI Struct. J., vol. 97, no. 6, pp. 684–689, 2000.
[5] D. Campbell-Allen, “Strength of concrete under combined stresses,” Constr. Rev., vol. 35, no. 4, pp.
29–37, 1962.
[6] D. J. Smee, “The Effect of Aggregate Size and Concrete Strength on the Failure of Concrete under
Triaxial Compression,” Civ. Eng. Trans. Inst. Eng. Aust., vol. 9, no. 2, pp. 339–344, 1967.
[7] K. K. B. Dahl, “A Failure Criterion for normal and High Strength Concrete, Series R No 286,”
Department of Structural Engineering, Technical University of Denmark, Lyngby, 1992.
[8] D. W. Hobbs, “Strength and deformation properties of plain concrete subject to combined stress -
Part 1: Strength results obtained on one concrete,” London, Cement and Concrete Association,
Report 42.451, 1970.
[9] K. Lahlou, P. C. Aïtcin, and O. Chaallal, “Behaviour of high-strength concrete under confined
stresses,” Cem. Concr. Compos., vol. 14, no. 3, pp. 185–193, 1992.
[10] J.-C. Chern, H. Yang, and H.-W. Chen, “Behavior of Steel Fiber Reinforced Concrete in Multiaxial
Loading,” ACI Mater. J., vol. 89, no. 1, pp. 32–40, 1992.
[11] J. Xie, A. E. Elwi, and J. G. MacGregor, “Mechanical Properties of Three High-Strength Concretes
Containing Silica Fume,” ACI Mater. J., vol. 92, no. 2, pp. 135–145, 1995.
[12] I. Imran and S. J. Pantazopoulou, “Experimental Study of Plain Concrete under Triaxial Stress,” ACI
Mater. J., vol. 93, no. 6, pp. 589–601, 1996.
[13] F. Ansari and Q. Li, “High-Strength Concrete Subjected to Triaxial Compression,” ACI Mater. J., vol.
95, no. 6, pp. 747–755, 1998.
[14] X. Lu and C. T. Hsu, “Stress-Strain Relations of High-Strength Concrete under Triaxial Compression,”
J. Mater. Civ. Eng., vol. 19, no. 3, pp. 261–268, 2007.
[15] F. Richart, A. Brandtzaeg, and R. L. Brown, “A Study of the Failure of Concrete under Combined
Compressive Stresses, Bulletin No. 26,” Univ. Illinois Bull., pp. 1–104, 1928.
[16] G. Balmer, “Shearing strength of concrete under high triaxial stress -- computation of Mohr’s
envelope as a curve, Structural Research Laboratory Report No. SP-23,” Denvar, Colorado, 1949.
[17] T. N. W. Akroyd, “Concrete under triaxial stress,” Mag. Concr. Res., vol. 13, no. 39, pp. 111–118,
1961.
[18] N. Gardner, “Triaxial Behavior of Concrete,” ACI J., vol. 66, no. 2, pp. 136–158, 1969.
[19] D. Sfer, I. Carol, R. Gettu, and G. Etse, “Study of the behavior of concrete under triaxial
compression,” J. Eng. Mech., vol. 128, no. 2, pp. 156–163, 2002.
[20] Fardis M., Seismic Design, Assessment and Retrofitting of Concrete Buildings: based on EN-
Eurocode 8, Springer, 2009, 742 p.
[21] Moccia F, Yu Q, Fernández Ruiz M, Muttoni A., “Concrete compressive strength: From material
characterization to a structural value”. Structural Concrete. 2020; 21 p..
Background document to
clauses 8.2.1 and 8.2.2
Shear in members without shear
reinforcement
For citations, copying and referencing see Note on Front page! page 298
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.2.1 and 8.2.2 Background to FprEN 1992-1-1:2023
Introduction
This document provides the background to clauses 8.2.1 and 8.2.2 and details the derivation of the closed-form
formulae for shear design of members without shear reinforcement based on the original formulation of the
Critical Shear Crack Theory (CSCT).
where:
VRd,c is the shear resistance;
CRd,c is a constant equal to 0.18/γc (recommended value)
k is the size factor which is equal to 1+(200/d)1/2≤2;
ρl is the longitudinal reinforcement ratio (ρl≤2%);
fck is the characteristic compressive strength of concrete measured in cylinder;
k1 is a constant equal to 0.15 (recommended value);
σpd is the axial stress and it is equal to NEd/Ac < 0.2 fcd;
NEd is the axial force (NEd<0 in compression);
bw is the width of the cross section;
d is the effective depth of the member.
0
0 0.50 1.0 1.5 2.0 -20 -10 0 10 20
d [m] σp [MPa]
Figure C8.1: Ratio test/EN 1992-1-1: 2004 for the case of simply supported beams subjected to point load: (a) for
members without axial force as a function of the effective depth d (database by Reineck et al. [C8.3]); (b) for
members with axial force, as a function of the axial stress σp (database by Cavagnis [C8.6]); for comparison with
database, average material properties (fcm) have been used and the partial safety factor is assumed equal to 1.
where:
VR,c represents the shear resistance;
εv refers to the strain in the reinforcement;
kdg accounts for the concrete type and can be calculated as kdg = 48/(16+Dlower) = 48/ddg;
fc is the compressive strength of concrete measured in cylinder;
bw is the width of the cross section;
d is the effective depth of the member.
According to Muttoni and Fernández Ruiz [C8.1], a linear relationship between the acting bending moment and
the strain in the reinforcement (εv) can be assumed. The strain in the reinforcement (εv) can thus be calculated
as:
ME ME VE acs
v (C8.3)
z As Es z l bw d Es z l bw d Es
where:
ME is the acting bending moment at the control section;
VE is the acting shear force at the control section;
acs= │ME/VE│ is the effective shear span at the control section;
As is the area of the longitudinal reinforcement;
l is the longitudinal reinforcement ratio;
z is the effective level arm of the longitudinal internal forces;
Es is the modulus of elasticity of the longitudinal reinforcement.
Shear design by using the failure criterion (C8.2) and the load-deformation relationship (C8.3) is very simple as
shown in Figure C8.2, since it can be performed by checking the inequality VEd ≤ VR,c. Calculating the actual
resistance (VEd =VRd,c) requires yet solving the set of Formulae (C8.2) and (C8.3). This can be done iteratively or
by directly solving, in a closed-form manner, the resulting quadratic Formula (Fernández Ruiz et al. [C8.10])
which is not sufficiently simple for a code formula.
0.4
0.3
[√ MPa [
CSCT failure criterion
0.2
bd √ fc
VR,c
load-deformation
relationship
0.1
0
0 1 2 3 4 5
ε v·d·kdg [-]
Figure C8.2: Design using the CSCT failure criterion and the load-deformation relationship.
This problem can be solved on the basis of a recently published refinement of the mechanical of the CSCT
[C8.4, C8.5, C8.9], where an improved general failure criterion can be formulated as follows:
12
fc ddg
VR,c k bw d VR,c 0 (C8.4)
ε d
v
where:
k depends on the location and shape (inclination) of the critical shear crack and, based on [C8.4,C8.5,
C8.6], can be estimated to depend upon the ratio acs/d resulting into k = 0.015ꞏ(acs/d)1/4;
ddg = 16 mm+Dlower ≤ 40 mm refers to the average roughness dimension of the critical shear crack (see
definition in 8.2.1(4));
VR,c0 is the maximum shear resistance.
This failure criterion, in combination with the load-deformation relationship of Formula (C8.3), can be used to
derive closed-form expressions to calculate the shear resistance, as it will be shown in the following.
Figure C8.3: Comparison of hyperbolic (curve in black, Formula (C8.2) and power-law (curve in red, Formula
(C8.4) with a value of acs/d=2 (green), 4 (blue) and 6 (red)) failure criteria.
(Formula (C8.4)), it results (assuming a value of acs/d = 4 for the slenderness ratio):
12
Es fc ddg
VR,c 0.021 bw d (C8.6)
fy d
Considering the value of the modulus of elasticity for the reinforcing steel as Es = 200’000 MPa and the fact that
according to Formulae (8.10a) and (8.10b), the design stress is calculated using the level arm z ≈ 0.9d, Formula
(C8.6) can be written in a design format as:
12
11 fck ddg
Rdc,min (C8.7)
V fyd d
where the partial safety factor V, it is defined in 4.3.3(1) and its justification can be found in the related
background document.
It has to be noted that the yield strength used in Formula (C8.7) is the strength that has been assumed for the
flexural design of the cross section. This means that if, for an old existing structure, the longitudinal
reinforcement has been designed using a lower yield strength than those used nowadays, the member will have
a higher mechanical longitudinal reinforcement ratio and, consequently, a higher shear resistance.
Physical meaning of the minimum shear resistance, cases with over-designed flexural
reinforcement or axial forces
It is important to note that, according to its definition, the minimum shear resistance ensures a member to reach
yielding of the flexural reinforcement and the shear resistance at the same load level. This presumes that
MEd = MRd at the control section for the shear verification. However, the available flexural reinforcement is
usually larger than the required one for bending conditions (i.e. MRd > MEd, verification for shear of sections
where bending is not governing for design, or cases where the reinforcement has been designed for other
criteria as SLS or fatigue for instance, or where the reinforcement layout has been simplified on the safe side).
For such cases, when the available flexural reinforcement (av) is larger than the required reinforcement for
flexural design (req), the minimum shear resistance could be increased as follows:
12
11 fck ddg av
Rdc,min (C8.8)
V fyd d req
or
12
11 fck ddg
Rdc,min (C8.9)
V sd d
where sd refers to the reinforcement stress in the reinforcement for the design situation governing for the shear
verification.
This potential refinement is not included in clause 8.2.1 of FprEN 1992-1-1:2023. Formula (C8.7) is thus a
simple and safe estimate of the resistance, yet covering the shear demands for most usual members. Cases
with over-designed flexural reinforcement and where the shear resistance given by Formula (C8.7) might not be
sufficient, are addressed in a general manner in clause 8.2.2 (“detailed verification for members not requiring
shear reinforcement”) where the actual reinforcement ratio is accounted for explicitly.
The same comment applies also to members subjected to axial forces. For a given section, applying a
compressive axial force to a member whose reinforcement has been designed without accounting for it, would
reduce the strain in the longitudinal reinforcement, and would therefore increase the shear resistance. This can
be seen in Formula (C8.8), where applying a moderate compressive axial force reduces the amount of required
flexural reinforcement (req) for a given effective shear span, and allows thus to increase the value of Rdc,min if
the available flexural reinforcement (av) remains unchanged. In the case where the available reinforcement has
been designed accounting for the axial force, the available reinforcement is smaller, and the corresponding
shear resistance will consequently remain unchanged. This explains the reason why the effect of axial force
should not appear explicitly in Formula (C8.7).
Figure C8.4: Shear resistance as a function of the plastic deformation (normalized plastic rotation on the
abscissa; all tests with constant cross section and longitudinal reinforcement ratio; figure from [C8.12]).
d1
d d d
d2
d d d
d2
d1
control sections
Figure C8.5: Locations of the control sections according to the CSCT (figure from [C8.6]).
MEd
acs d (8.19)
VEd
where the limit acs ≥ d replaces the upper limit VRc,0 in Formula (C8.4). Formula (C8.10) can be easily solved to
which can be written in an analogous format to EN 1992-1-1:2004 (considering the value of the modulus of
elasticity for the reinforcing steel as Es = 200’000 MPa) as [C8.6]:
13
ddg
VR,c 0.60 100 ρl fc bw d (C8.12)
acs
d
4
For the code formulation, Formula (C8.12) needs to be adapted by using characteristic values and partial safety
factors. In addition, to enhance the ease of use, the calculation of the shear resistance can be performed in a
first step by adopting a value acs = 4ꞏd (generally safe for members subjected predominantly to distributed
loading when shear is governing considering the exponent associated with the parameter acs). In that case,
Formula (C8.12) can thus be written in a design format as:
13
VRd ,c VRd ,c 0.66 ddg
Rd ,c 100 l fck (8.16)
bw z bw 0.9 d V d
where the value 0.66 replaces 0.60 in Formula (C8.12) since Rd,c is calculated on the basis of the lever arm
z 0.9 d according 8.2.1(3).
For a more refined calculation, parameter d in Formula (8.16) may be replaced by av acs d / 4 d (see
8.2.2(3).
Table C8.1: Summary of experimental-to-calculated shear resistance ratio of 669 simply supported beams or
cantilevers subjected to point load (database of Reineck et al. [C8.3], completed with the tests by Cavagnis et
al. [C8.8]).
Number of tests Approach Mean *) COV [%] *)
Figure C8.6: Ratio R,test/R,FprEN2023 as a function of the main mechanical and geometrical parameters for the
case of simply supported beams subjected to point load: (a) shear span-to-effective depth ratio a/d; (b)
longitudinal reinforcement ratio ρ; (c) effective depth d; (d) width-to-effective depth ratio b/d; (e) cylinder
compressive strength of concrete fc; (f) aggregate size Dlmax
ME NE z d h / 2 ME NE d / 3
v (C8.13)
z Al Es z Al Es
This formula can be expressed by defining an effective shear span which accounts for the presence of the axial
force:
ME N d
acs,N E (C8.14)
VE VE 3
It can be noted that the effective shear span accounting for the axial force has a physical meaning as shown in
Figure C8.7(a).
To allow for the simplified use of Formula (8.16) (using d instead of av), the corrected shear span acs,N of
previous Formula can be replaced by coefficient kvp which may be used to reduce the value of d in Formula
(8.16) or av acs d / 4 in case of a refined calculation:
NEd d
kvp 1 0.1 (8.20)
VEd 3 acs
In case of tensile axial force, the same coefficient should be used to increase d or av.
NEd d
acs = a' +
VEd 3 d
(a)
h d
NEd /VEd z-d+ ~
2 3
- NEd z
a
VEd VEd a' + NEd z - d + h/2 =VEd cs
z z z
a'
Pd
acs = a' - z
VEd d
(b)
MEd = VEd ∙a' - Pd∙(dp - h/2)
h
2 Pd Pd
dp acs = a' - (dp - h/2) - (z - d + h/2)
VEd VEd
ep Pd Pd
= a' - z ~ a' - (d/3 + ep)
VEd VEd
Pd
VEd
a'
acs d
(c)
Hd
acs unchanged
VEd
Figure C8.7: equilibrium of internal forces: members subjected to (a) centred axial force; (b) prestressing force
on the tension side; (c) prestressing force on the compression side (from [C8.9]).
Consideration of prestressing
Prestressing has three potential influences on the shear resistance:
(i) the vertical component of the prestressing force related to the longitudinal axis may be reduced from
the design value of the shear force VEd according to 8.2.1(8) in case prestressing is considered as an
imposed deformation between prestressing steel and concrete (see approach (a) in 7.6.1(1)). This
effect is accounted for implicitly by considering prestressing as a set of self-equilibrated system of
actions (see approach (b) in 7.6.1(1)). In this case, the shear force VEd is reduced by the effect of
deviation forces and the vertical component of anchorage forces;
(ii) the horizontal component influencing the axial force NEd, and thus the shear resistance according to
8.2.2(4). This effect is accounted for implicitly if prestressing is considered according approach (b) in
7.6.1(1) (as external action) or by adding the term -Pd∙cosp to NEd if prestressing is considered
according approach (b) in 7.6.1(1).
(iii) the eccentricity of the tendon, influencing the reinforcement force and thus the shear resistance. This
effect is accounted for implicitly if prestressing is considered according approach (b) in 7.6.1(1) (as
external action) or by adding the term -Pd∙ep∙cosp to MEd if prestressing is considered according
approach (b) in 7.6.1(1).
Figure C8.7(b) shows that the effect of tendon eccentricity is relevant for prestressed members, as it reduces
the shear span with respect to members subjected to a centred axial force (Figure C8.7(a)). On the contrary,
prestressing on the compression side of the member shows negligible influence on the shear response (Figure
C8.7(c)).
In conclusion, the three effects (i-iii) described above are accounted for implicitly if prestressing is considered as
an external action (approach (b) in 7.6.1(1)) or, if prestressing is considered as an imposed deformation
between prestressing steel and concrete (approach (a) in 7.6.1(1)), by adapting the internal forces according to
8.2.1(8).
Figure C8.8: Ratio R,test/R,FprEN 1992-1-1:2023 for simply supported beams subjected to point load and to centred
axial force (database by Cavagnis [C8.6]).
d d Δx d
q q
d
MEd MEd
VEd VEd
lc lc lt
Figure C8.9: Definition of Δx for the case of distributed loads acting (a) near simple supports of end spans
(Δx=0) and (b) near intermediate supports or on cantilevers (Δx=d).
Figure C8.10: Ratio R,test/R,FprEN 1992-1-1:2023 for members with distributed loads near simply supports of end
spans (load on the compression side) as a function of (a) acs/d; (b) the span-to-effective depth ratio lc/d (refer to
Figure C8.9); (c) the effective depth d (database by Cavagnis [C8.6]).
Figure C8.11 illustrates the comparisons of the shear resistances calculated by using Formula (8.16) (again,
accounting for the value of the mechanical shear slenderness av according to 8.2.2(3)) against the test results of
members with distributed loads near intermediate supports or on cantilevers as a function of control section
shear span-to-effective depth ratio acs/d (Figure C8.11(a)) and length-to-effective depth ratio lt/d (Figure
C8.11(b)), where lt is the distance from the intermediate or fixed support to the section of zero shear force (see
Figure C8.9(b)). In this comparison, also the load VEd according to 8.2.2(8) is accounted for. The results show
that this approach agrees consistently with the experimental results.
Table C8.2 presents the summary of the experimental-to-calculated shear resistance ratios of members with
distributed loads (both near simply supports of end spans and intermediate supports or cantilevers), where the
shear resistance is calculated with Formula (8.16) without (approach (1), acs=4d) and with an explicit calculation
of the mechanical shear slenderness (approach (2) accounting for the actual value of acs ≤ 4). The results show
again that a satisfactory estimation of the shear resistance can be first obtained considering acs=4d, while an
explicit calculation of this parameter will normally lead to a better prediction (both in terms of average and
coefficient of variation).
Figure C8.11: Ratio R,test/R,FprEN 1992-1-1:2023 for members with distributed loads near intermediate supports or on
cantilevers (load on the tension side) as a function of (a) acs/d; (b) the span-to-effective depth ratio lt/d defined in
Figure C8.9 (database by Cavagnis [C8.6]).
Table C8.2: Summary of experimental-to-calculated shear resistance ratio of members with distributed loads
near simply supports of end spans and members with distributed loads near intermediate supports or
cantilevers (databases by Cavagnis [C8.6]).
Number of tests Approach Average COV [%]
1.18 14.5
(1)
Members with distributed loads near simply lc / d ≥ 5 : 68 (1.36) (23.3)
supports of end spans (all tests: 117) 1.00 13.5
(2)
(1.16) (24.3)
Members with distributed loads near (1) 1.30 16.3
43
intermediate support of cantilevers (2) 1.12 13.1
(1) Without calculation of av, Formula (8.27) on the basis of d according to 8.2.2(2)
(2) With calculation of av, Formula (8.27) with d replaced by av acs d / 4 according to 8.2.2(3)
[C8.1] Muttoni A., Fernández Ruiz M., Shear strength of members without transverse reinforcement as
References
function of critical shear crack width. ACI Structural Journal, Vol. 105, N° 2, 2008, pp. 163-172.
[C8.2] Muttoni A., Shear and Punching Strength of Slabs Without Shear Reinforcement, (in German:
Schubfestigkeit und Durchstanzen von Platten ohne Querkraftbewehrung, Beton- und Stahlbetonbau)
Beton-und Stahlbetonbau, Vol. 98, 2003, pp. 74-84.
[C8.3] Reineck K., Bentz E.C., Fitik B,. Kuchma D.A., Bayrak O., ACI-DAfStb Database of Shear Tests on
Slender Reinforced Concrete Beams without Stirrups. ACI Structural Journal, Vol. 110, N° 5, 2013,
pp. 867–876.
[C8.4] Cavagnis F., Simões J., Fernández Ruiz M., Muttoni A., “Shear Strength of Members without
Transverse Reinforcement Based on Development of Critical Shear Crack.”, ACI Structural Journal
(2020), 103-118
[C8.5] Cavagnis F., Fernández Ruiz M., Muttoni A., “A mechanical model for failures in shear of members
without transverse reinforcement based on development of a critical shear crack.”, Engineering
Structures (2018), 200-315
[C8.6] Cavagnis F., Shear in reinforced concrete without transverse reinforcement: from refined
experimental measurements to mechanical models. PhD Thesis. Thesis No. 8216. Lausanne,
Switzerland: EPFL, 2017.
[C8.7] Cavagnis F., Fernández Ruiz M., Muttoni A., Shear failures in reinforced concrete members without
transverse reinforcement: An analysis of the critical shear crack development on the basis of test
results, Engineering Structures, Vol. 103, 2015, pp.157–173.
[C8.8] Cavagnis F., Fernández Ruiz M., Muttoni A., An analysis of the shear transfer actions in reinforced
concrete members without transverse reinforcement based on refined experimental measurements,
Structural Concrete, 2017, doi: 10.1002/suco.201700145
[C8.9] Muttoni A., Fernández Ruiz M., Cavagnis F., Shear in members without transverse reinforcement:
from detailed test observations to a mechanical model and simple expressions for codes of practice,
fib International Workshop on Beam Shear, Zurich, Switzerland, September 5-6, 2016, fib Bulletin 85.
[C8.10] Fernández Ruiz M., Muttoni A., Sagaseta J., Shear strength of concrete members without transverse
reinforcement: a mechanical approach to consistently account for size and strain effects. Engineering
Structures, Vol. 99, 2015, pp. 360-372.
[C8.11] Leonhardt F., Walther R., Shear Tests on Beams With and Without Shear Reinforcement, Deutscher
Ausschuss für Stahlbeton, No. 151, 1962, 83 pp.
[C8.12] Vaz Rodrigues R., Muttoni A., Fernández Ruiz M., “Influence of Shear on Rotation Capacity of
Reinforced Concrete Members Without Shear Reinforcement”, ACI Structural Journal, V. 1072010.
[C8.13] Muttoni A., The Applicability of the Theory of Plasticity in the Design of Reinforced Concrete, (in
German: Die Anwendbarkeit der Plastizitätstheorie in der Bemessung von Stahlbeton), Report No.
176, Zürich, Switzerland, 1989, 159 pp.
[C8.14] Muttoni A., Thürlimann B., Shear Tests on Beams and Slabs Without Shear Reinforcement, Institut
für Baustatik und Konstruktion, Zürich, Switzerland, 1986, 12 pp.
[C8.15] Natário F., Fernández Ruiz M., Muttoni A., Shear strength of RC slabs under concentrated loads near
clamped linear supports. Engineering Structures, Vol. 76, 2014, pp. 10-23.
[C8.16] Pérez Caldentey A., Padilla P., Muttoni A., Fernández Ruiz M., Effect of load distribution and variable
depth on shear resistance of slender beams without stirrups, ACI Structural Journal, Vol. 109, N°. 5,
2012, pp. 595-603.
[C8.17] Reineck, K.-H.; Dunkelberg, D., ACI-DAfStb databases 2015 on shear tests for evaluating
relationships for the shear design of structural concrete members without and with stirrups. Report for
Research Project DAfStb V479. (Eds.). DAfStb H. 617, Beuth Verl. Berlin, 2015.
[C8.18] Reineck K., Todisco L., Database of Shear Tests for Non-slender Reinforced Concrete Beams
without Stirrups. ACI Structural Journal, Vol. 111, N° 6, 2014, pp. 1363.1372.
[C8.19] Sigrist V., Bentz E. C., Fernández Ruiz M., Foster S. J., Muttoni A., Background to the Model Code
2010 Shear Provisions - Part I: Beams and Slabs. Structural Concrete, Vol. 14, No. 3, 2013, pp. 204-
214.
03.01.2023
For citations, copying and referencing see Note on Front page! page 311
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.2.2 (5) Background to FprEN 1992-1-1:2023
Introduction
This document has been prepared in order to give a detailed background to the proposed EC2-provisions for
considering the effects of normal compressive forces on the shear capacity of members without shear
reinforcement. The shear capacity is increased by the stress level required for decompression in the control
section.
[1] EN 1992-1-1:2004: Eurocode 2: Design of concrete structures - Part 1-1: General rules and rules for
References
buildings. 2004.
[2] Zsutty, T.C.: Beam Shear Strength Prediction by Analysis of Existing Data. In: ACI Journal, Vol. 65
(1968), Iss. 11, pp. 943-951.
[3] Hedman, O.; Losberg, A.: Design of Concrete Structures with Reagrd to Shear Forces. In: Regan,
P.E.; Taylor, H.P.J. (Hrsg.): Shear and Torsion – Explanatory and Viewpoint Papers on Model Code
chapters 11 and 12 prepared by members of CEB Commission V, CEB Bulletin d'Information 126,
1978, S. 184-209.
[4] Walraven, J.C.: Background document for EC-2, Chapter 6.2 Shear. Delft University of Technology,
Delft 2002.
[5] Miguel, P.; Fernández, M.; Hegger, J.; Schmidt, M: Shear resistance of members without shear
reinforcement in in presence of compressive axial forces in the next Eurocode 2. In: Hormigón y
Acero, 2023
For citations, copying and referencing see Note on Front page! page 313
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.2.3 Background to FprEN 1992-1-1:2023
Introduction This document has been prepared to clarify the background of the changes introduced in the design provisions
of FprEN 1992-1-1:2023 with respect to Section 8.2.3 (“Members requiring design shear reinforcement”). These
provisions inherit the approach of former Section 6.2.3 of EN 1992-1-1:2004, maintaining the variable-angle
truss model approach grounded on limit analysis. It however updates the values for some parameters, notably
of efficiency factors, as well as the provisions for direct support.
Reasons for change Current provision for shear design of members with more than minimum amount of shear reinforcement in EN
1992-1-1:2004 (clause 6.2.3) is based on a variable-angle truss model (see a detailed description of the method
in [C8.5.1-3]). This model is a lower-bound solution based on limit analysis, allowing to select the angle of
inclination of the compression field. Such approach is sound, provides the engineer with a clear understanding
of the load-carrying actions and is consistent with section 8.5 (design based on strut-and-tie and stress field
models). In addition, it allows considering also members with variable depth, geometric discontinuities (e.g.
dapped-end beams and beams with openings), concentrated force introductions (loads and anchor forces),
indirect supports and other cases which are not uncommon in practice. The content of this subsection is also
fully consistent with Annex G (Design of membrane-, shell- and slab elements).
For FprEN 1992-1-1:2023, the core of the EN 1992-1-1:2004 approach is maintained, but some changes are
nevertheless introduced to update them to current state-of-the-art with respect to a number of topics:
i. Minimum allowable angle: as lower angles than those provided in EN 1992-1-1:2004 have been
consistently observed in experimental programmes of prestressed beams [C8.5.4,5]
ii. Efficiency factors for concrete compressive strength: allowing for a detailed consideration of the strain
state, consistently with Section 8.5 and Annex G of prEN 1992-1-1:2020-11 (D7)
iii. Consideration for loads applied near supports: providing a design model based on the same theory as
the variable-angle truss model [C8.5.6], replacing current empirical formulation of Section 6.2.3(8) in
EN 1992-1-1:2004, which provides in addition poor fitting to available experimental data.
Minimum angle of the The minimum angle of the compression field has been reduced for members subjected to significant axial
compression field compression forces. This change is ground on several experimental evidences [C8.5.4,5] and allows for a more
economic design.
8.2.3(4)
Conversely, the angle has been limited for members in tension, following the considerations of MC2010
[C8.5.7,8], which makes the provision also consistent with strain-based analyses.
In addition, the minimum angle of the compression field has been increased (cotmin reduced) for the case of
brittle shear reinforcement (ductility class A). The reasons for this change are described in detail in the
background document to clause 12.2(4) (minimum shear reinforcement).
Design formulae for shear These provisions have been reworked with respect to the formulation of current EN 1992-1-1:2004. Their
resistance verification of theoretical ground remains however unchanged (variable-angle truss model). The proposed formulation allows
concrete stress in the for a direct calculation of the optimum angle of the compression field (simultaneous crushing of compression
compression field field and yielding of stirrups, minimising the amount of shear reinforcement to be arranged, refer to Formula
8.2.3(5) (8.31)).
Efficiency factor for the The value of the efficiency factor is modified with respect to EN 1992-1-1:2004:
effective concrete
The influence of concrete brittleness (reduction of for high concrete compressive strength) is
strength in the
considered in FprEN 1992-1-1:2023 in the definition of fcd and thus has been removed from the
compression field
efficiency factor of the compression field;
8.2.3(6-7)
The proposed constant value of the efficiency factor ( = 0.5, 8.2.3(6)) has been adapted to match the
strain-based definition of that factor in 8.2.3(7) (similarly as performed in MC2010 [C8.5.7,8], following
the researches of Vecchio and Collins [C8.5.9]) by considering that the flexural reinforcement is at
yielding (x ≈ 0.002/2 = 0.001, safe estimate), refer to Figure C8.2.3.1.
0.8
0.6
0.5
Constant value §8.2.3(6)
0.4
0.3
0.2
0.1
0
0 10 20 30 40 50 60
[°]
Figure C8.2.3.1: Comparison of constant and variable value of concrete efficiency factor
Disturbance by ducts in A new case has been introduced to consider the disturbance of plastic ducts in the resistance of the
the web compression field as considered in MC 2010 [C8.5.7] (for details, refer to Muttoni et al. [C8.5.9]).
8.2.3(10)
Concentrated loads near Remark: The material presented hereafter is a concise version of [C8.5.6] which is referred to for further details.
to supports
8.2.3(12) This clause replaces the empirical approach of EN 1992-1-1:2004 (§6.2.3(8)) with an approach based on a
stress field consistent with the variable-angle truss model and the stress field approach presented in 8.5.
The approach of EN 1992-1-1:2004 (§6.2.3(8)) is based on a reduction of the action for concentrated loads
applied in between 0.5d and 2d from a support to account for a direct strutting (so called “arching action”,
development of a direct strut between load and the support). In addition, only 75% of the shear reinforcement in
the region between support and concentrated load can be accounted for. Such approach has an empirical
basis, and its performance is in addition poor when compared to test results, see Figure C8.2.3.2. This is related
to the fact that there is no transition between the case with direct strutting (av<2d) and the case without direct
strutting (av≥2d). As a consequence, for concentrated loads in between 1.2 and 2d, the model is overly
conservative (due to the fact that only 75% of the reinforcement is accounted for).
Figure C8.2.3.3: Considered stress field in prEN 1992-1-1:2020-11 (D7): (a) detail of stress field; and (b)
consideration of concrete and reinforcement contributions (from [C8.5.6])
On that basis, by analysing the free-body shown in Fig. C8.2.3.3b, the total shear resistance can be calculated
as the sum of a contribution related to the direct strut (drawn in blue in Fig. C8.2.3.3b) plus a contribution related
to the stirrups (drawn in red in Fig. C8.2.3.3b):
As it can be noted, the inclination of this free-body corresponds to the angle defining the clear shear span and
thus is not equal to (angle of the compression field) as usually performed for design of stirrups in slender
members. By introducing the value of the two components, it results (see Fig. C8.2.3.3b):
consistently with the design of slender members), Asw to the area of stirrups at a given horizontal position (sum
of multiple legs if applicable), s to the stirrup spacing and swd to the stress in the stirrups. This Formula can be
expressed in terms of the shear stress resistance (Rd = V Rd /(bwz)) as:
cot cot
Rd f cd w swd cot (C8.2.3.3)
1 cot 2
The term swd can be calculated in a general manner accounting for the compatibility of deformations in the clear
shear span region (by making use of the Mohr’s circle for instance, see Figure C8.2.3.4):
Figure C8.2.3.4: Calculation of vertical strain in the clear shear span region (from [C8.5.6])
For practical purposes, when cot> 1, this term can be assumed to be equal to the yield stress of the
reinforcement (swd ≈ fywd), which is in agreement to experimental evidence and also elastic-plastic analyses of
this region. Thus, the resulting expression results:
cot cot
Rd f cd w f ywd cot (C8.2.3.5)
1 cot 2
which corresponds to Formula (8.55). In addition, by performing a free-body limited by a vertical line (parallel to
the stirrups) at mid –distance of the clear shear span, it will also result the crushing condition of the fans and
direct strut (all having the same inclination ):
where the latter inequality condition will refer to the strength outside of this region as governing. For the case of
variable values of (8.2.3(7), Formula (8.45), where can be calculated on the basis of the compression field
inclination and the longitudinal strain x), an analytical solution is not possible and the optimum compression
field inclination could be obtained by numerical iteration. A first estimate of this value can be obtained in a
simple manner by approximating as follows:
As it can be observed, for values cot<1, design can be performed on the basis of Formula (C8.2.3.2),
combined with Formula (C8.2.3.4) and by optimizing the resistance (for a variable or constant value of ),
depending on the case and performed approach.
The comparison of the FprEN 1992-1-1:2023 to pertinent tests [C8.5.6] shows a neat improvement on the
accuracy of the model, as well as a uniform level of safety for squat and slender members, see Figure C8.2.3.6.
It has to be noted that for the approach of prEN 1992-1-1:2020-11(D7), the limit between with and without direct
strutting depends on the inclination cot with respect to cot. This means that for members with low shear
reinforcement ratio, this limit can be significantly higher than a/d = 2.
Figure C8.2.3.6: Comparison of FprEN 1992-1-1:2023 to test data; (a) case with constant efficiency factor
(0.50) ad (b) case with variable effciency factor according to formula (8.32), (red marks
refer to tests with direct strutting (cot < cot); blue marks indicate no direct strutting analysed
according to 8.2.3(3-6)) (from [C8.5.6])
Other considerations, as the increase of tension force in the flexural reinforcement (introduced as an increase of
the design bending moment MEd) or the consideration of inclined shear reinforcement, can be obtained on the
basis of the same stress field.
Inclined shear For members with inclined shear reinforcement (45° ≤ w < 90°), design formulae are derived on the basis of the
reinforcement stress field approach.
8.2.3(13) It has to be noted, that these provisions shall not be used in case of shear reinforcement arranged with an
inclination w > 90°. This is due to the fact that in this case, the principal tensile strain can be much larger than
for the case of vertical shear reinforcement (vertical with respect to the member axis), leading to wide cracks at
SLS and reduced efficiency factor at ULS. For these reasons, this situation, which could arise in slightly
inclined beams where the stirrups would be arranged vertically or for members with inclined shear reinforcement
with reverse shear force, should be avoided.
References [C8.5.1] Nielsen M. P., Braestrup M. W. and Bach F. “Rational Analysis of Shear in Reinforced Concrete
Beams.” IABSE Colloquium Proceedings P-15, Vol. 2, 1978, 16 p.
[C8.5.2] Thürlimann B. “Plastic Analysis of Reinforced Concrete Beams.” IABSE Colloquium - Plasticity in
Reinforced Concrete, Reports of the working commissions, Vol. 28, 1979, p. 71-90.
[C8.5.3] Nielsen, M. P., and Hoang, L. C. (2011), Limit Analysis and Concrete Plasticity, third edition, CRC
Press, Boca Raton, FL, 2011, 796 pp.
[C8.5.4] Fernández Ruiz, M., Muttoni, A., Shear strength of thin-webbed post-tensioned beams, American
Concrete Institute, Structural Journal, Vol. 105, No. 3, 2008, pp. 308—317
[C8.5.5] Rupf, M., Fernández Ruiz, M., Muttoni, A., Post-tensioned girders with low amounts of shear
reinforcement: shear strength and influence of flanges, Engineering Structures, Elsevier, Vol. 56,
2013, pp. 357—371
[C8.5.6] Pejatovic, M., Fernández Ruiz, M., Muttoni, A., Design of slender and squat reinforced concrete
members with shear reinforcement, Structural Concrete, November 2022.
https://doi.org/10.1002/suco.202200887
[C8.5.7] International Federation for Structural Concrete, “fib Model Code for Concrete Structures 2010,”
Ernst & Sohn, Berlin, Germany, 434 p., 2013.
[C8.5.8] Sigrist, V., Bentz, E., Fernández Ruiz, M., Foster, S.J., Muttoni, A., Background to the Model
Code 2010 Shear Provisions - Part I: Beams and Slabs, Structural Concrete, Vol. 14, No. 3,
2013, pp. 204--214
[C8.5.9] Vecchio, F. J., and Collins, M. P., “The Modified Compression-Field theory for Reinforced
Concrete Elements Subjected to Shear,” ACI Journal Proceedings, V. 83, No. 2, Mar.-Apr. 1986,
pp. 219-231.
[C8.5.10] Muttoni, A., Burdet, O. L., and Hars, E., “Effect of duct type on shear strength of thin webs”, ACI
Structural Journal, V. 103, No. 5, 2006, pp.729-735.
Background Document to
Clause 8.2.3(9): Members with circular
cross section
03.02.2023
For citations, copying and referencing see Note on Front page! page 319
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.2.3 (9) Background to FprEN 1992-1-1: 2023
Introduction
This document has been prepared to clarify the background for the provisions introduced in clause 8.2.3(9) of
prEN-1992-1-1:2021 for shear design of members with circular cross section.
Theoretical background
The design rules given in clause 8.2.3(9) of FprEN-1992-1-1:2023 are based on a rigid-plastic lower bound
model presented in [7], see also [8 & 9] for details. To illustrate the model, a circular beam subjected to constant
shear is considered (Figure 1). It is assumed that the shear force V is transferred to the shear span, a, by
means of a sufficiently strong diaphragm (to simplify the boundary conditions). The shear reinforcement consists
of circular stirrups with diameter Dh, closely placed at spacing s. In the longitudinal direction, reinforcement bars
are evenly distributed along the inner perimeter of the stirrups. The first step to establish an admissible stress
field (lower bound solution) to carry V is to assume that the stirrups are at yield and confine the concrete within
Dh, i.e. the shaded area shown in Figure 1(b). The concrete cover, which is not confined, may spall of prior to
the ultimate state. The cover is therefore disregarded.
Figure 1. Cantilever beam with circular cross section and reinforced with circular stirrups.
(a) (b)
Figure 2. Average confinement pressure p = fc on the concrete core within Dh.
As illustrated in Figure 2(a), the average confinement pressure p due to yielding of the stirrups can be
determined by:
L. C. Hoang 24.05.2021
For citations, copying and referencing see Note on Front page! page 320
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.2.3 (9) Background to FprEN 1992-1-1: 2023
that encapsulate the core. Within the top- and bottom segment, concentrated chord forces, C and T, are
assumed to develop to carry the bending moment. The two remaining segments have no function, other than to
provide confinement to the inner core. This also means that the longitudinal bars in these two segments are not
interacting with the inner core. They can therefore, in this lower bound model, not be used to carry the bending
moment.
The inner core, with dimension h and bw, is the cross section of the shear zone (Figure 3(b)). In this zone, a
homogenous tri-axial concrete stress field is assumed to develop, see Figure 3(a), where the chords carrying T
and C are also shown. It is assumed that the internal lever arm for the chords is approximately the same as the
height of the shear zone, i.e. z h. In the x,y-plane, the shear zone carries a compressive normal stress, fcc,
inclined at the angle with the beam axis, and a compressive normal stress, con, acting perpendicular to fcc.
These stresses are superimposed by the above-mentioned confinement pressure,fy, acting in direction of the
z-axis.
Figure 3. Triaxial stress field in circular beam subjected to constant shear (adapted from [7-9]).
Free body diagrams of the shear zone, the tension and compression chords, the load introduction diaphragm
and the four circular segments are shown in Figure 3(c – d). As illustrated in the free body diagram for the shear
zone, the boundaries are acted upon by the constant shear stress, , which from equilibrium conditions is given
by = V/bwh V/bwz. Furthermore, the horizontal boundaries also have to transfer the above-mentioned normal
pressure fy while the vertical boundaries parallel to the y,z-plane carry the compressive normal stress |x|. For
any chosen value of , the available equilibrium conditions can be used to establish relations that involve the
quantities , fcc, con, fy, C and T, see [7-9] for details. Furthermore, by applying the classical Modified Coulomb
failure criteria, see e.g. [9], and by requiring fcc ≥ fy ≥ con, the following relationship between fcc and con can be
established:
f cc f c 4 con (2)
where fcc can be identified as the tri-axial crushing strength of the concrete and fc is the effective uniaxial
compression strength of concrete (as known from the shear provisions for rectangular beams).
The assumed stress field will lead to a lower bound for the shear capacity of the beam, where failure is
governed by crushing of the concrete core and yielding of the stirrups. This lower bound solution is only valid if
the chords are able to carry C and T. For the sake of determining the maximum shear capacity, it will in the
L. C. Hoang 24.05.2021
For citations, copying and referencing see Note on Front page! page 321
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.2.3 (9) Background to FprEN 1992-1-1: 2023
following be assumed that the beam is not critical to bending failure, i.e. C and T can be carried.
To determine the maximum shear capacity, optimisation of the stress field with respect to the angle as well as
the stress con must be performed. The detailed calculations may be found in [7-9]. The optimisation leads to the
following results:
1
1 ; (with con 0) (a)
5 (3)
u
fc
1 3 1
; (with con (5 f c f c ) / 8) (b)
4 4 5
/
; 1/ 5
1 / (4)
tan
1
; 1/ 5
2
where u = Vu/bwz represents the maximum average shear stress that the shear cross section can carry. As can
be seen, for shear reinforcement content up to /0.20, it is optimal for the beam to carry the shear action
without the need of the in-plane confinement stress, i.e. con = 0. Solution (3a) is therefore identical to the so-
called web crushing criterion for shear in rectangular members, which is obtained by Expressions (8.44) and
(8.45) in prEN-1992-1-1: 2021. The web crushing criterion has been illustrated in Figure 4 as the dotted line,
which coincides with solution (3a). As indicated in the figure, for /0.20 the circular beam carries the shear
force by means of a biaxial compression field (fcc in the x,y-plane and fy in the direction of the z-axis) while the
tri-axial strength is only beneficial for /0.20. Comparison of solution (3) with test results with circular
members can be found in [7]. It should be noted that solution (3) is also valid when the entire circular area
within Dh is utilised to carry . This, however, requires that the bending moment must be carried solely by the
longitudinal reinforcement (both in compression as well as tension) and that all the longtidinal bars are engaged.
In this case, the lower bound solution will also be an exact solution as it coincides with the upper bound solution.
Detailed discussions of this exact solution can be found in [8].
f c cot (5)
L. C. Hoang 24.05.2021
For citations, copying and referencing see Note on Front page! page 322
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.2.3 (9) Background to FprEN 1992-1-1: 2023
x cot (7)
By inserting as defined in Equation (1), Equation (5) takes the following form:
2 Ash (8)
f y cot
Dh s
Furthermore, by adopting the definition of the shear reinforcement area, Asw, according to FprEN-1992-1-
1:2023, i.e. Asw = 2Ash, Equation (8) can be rewritten as:
bw Asw (9)
f y cot
Dh bw s
It can be seen that the term within the bracket is identical to Expression (8.42) in FprEN-1992-1-1:2023, i.e. the
shear capacity of beams with rectangular web in case where the transverse stirrups are rectangular and at yield.
Hence, according to Equation (9), the shear capacity of circular beams reinforced with circular stirrups can be
determined by means of Expression (8.42) in FprEN-1992-1-1:2023, provided that the area Asw is multiplied by
the ratio bw/Dh and provided that the nominal shear stress is taken as = V/zbw, where bw can be chosen freely,
however not greater than Dh (see Figure 6). Equation (9) is thus the theoretical background for the parts of
clause 8.2.3(9) in FprEN-1992-1-1:2023, which deal with members with circular cross section. As noted in this
clause, the internal lever arm, z, is based on a cross section fitted into the circular cross section, where the
tension bars within bw define the tension chord. Fulfilment of this clause would correspond to the fulfilment of the
assumptions made, when developing the lower bound model as explained in the above.
Figure 6. Effective cross section, where the tension bars within the chosen width bw define the tension chord.
It appears that Equation (6) is identical to Expression (8.44) in FprEN-1992-1-1:2023. Verification of the stress
in the compression field for members with circular cross section can therefore be performed as for rectangular
members.
Finally, according to Equation (7), the compression field in the shear zone results in a compressive axial force,
which due to equilibrium must be outbalanced by tensile axial forces in the chords. Hence, by multiplying both
sides of Equation (7) with the cross section area of the shear zone, zbw, the following expression for the
additional tensile force due to shear is obtained:
N v V cot (10)
which is identical to Expression (8.50) in FprEN-1992-1-1:2023. This force can be added to the chords in the
same manner as described in clause 8.2.3(8) of FprEN-1992-1-1:2023.
L. C. Hoang 24.05.2021
For citations, copying and referencing see Note on Front page! page 323
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.2.3 (9) Background to FprEN 1992-1-1: 2023
References
[1] Ang YB, Priestley MJN and Paulay T. Seismic shear design of circular reinforced concrete columns.
ACI Structural Journal, Vol. 86, No. 1, 54-59, 1989.
[2] Schwartz J. Stress field design of reinforced concrete members with circular cross section. Structural
Engineering International, Vol. 12, No. 1, 36-39, 2002.
[3] Kim JH, Mander JB. Theoretical shear strength of concrete due to transverse steel. Journal of
Structural Engineering, Vol. 131, No. 1, 197-199, 2005.
[4] Kim JH, Mander JB. Closure of theoretical shear strength of concrete due to transverse steel. Journal
of Structural Engineering, Vol. 132, No. 7, 1178-1179, 2006.
[5] Turmo J, Ramos G, Aparicio AC. Shear truss analogy for concrete members of solid and hollow
circular cross section. Engineering Structures, Vol. 31, No. 2, 455-465, 2009.
[6] Orr JJ. Darby AP, Ibell TJ, Denton SR, Shave JD. Shear design of circular cross sections using the
EC2 truss model. The Structural Engineer, Vol. 88, No. 23/24, 26-33, 2010.
[7] Hoang LC, Jensen UG. Rigid plastic solutions for the maximum shear capacity of confined RC
members. Magazine of Concrete Research, Vol. 62, No. 9, 625-636, 2010.
[8] Jensen, UG: Limit analysis of reinforced concrete bridge substructures – Shear strength of piers and
piles with circular cross section and strength of pile caps. PhD thesis, Inst. of Technology and
Innovation, University of Southern Denmark, 242 pp., 2010.
[9] Nielsen, MP, and Hoang, LC: Limit Analysis and Concrete Plasticity. Third edition, CRC Press, Boca
Raton, FL, 796 pp., 2011.
L. C. Hoang 24.05.2021
For citations, copying and referencing see Note on Front page! page 324
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.2.6 Background to FprEN 1992-1-1:2023
01.03.2023
Introduction
This document provides the background to subsection 8.2.6 (Shear at interfaces).
8.2.6(5) Design shear resistance without reinforcement or with sufficiently anchored reinforcement
The design shear resistance is defined complying with current EN 1992-1-1. An introductory clause is added to
clarify the use of the design expression. Compared to current EN 1992-1-1, the adhesion term of Eq. (8.76) is
expressed as a function of √fck/C instead of fctd to simplify the design procedure:
Such approach requires full anchorage of the reinforcement crossing the interface. The comparison between the
revised (FprEN 1992-1-1) and current (EN 1992-1-1) design shear resistance without interface shear
reinforcement is shown in Figure C8.3. As a result of the definition of fctd with respect to fck, the ratio of Rdi
according to FprEN 1992-1-1 and EN 1992-1-1 varies with varying concrete strength fck. While both expressions
yield similar results for fck ≈ 30 MPa, larger differences can be observed for lower and higher concrete strengths.
Nevertheless, as the differences are moderate (between +10% and -8%) for common concrete strength classes
(C20/25 – C60/70), this simplification seems justified.
4,0
very smooth EN 1992‐1‐1
smooth prEN 1992‐1‐1
3,0
τRdi (adhesion) [MPa]
rough
very rough
2,0 keyed
1,0
0,0
0 20 40 60 80 100 120
fck [MPa]
2,0
+10%
1,0
‐8%
0,5
0,0
0 20 40 60 80 100 120
fck [MPa]
Figure C8.3: Comparison between revised (FprEN 1992-1-1) and current (EN 1992-1-1) design shear
resistance without shear reinforcement (Eq. 8.76)
The shear reinforcement may be staggered as shown in Figure 8.15d. The lower limitation for the inclination of
interface reinforcement was extended to 35°≤ ≤ 135° (current EN 1992-1-1: 45°≤ ≤ 90°) for all surfaces
except for very smooth surfaces where 35°≤ ≤ 90°, since the crack opening is not relevant for reinforcement
tension. According to product standards and due to experience, lattice girders may be considered also for
smooth surfaces. The angle limitations are consistent with EN 13747:2010. However, for shear reinforcement,
the current limits (45°≤ ≤ 90°) still apply.
achieved in case of limited and insufficient anchorage length in the old and / or new concrete layer (e.g.
concrete toppings). For these cases, reduced anchorage requirements apply and the resisting shear stress at
the interface can be described by a combination of clamping effect and dowel action of the reinforcement
perpendicular to the interface. In the revised subsection, these effects are considered by the following
expression:
where cv2, v, kv, and kdowel are factors depending on the roughness of the interface, fck is the smallest strength of
the concretes at the interface, n is the stress caused by the minimum external normal force across the interface
that acts simultaneously (permanent stresses caused by confinement of surrounding structural parts may be
taken into account), is the degree of interface reinforcement, fyd is the design value of the yield strength of the
interface reinforcement, and fcd is the design value of the concrete compressive strength.
Again, the upper limitation of Eq. 8.77 is required according to RANDL [C8.1].
The new approach follows up on the work by RANDL AND WICKE and the background of the design expression is
described in detail in [C8.1], [C8.2], and [C8.3]. Interface shear design following this approach may result first of
all in somewhat higher degrees of interface reinforcement. This is however in many situations compensated by
the following two accompanying clauses as per FprEN 1992-1-1:2023 :
- In the case of horizontal shear transfer in slab members with cast-in-place structural toppings and
rough or very rough interfaces, the coefficient cv2 may be increased by a factor of 1,2 for determining
the design value τRdi of the interface shear resistance.
- The interface reinforcement should be anchored for a stress of at least 0,5fyd with a minimum length of
embedment of 8φ if no other methods of anchorage than by straight bars are applied.
This clause leads therefore to a reduced anchorage length in comparison to Eq. 8.76 and thus allows for shorter
reinforcing bars.
8.2.6(9) Detailing
Compared to current EN 1992-1-1, additional spacing rules are defined in clause 8.2.6(9) if interface
reinforcement is required in composite slabs to ensure sufficient detailing. In this context, the spacing of the
bars crossing the interface shall not exceed 2.5h ≤ 300 mm in the direction of shear transfer and 5h ≤ 750 mm
(≤ 375 mm) perpendicular to the direction of shear transfer (with h being the slab depth). The provisions were
derived by FURCHE AND BAUERMEISTER [C8.4] and were also incorporated in current German Annex of EN 1992-
1-1.
To avoid delamination of concrete toppings along free edges, an additional minimum interface reinforcement is
defined. The minimum reinforcement as,min shall be calculated as:
where tmin is the smaller value of the thickness of both concrete layers, fctm is the mean value of the concrete
tensile strength of the respective concrete layer, and fyk is the characteristic value of the yield strength of the
interface reinforcement. Such reinforcement along edges of concrete layers covers peeling forces (tensile forces
perpendicular to the interface) resulting from differential shrinkage or temperature effects, moreover adhesive
bond tends to be somewhat weaker towards edge regions than inside the member. The approach is based on
an 45°-inclined compressive strut equilibrating the horizontal forces along the edge; with its vertical component
to be anchored by as,min.
There exist however various solutions to cope with the peeling forces along such edges, even structural
solutions are possible (Fig. C8.26).
References [C8.1] Randl, N., Design recommendations for interface shear transfer in fib Model Code 2010, Structural
Concrete, Vol. 14, No. 3, 2013, pp. 230-241.
[C8.2] Randl, N., Tragverhalten einbetonierter Scherbolzen, Beton- und Stahlbetonbau, Vol. 100, No. 6,
Background Document to
Subsection 8.3: Torsion and Combined
Actions
08.02.2023
For citations, copying and referencing see Note on Front page! page 332
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.3 Background to FprEN 1992-1-1: 2023
Introduction
This document has been prepared to clarify the background for the provisions and changes introduced in
subsection 8.3 of FprEN-1992-1-1:2023. The subsection provides methods for ULS design of members
subjected to torsion as well as combinations of torsion, bending, shear and normal forces.
Figure 1: Roadmap to link the various provisions in Section 8 with Annex G. Topics written in red font are new
and not included in EN 1992-1-1:2004.
Subsection 8.3.1
The navigation clauses in EN 1992-1-1:2004 have been reworded and slightly changed. Compared to EN 1992-
1-1:2004, clause 8.3.1(3) allows the distribution of the acting torsional moments over the sub-sections in
proportion to their uncracked torsional stiffnesses or in proportion to their maximum possible torsional capacity.
Both methods give similar result but the distribution in proportion to the maximum strength of each sub-wall is
much easier for design.
Subsection 8.3.4
Subsection 8.3.4 presents the complete set of torsional capacity equations, where the equation to determine
t,Rd,sw is missing in EN 1992-1-1:2004. Note that the equations are formulated as capacity equations (and not
design equations) to make it easier when dealing verification of members subjected to combined actions. The
capacity has been formulated in terms of a shear stress to be consistent with subsection 8.3.2 and Annex G.
L. C. Hoang 18.05.2021
For citations, copying and referencing see Note on Front page! page 333
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.3 Background to FprEN 1992-1-1: 2023
To avoid the risk of overestimation of the torsional capacity due to partial spalling of the concrete cover prior to
the ultimate state (spalling of corners), clause 8.3.4(5) introduces a maximum nominal cover, c, to be used
when determining the thin-walled cross section that carries torsion. The well-known negative effect of corner
spalling on the torsional capacity is most pronounced for square cross sections. Explanation of this
phenomenon can be found in e.g. [1] while an extensive comparison of test results with calculations
demonstrating this effect can be found in [2].
Subsection 8.3.5
The provisions in EN-1992-1-1:2004 for combined actions are incomplete and have too narrow scope. These
provisions can for instance only be used for case (a) shown in Figure 2, but not for case (b). Even in case (a),
the provisions of EN-1992-1-1:2004 do not provide a method to verify the compression field in all of the four thin
walls, but only a check of the two walls carrying V and T. This can be seen in Figure 3, which shows the
procedure to verify the compression field in the left sub-wall for a cross section subjected to shear and torsion.
The calculation of cd follows the provisions of subsection 8.3.2 and Annex G. As stated in the figure, the check
is identical to Eq. (6.29) in EN-1992-1-1:2004. However, a similar verification of the bottom wall should also be
performed, since the -factor at this position may be significantly reduced due to large tension strains generated
by M and T. Such a verification is not included in Eq. (6.29) of EN-1992-1-1:2004.
Figure 2: Square cross sections subjected to combinations of shear, bending and torsion (V,M,T).
Figure 3: Procedure to verify the compression field in a wall of a cross section subjected to combinations of
bending, shear and torsion.
To enable design and capacity verification in the general case with combinations of biaxial bending, biaxial
shear, torsion as well as normal forces (My, Mz, Vy, Vz, T, N), clause 8.3.5(2) allows the designer to use
equilibrium methods to determine stress distributions, which are statically equivalent to the sectional forces,
without the need to fulfill compatibility conditions. The stresses can thereafter be used to design each thin-
walled part by applying Annex G. As an example of this concept, which is grounded on the lower bound
theorem of limit analysis, Figure 4 shows simple distributions of normal- and shear stresses, which are statically
equivalent to the actions (My, Mz, Vy, Vz,, T). Note that the shown stress distributions are just one of many
admissible sets of stress distribution. A linear elastic calculation with fulfilment of compatibility conditions would
lead to the well-known Navier- and Grashof distribution for normal stresses and shear stresses, respectively.
L. C. Hoang 18.05.2021
For citations, copying and referencing see Note on Front page! page 334
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.3 Background to FprEN 1992-1-1: 2023
Figure 4. Example of a set of normal- and shear stress distributions that is statically equivalent with the
sectional forces.
Subsection 8.3.6
The general method given in clause 8.3.5(2) is especially useful for the design of members in civil
engineering structures such as box bridge girders. For members, where torsion is not dominant, it will in
most cases be sufficient to adopt a simpler and conservative approach. For this reason, FprEN-1992-1-
1:2023 also includes a simple linear interaction formula for verification. The following formula is given in
clause 8.3.6(1),
S Ed
S 1
(1)
Rd i
where SEd and SRd represent the individual design action and, respectively, the corresponding individual
design resistance of the cross section. The linearized interaction formula has been illustrated in Figure
5 for combinations of shear, bending and torsion. The formula is conservative since yield conditions in
limit analysis are in general convex. As noted in clause 8.3.6(2), the interaction between shear action
and bending action do not need to be verified, provided that the bending capacity is based on an area
L. C. Hoang 18.05.2021
For citations, copying and referencing see Note on Front page! page 335
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.3 Background to FprEN 1992-1-1: 2023
of reinforcement, that does not include the portion of the reinforcement reserved to carry the additional
tension, Nvd, generated by the shear action. In this case, the interaction formula illustrated in Figure 5 is
split into the following two criteria:
M Ed TEd (2)
1
M Rd TRd
It appears that Equation (3) is similar to Eq. (6.29) in EN-1992-1-1:2004. The linear interaction formula
in clause 8.3.6(2) therefore contains the interaction rule in EN-1992-1-1:2004 as a special case.
Finally, it is noted that the fully linearized interaction formula can be replaced by other approximate
interaction diagrams developed on the basis of the provisions of clause 8.3.5(2). This is stated in clause
8.3.6(3). This clause basically allows the designer to use more accurate interaction models in
combination with linearization. The idea is illustrated in Figure 6 showing a (M,T,N)-interaction diagram,
where the knowledge of the accurate (M,N)-diagram can be used when adding a linear interaction with
torsion.
Figure 5. Illustration of linear interaction diagram for combinations of bending, shear and torsion.
Figure 6. Example of an alternative interaction diagram for combinations of bending, axial compression
and torsion, allowed according to clause 8.3.6(3).
L. C. Hoang 18.05.2021
For citations, copying and referencing see Note on Front page! page 336
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.3 Background to FprEN 1992-1-1: 2023
References
[1] Nielsen, MP, and Hoang, LC: Limit Analysis and Concrete Plasticity. Third edition, CRC Press, Boca
Raton, FL, 796 pp.
[2] Feddersen, B: Jernbetonbjælkers bæreevne -rene og kombinerede påvirkninger [Load-bearing
capacity of reinforced concrete beams subjected to pure and combined actions], PhD thesis,
Department of Structural Engineering, Technical University of Denmark, Report No. R-251 and R-
252, 1990.
L. C. Hoang 18.05.2021
For citations, copying and referencing see Note on Front page! page 337
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
For citations, copying and referencing see Note on Front page! page 338
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.4 Background to FprEN 1992-1-1:2023
Introduction
This document, which is a concise version of reference [C8.4.37], provides the background of the provisions
presented in Clause 8.4 of FprEN 1992-1-1:2023 for punching shear design. The main formulae for the punching
shear resistance are derived from the fib MC2010 [C8.4.1] and are based on the Critical Shear Crack Theory
(CSCT) [C8.4.2,C8.4.3,C8.4.4,C8.4.5,C8.4.6, C8.4.7,C8.4.8,C8.4.9,C8.4.11,C8.4.12].
This document introduces first the reasons for changing the current design method (EN 1992-1-1:2004) and briefly
presents the theoretical background of the proposed approach. Thereafter, the newly introduced design formulae
in Clause 8.4 of FprEN 1992-1-1:2023 are presented, justified and compared with experimental results.
Theoretical background
The new provisions for punching shear design of FprEN 1992-1-1:2023 are based on fib Model Code 2010
[C8.4.1], which has a pre-normative character. A detailed description of their theoretical bases can be consulted
in references [C8.4.4,C8.4.5,C8.4.6,C8.4.8,C8.4.9,C8.4.11,C8.4.12]. For further details on the provisions for
punching shear of fib Model Code 2010, please refer to reference [C8.4.7]. In the following, only the main
equations grounding the analytical derivation of the expressions included in the FprEN 1992-1-1:2023 are briefly
presented.
Figure C8.4.1: (a) Calculation of punching shear resistance and rotation at failure, with different possible punching
shear failure regimes (figures from [C8.4.8]); (b) experimental validation of the hyperbolic failure criterion of CSCT
(Eq. (C8.4.2)) proposed by Muttoni in 2008 [C8.4.4] .
Figure C8.4.2: Failure criteria associated to the different failure modes (Eqs. (C8.4.3) to (C8.4.5)) and intersection
with load-rotation relationship (figure adapted from [C8.4.11]; failure modes adapted from [C8.4.5])
formula specifically derived from the general law (based on a quadri-linear moment curvature relationship) for
practical purposes (detailed analytical derivation in [C8.4.7]) was retained:
𝜓 𝑘 (C8.4.1)
where ap refers to the distance between the axis of the supported area and the line of zero radial moment, d is
the effective depth, fy and Es are respectively the yield strength and the modulus of elasticity of the flexural
reinforcement, km is a factor depending on the level of the refinement of the approach used to estimate the acting
bending moment in the support width, ms is the acting bending moment in the support strip width (bs) and mR is
the average moment capacity in the support width. According to the fib Model Code 2010 [C8.4.1], the acting
moment in the support strip width mS can be calculated using a linear Finite Element Analysis or, in a first step,
can be estimated as a function of the acting shear force.
where VR,c is the punching shear resistance (concrete contribution), ke is the reduction factor (≤1) accounting for
the concentration of shear forces along the control perimeter [C8.4.6], b0,5 is the control perimeter at a distance
of dv/2 from the column face (round corners in case of square or rectangular columns), fc refer to the cylinder’s
concrete compressive strength and dg0 is the reference aggregate size (dg0 = 16 mm for normal weight concrete).
Intensive research carried out in the last 20 years has shown that Eq. (C8.4.2) has a very general scope and it
can be used to calculate the punching shear resistance in a general manner [C8.4.6] (see also Figure C8.4.1(b)).
One interesting advantage for the practical use of the CSCT is that the slab’s load-rotation relationship (which
results mostly from the slab flexural curvatures and thus depends on the load level) can be calculated in different
manners as described above.
criterion to be used for failures outside the shear-reinforced area is therefore similar to the one used to calculate
the punching resistance of slabs without shear reinforcement:
, .
⋅ (C8.4.4)
, ⋅ , , ⋅ , ⋅
where b0,5,out refers to the outer control perimeter (defined at 0.5dv,out from the last perimeter of shear reinforcement
and accounting for some limitations in the effective perimeter between the shear reinforcement units), dv,out is the
outer shear-resisting effective depth (thus accounting for the development of a horizontal crack along the
compression reinforcement that reduces the shear-resisting effective depth (see [C8.4.5] for details) and ke,out is
the factor accounting for the concentration of shear forces along the outer control perimeter. Conservatively, all
rotations are assumed to develop outside of the shear-reinforced area according to [C8.4.5].
𝜎 𝜓 𝑓 𝑓 (C8.4.7)
where fb is the average bond stress, w and fyw are respectively the diameter and the yielding strength of the shear
reinforcement. For members governed by shear deformations (as foundations or prestressed slabs)
[C8.4.31,C8.4.35], the maximum punching resistance can be found for large shear deformations with the concrete
contribution cancelling (Vc→0) and the stress in the shear reinforcement tending to the yielding strength (σs→fy;
shear strength governed by yielding of shear reinforcement).
Different effects due to the introduction of prestressing influence the punching resistance [e.g., C8.4.34]: (i)
introduction of in-plane forces; (ii) moments due to eccentric prestressing cables; (iii) deviation of forces in the
shear-critical region. As discussed by Clément et al. [C8.4.34,C8.4.35], the fib Model Code 2010 [C8.4.1] and the
CSCT account for all these effects.
As previously shown, the deformation and load carrying capacities can be calculated in the framework of the
CSCT by equalling the shear demand to the shear strength [C8.4.4]. This procedure consists on intersecting the
load-rotation relationship (representing the shear demand) with the failure criterion (corresponding to the shear
capacity). A simplified load-rotation relationship was proposed in the framework of the CSCT to account for the
effects of prestressing (eccentricity of the tendons and in-plane stresses) as follows [C8.4.34] (this formula is an
extension of a previous work of Muttoni [C8.4.4] for slabs without prestressing (Eq. (C8.4.1))):
𝜓 𝑘 (C8.4.8)
where mD is the decompression moment, calculated according to [C8.4.34] (mD positive if it leads to compressive
stresses on the top face of the slab):
𝑚 𝑛 𝑒 (C8.4.9)
where n represents the axial force (σdꞏh, with h referring to the height of the slab) and e is the eccentricity of the
normal forces. The punching shear design of prestressed slabs according to the fib Model Code 2010 [C8.4.1]
and SIA 262:2013 [C8.4.36] (which are based on the CSCT) consist on the application of Eq. (C8.4.8) (with
km=1.5) and an hyperbolic failure criterion similar to Eq. (C8.4.2) (with characteristic values).
A refined calculation of the punching shear resistance of prestressed slabs in the framework of the CSCT was
also proposed by Clément et al. [C8.4.35]. By considering the effects of the prestressing as external forces, the
mentioned authors proposed the calculation of the load-rotation relationship of the slab based on compatibility
and equilibrium conditions of a slab sector and by adopting a multi-linear moment-curvature relationship for the
sectional response (further details on the calculation of the load-rotation relationship may be found in [C8.4.35]).
8.4.3 (1) through (3) Derivation of closed-form formulae for punching shear design of flat slabs and foundation
Formulae (8.92) to (8.98) slabs without shear reinforcement
General considerations
To enhance the ease-of-use while applying the fib Model Code 2010 [C8.4.1] punching shear provisions (based
on the strain-based approach of the CSCT) in the FprEN 1992-1-1:2023, it was decided to keep the consistency
with the format of the current EN 1992-1-1:2004 and to formulate the punching resistance in terms of closed-form
solutions instead of intersecting the failure criteria and load-rotation relationships (strain-based approach). In
addition, it was further decided:
to use systematically the shear-resisting effective depth dv, instead of distinguishing between the flexural
effective depth (d, theoretically to be used in the load-rotation relationship) and the shear-resisting
effective depth (dv, governing for the location of the control section, length of the control perimeter and
calculation of the acting shear stress E)
to consider the effect associated with the concentration of shear forces along the control perimeter on the
action side (increasing the acting shear stress by a factor βe>1; similarly to the approach of the current EN
1992-1-1:2004), rather than including it on the resistance side (by reducing the control perimeter by a
factor ke<1) as included in the fib Model Code 2010 [C8.4.1].
The closed-form solution is first presented in a general format (the analytical derivation can be found in references
[C8.4.11,C8.4.37]) and simplified thereafter based on the two simplifications stated above (d=dv and ke=1). A
discussion on the impact of such simplifications on the derived equations and results is presented elsewhere
[C8.4.11,C8.4.37].
𝜓 𝑘 ⋅ 8 (C8.4.10)
Failure criterion associated to the concrete contribution adopted for the derivation of
provisions of Section 8.4
Closed-form solutions for the punching resistance cannot be derived using a hyperbolic failure criterion. To that
aim, Muttoni et al. [C8.4.8, C8.4.38] showed that the hyperbolic failure criterion of Eq. (C8.4.2) can be replaced
by a power-law equation with very similar results (see comparison in Figure C8.4.3):
where ddg represents the reference value of roughness of the critical shear crack and is computed as follows
[C8.4.8, C8.4.39]:
Figure C8.4.3: Comparison of original hyperbolic failure criterion (Eq. (C8.4.2)) with failure criterion based on a
power law (Eq. (C8.4.11)) (figure adapted from [C8.4.8]).
,
4 𝑎 100𝜌 𝑓 0.50 𝑓 (C8.4.13)
, ⋅ , ⋅
Where:
a = Vflex/mR;
. .
mR is the flexural resistance which is approximated as 𝑚 0.75𝑑 𝜌 . 𝑓 𝑓 [C8.4.38, C8.4.8];
. .
𝜏 , 𝑘 100𝜌 𝑓 𝑓 (C8.4.14)
where ddg is calculated according to Eq. (C8.4.12), l is the longitudinal flexural reinforcement ratio (with 𝜌
𝜌 , ⋅ 𝜌 , , where subscripts x,y refer to two orthogonal directions) and Rd,c refers to the design punching shear
stress (to be compared with the acting shear stress Ed=e·VEd/(b0,5·dv) considering the effect associated to the
concentration of shear forces on the action side). With respect to coefficient kpb in Eq. (C8.4.14), it accounts for
the strength enhancement due to the shear field gradient in the control section, and in the general format (i.e.,
distinguishing between d and dv, and considering ke) it is written as:
𝑘 𝑎 (C8.4.15)
. ,
𝑘 5 𝑎 5 𝜇 (C8.4.16)
, ,
where p accounts for the shear force-to-bending moment ratio in the shear-critical zone as well as for shear force
concentrations (considering that a=VEd/msd , where the value of msd is the acting moment computed in the width
of the support strip [e.g. C8.4.1, C8.4.4, C8.4.7], and accounting for the fact the factor ke considers the
concentration of shear forces along the control perimeter). The definition of factor p thus yields:
𝜇 𝑎 (C8.4.17)
For assessment of existing structures, the calculation of μp can be done in a tailored manner by using Eq.
(C8.4.17). Nevertheless, for design purposes, this formulation can be further simplified adopting the following
values of μp for the different cases (see [C8.4.1] for details):
p=8 for internal columns;
p=4 for edge columns and ends of walls;
p=2 for corner columns and corners of walls;
This factor p can also be expressed as a function of the two main control perimeters defined in FprEN 1992-1-
1:2023, namely b0 and b0,5, where b0 corresponds to the perimeter of the support region (perimeter at the column
edge minimised for re-entrant corners and columns near slab edges, see Figure 8.18 of FprEN 1992-1-1:2023)
and b0,5 is the length of the control perimeter for the shear verification located at a distance 0,5dv from b0.
𝜇 ⋅ ⋅𝑑 𝑏 , 𝑏 (C8.4.18)
Introducing Eq. (C8.4.18) in (C8.4.16), a simplified formula for kpb based on a geometrical rule can be derived as
follows (8.4.3(1) of FprEN1992-1-1:2023):
It can be noted that the shear-gradient enhancement factor kpb represents a smooth transition between one- and
two-way shear. The upper bound of kpb is required to limit the increase of shear resistance for very small supported
areas.
According to 8.4.3(2) of FprEN1992-1-1:2023, for distances between the axis of the supporting area and the
location of the zero radial bending moment smaller that ap≤8d, the punching resistance calculated according to
𝑎 𝑎 , ⋅𝑎 , 𝑑 (C8.4.20)
with ap referring to the location where the radial bending moment is equal to zero with respect to the support axis.
Finally, with respect to the parameter βe accounting for the concentration of shear forces due to moment transfer
between the slab and the column, it is now defined in FprEN 1992-1-1:2023 for wall supports and slab-column
connections. In the case of wall corners and ends of walls, parameter βe has been derived from the provision of
fib Model Code 2010 [C8.4.1] of eccentricity parameter ke, and is proposed as a constant value which is different
for ends of walls and wall corners. For columns, it can be estimated using an approximate approach, in which the
values for internal, edge and corner columns consist of a constant value, equal to the values in the current version
of EN 1992-1-1:2004, which are as well consistent with the provisions of fib Model Code 2010 [C8.4.1].
Additionally, in the case of slab-column connections, parameter βe can also be calculated using a more refined
procedure based on the work by Vollum and co-workers [C8.4.40]. In this work, an assessment of typical building
arrangements (in terms of span lengths, slab thickness and column size and type) was performed systematically
by means of Finite Element analyses. For each case, parameter βe was computed as the ratio between the shear
stress calculated using the Finite Element analysis (making use of shear fields) and the average shear stress
around the reduced control perimeter as per EN 1992-1-1:2004. The values obtained allowed the authors to
propose an expression for the calculation of parameter βe for corner columns, that was made extensible for inner
and edge connections and incorporated to FprEN 1992-1-1:2023, as shown in Eq. (C8.4.21). The proposal is as
well compared with current provisions of EN 1992-1-1:2004, showing that it is slightly more conservative at the
same time as it covers a wider range of scenarios than the current constant values.
𝛽 1 1.1 ⋅ 1.05 (C8.4.21)
where eb is the eccentricity of the resultant of shear forces with respect to the centroid of the control perimeter,
calculated for internal, edge and corner columns with Eqs. (C8.4.22a) through (C8.4.24c) respectively; and bb
represents the dimensions of the control perimeter and is calculated as the geometric mean of the minimum and
maximum overall widths of the control perimeter. With respect to the eccentricities, according to [C8.4.40], they
can be defined as follows:
𝑒 𝑒 𝑒 for internal columns (C8.4.22a)
Table C8.4.1: Summary of experimental-to-calculated punching shear resistance for the different investigated
cases of specimens without shear reinforcement.
Number of tests Approach Average COV [%]
The experimental-to-calculated resistance ratios are shown as a function of the main parameters governing the
punching shear resistance in Figures C8.4.4 to C8.4.6 for slabs and foundation slabs with concentric loading and
slabs with unbalanced moment, respectively. In the mentioned figures, the punching resistance was calculated
using Eq. (C8.4.14), still accounting for the favourable effect of the slenderness. It is important to note that no
noticeable trend can be clearly identified.
When the database of slabs without shear reinforcement is compared to EN 1992-1-1:2004, it is observed that
the accuracy of the punching resistance predictions is increased (in terms of coefficient of variation, 13.9% using
EN 1992-1-1:2004 compared to 11.2% for the same database) and no noticeable trends can be observed
(contrary to the case of EN 1992-1-1:2004, where, according to Kueres et al. [C8.4.19] using a different database,
trends are observed for the column size-to-effective depth ratio and depth of the member).
Figure C8.4.4: Experimental-to-calculated punching resistance ratio for a total of 121 specimens (database of
[C8.4.8]) representing slender slabs without shear reinforcement subjected to a concentric loading (punching
resistance calculated with Eq. (C8.4.14) accounting for the effect of slenderness replacing dv by (dvꞏap/8)1/2).
Figure C8.4.5: Experimental-to-calculated punching resistance ratio for a total of 34 specimens (database of
[C8.4.8]) representing foundation slabs without shear reinforcement subject to an uniform loading (punching
resistance calculated with Eq. (C8.4.14) accounting for the effect of slenderness replacing dv by (dvꞏap/8)1/2).
Figure C8.4.6: Experimental-to-calculated punching resistance ratio for a total of 47 specimens (database of
[C8.4.42]) representing slender slabs without shear reinforcement subject to a combined effect of shear forces
and unbalanced moment (factor e according to Eq. (C8.4.21) with eb=MR,test/VR,test; punching resistance calculated
with Eq. (C8.4.14) accounting for the effect of slenderness replacing dv by (dvꞏap/8)1/2).
8.4.3 (4) Punching shear resistance of prestressed slabs or slabs with axial forces
Formulae (8.99 to 8.103) Prestressing in slabs can influence the punching shear resistance due to three effects (see e.g. [C8.4.35]): (i)
deviation forces associated with prestressing tendons; (ii) moments due to eccentric prestressing forces and (iii)
compressive axial forces. The first effect is accounted for by calculating the acting shear forces at the control
perimeter (thus allowing to remove deviation forces inside of it, or considering the shear forces at the control
perimeter). The second effect is normally neglected in code formulations, while the third is often accounted for as
an increase of the unitary shear resistance as a function of the axial compressive stresses (see e.g., [C8.4.34] for
a discussion on the topic). In the fib Model Code 2010 [C8.4.1] and in the framework of the CSCT, the second
and third effects are accounted for by means of a decompression moment, which modifies the simplified load-
rotation relationship [C8.4.34, C8.4.35]. However, the analytical derivation of closed-form solutions for the
punching shear design of prestressed slabs combining the load-rotation relationship of Eq. (C8.4.8) (considering
the second and third previously mentioned effects) and the failure criterion of Eq. (C8.4.11) leads to complex
expressions that are not suitable for daily practice.
For that reason, the approach proposed in FprEN 1992-1-1:2023 results from an approximation of the results of
the mechanical model of the CSCT for slabs with in-plane forces [C8.4.35]. The effects of prestressing or axial
forces (compressive or tensile) are considered in the calculation of the punching shear resistance by multiplying
the value of kpb by the factor kpp calculated as follows:
𝑘 𝑘 if σd < 0 (compressive axial forces) (C8.4.23a)
. | |
𝑘 1 1 6 (C8.4.23)
,
where ep is the eccentricity of the tendons at the axis of the supported area with the respect to the centre of gravity
of the cross section (it is to be considered as positive for tendons on the tensile side and equal to zero for
prestressing forces centred with the cross section), and d is the effective depth in bending.
8.4.4 (6) and I.8.5.1 (1) Influence of compressive membrane action and slab continuity
Formula (I.15) Einpaul et al. [C8.4.43, C8.4.44] have shown that the slab continuity and compressive membrane action might
have a significant beneficial influence on the punching behaviour and strength of slab-column connections. These
effects are neglected in current EN 1992-1-1:2004 but have been accounted for in FprEN 1992-1-1:2023 for the
assessment of existing structures (Annex I). This is done consistently with the references [C8.4.43, C8.4.44,
C8.4.38, C8.4.8] where it is shown that the influence of slab continuity and compressive membrane action can be
analytically introduced in the closed form solution presented in Eq. (C8.4.13) and, therefore, in the design form
Eq. (C8.4.14).
The analytical derivation briefly presented in the following is grounded on the works presented by Muttoni et al.
[C8.4.38, C8.4.8], which can be consulted for further details. Einpaul et al. [C8.4.44] suggested a modification of
the load-rotation curve presented in Eq. (C8.4.1) to account for the stiffer behaviour (resulting from the
consideration of slab continuity) by multiplying the expression by a factor kcs. Analogously, Eq. (C8.4.10) can be
adapted to account for slab continuity as follows:
𝜓 𝑘 𝑘 8 (C8.4.24)
where kcs is a factor accounting for the influence of slab continuity and compressive membrane action. Based on
these works, Muttoni and Fernández Ruiz [C8.4.38] suggested a formulation for this factor which is a function of
the ratio mcr/mR (supported on the evidence that the confinement at the column region is provided by the
surrounding concrete during the crack development stage):
𝑘 0.08 1 (C8.4.25)
where mcr is the cracking moment per unit length (without considering an axial force). Eq. (C8.4.13) can therefore
be rewritten accounting for slab continuity and compressive membrane action as (following the same assumptions
used in the derivation of Eq. (C8.4.13) and considering ke=1 and d=dv as previously introduced):
,
4𝑎 100𝜌 𝑓 0.50 𝑓 (C8.4.26)
, ⋅ , ⋅ ⋅
It can therefore be seen that Eq. (C8.4.26) is analogous to Eq. (C8.4.13), considering that the value of ap is
corrected (reduced) to account for the slab continuity and compressive membrane action by the factor kpm=kcs2.
Considering that the flexural capacity per unit length mR can be approximated as 𝑚 0.75𝑑 𝜌 . 𝑓 . 𝑓 .
[C8.4.38, C8.4.8] and that the cracking moment per unit length can be computed as 𝑚 ℎ (considering
h=d/0.9), the factor kpm can be simplified as follows:
. . .
.
𝑘 𝑘 ≅ 0.08 ⋅ 6 ⋅ 1.0 (C8.4.27)
Simplifying the exponents (with minor impact) and considering fc≈30 MPa, fy≈500 MPa, ρ≈0.01, and fct=0.53ꞏfck1/2,
Eq. (C8.4.27) can written as:
This enhancement due to compressive membrane action can be expressed in terms of a coefficient that multiplies
factor kpb, thus improving the ease of use of the formulation. This coefficient, defined as ηpm, can be directly
obtained as the ratio between Equations (C8.4.26) and (C8.4.13):
, ,
𝜂 0.95 (C8.4.29)
, ⋅
The punching shear resistance of slabs in existing structures can therefore be calculated accounting for slab
continuity and compressive membrane action using Eq. (C8.4.14) with the parameter kpb being replaced by kpbꞏηpm
where the factor ηpm is calculated with Eq. (C8.4.29).
allows the calculation of the stress in the shear reinforcement as a function of a number of known
parameters.
(4) Using the descending branch of the power-law failure criterion of Eq. (C8.4.11), and still considering
the relationship VR,c/VE=ηc established in (3), it can be shown that the concrete contribution VRc,E
(associated to the rotation ψE corresponding to the shear force VE) is in fact given by VRc,E= VR,c ∙ ηc
Based on the steps (1) to (4) above presented (please refer to [C8.4.11,C8.4.37] for details), it is possible to show
that the punching shear resistance associated to failures within the shear-reinforced region can be calculated as
follows (already expressed in a design format) [C8.4.11,C8.4.37]:
𝜏 , 𝜂 ⋅𝜏 , 𝜂 ⋅𝜌 ⋅𝑓 𝜌 ⋅𝑓 (C8.4.30)
where
𝜂 15 0.8 is a factor accounting for the activated shear reinforcement stress (upper
⋅ ⋅
limit referring to both yielding and amount of shear reinforcement that can
be effectively activated, see Figure C8.4.7)
,
𝜂 is the concrete contribution reduction factor accounting for the increased
deformation of slabs with shear reinforcement at ultimate limit state with
respect to slabs without shear reinforcement, see Figure C8.4.7
It should be noted that the additive factor in Eq. (C8.4.30) can only be accounted for provided that the
⋅
shear reinforcement consists of ribbed or indented bars.
Figure C8.4.7: Definition of contributions of concrete and shear reinforcement for punching failure within the
shear-reinforced region (from [C8.4.11]).
where dsys is defined as in Figure I.3 of FprEN 1992-1-1:2023 and represents the anchorage performance of the
punching reinforcement system and its detailing, b0 is the perimeter of the supported region and represents the
size of the column, and s0 is the distance from the column face to the axis of the first perimeter of shear
reinforcement.
To increase the ease-of-use, Eq. (C8.4.32) has been simplified in Clause 8.4 for studs, links and stirrups fulfilling
the detailing rules of Clause 12.5.1:
It has to be noted that for this coefficient, two different symbols are used:
- sys is the coefficient which applies to the resistance calculated with the closed-form formulation
(clauses 8.4 and I.8.5.2(2)) whereas
- sys,sb is the coefficient which applies to the failure criterion in the strain-based approach (clause 8
I.8.5.2(3)). Since the resistance results from the intersection between the failure criterion and the load-
rotation relationship, the ratio between the maximum punching resistance and the resistance of a similar
slab without shear reinforcement is always smaller than sys,sb .
8.4.4 (7) and (8) Development of a closed-form formula for outer control perimeter
Formula (8.112) The punching shear resistance outside the shear reinforced region shall be calculated in accordance to Eq.
(C8.4.14), considering the reduced shear-resisting effective depth dv,out (function of the shear reinforcement
system) and the outer control perimeter b0,5,out, located at dv,out/2 from the outer perimeter of shear reinforcement
with a length of the straight segments not exceeding 3ꞏdv,out. This approach is consistent with the model that
grounds the expressions, thus considering all the rotation to concentrate outside of the shear-reinforced region.
The expression therefore is (following similar considerations as previously adopted for the derivation of Eq.
(C8.4.14)):
. .
𝜏 , , 𝑘 , 100𝜌 𝑓 𝑓 (C8.4.35)
⋅
𝑘 , 5 𝜇 , (C8.4.36)
. ,
where dv is the shear-resisting effective depth and dv,out the outer shear-resisting effective depth (accounting for
the possibility of delamination in compression surface). It should be noted that the dv in the parts of the Eq.
(C8.4.35) related to kpb and to the size- and strain-effect is not replaced by dv,out as the dv in those cases is
associated to the flexural effective depth (simplified to the shear-resisting effective depth as discussed above).
The resulting Eq. (C8.4.35) thus comes from the analytical derivation of Eq. (C8.4.13) introducing d=dv but not by
dv,out in kpb and strain- and size-effect parts. Such consideration is on the safe side as shown in reference [C8.4.37]
(see also d/dv ratio in Eq. (C8.4.15)), where a discussion on this issue is presented.
Eq. (C8.4.36) which can be further simplified in the same manner as kpb in Eq. (C8.4.19), i.e., assuming that:
𝜇 , ⋅ ⋅𝑑 , 𝑏 , , 𝑏 , (C8.4.37)
where b0,out is the control perimeter at the outermost perimeter of shear reinforcement, with a length of the straight
segments not exceeding 3ꞏdv,out. Introducing Eq. (C8.4.37) in (C8.4.36), a simplified formula for kpb,out based on a
geometrical rule can be derived as follows (clause 8.4.4(4) of FprEN1992-1-1:2023) in an analogous manner to
the previous derivation of kpb in Eq. (C8.4.19):
,
1 𝑘 , 3.6 1 2.5 (C8.4.38)
, , ,
Comparing Eq. (C8.4.38) for the calculation of kpb,out to Eq. (C8.4.19) for kpb, it results that an increase of kpb by
multiplying it by factor (dv/dv,out)1/2 is possible for the verification outside the shear-reinforced region (as allowed in
Clause 8.4.4(4)). The consideration of such increase on the factor kpb can be shown to be justified and on the safe
side (see Eq. (C8.4.15) and reference [C8.4.37] for further discussion on this topic).
One can derive a closed-form solution to calculate the minimum length of the outer control perimeter to avoid a
failure outside the shear-reinforced region by equalling VRd,c,out (calculated with Eq. (C8.4.35)) to VEd as follows
(all rotation considered to develop outside the shear-reinforced area and strength criterion considered not to be
governing):
𝑑 𝑏 , 𝑘 , 𝑑 𝑏 , 𝑏 ,
𝜏 𝜏 , ⇒𝜏 𝜏 ,
𝑑 , 𝑏 , , 𝑘 𝑑 , 𝑏 , , 𝑏 , ,
⇒𝑏 , , 𝑏 , (C8.4.39)
,
The experimental-to-calculated resistance ratios for members with shear reinforcement are shown as a function
of the main parameters governing the punching shear resistance in Figures C8.4.8 and C8.4.9 for slabs and
foundation slabs with concentric loading, respectively. Again, no noticeable trend can be clearly identified.
With respect to approach (3) in Table C8.4.2 (refined coefficient ηsys according to I.8.5.2(2) for the assessment of
existing structures), in the case of flat slabs with details not fulfilling the rules of Clause 12, no notable increase
of the accuracy compared to the simplified formula in Clause 8.4 can be observed (approach (2) in Table C8.4.2).
Nevertheless, the detailed formula according to I.8.5.2(2) is particularly useful for the assessment of slabs not
fulfilling anymore the detailing rules of Clause 12.
Figure C8.4.8: Experimental-to-calculated punching resistance ratio for a total of 135 specimens (dataset of
database of [C8.4.45] completed with the tests of [C8.4.46]) representing slender slabs with shear reinforcement
subjected to a concentric loading, using approach (2) as defined in Table C8.4.2: 4 tests of [C8.4.47]; 1 test of
[C8.4.48]; 1 test of [C8.4.49]; 6 tests of [C8.4.50]; 1 test of [C8.4.51]; 7 tests of [C8.4.52]; 7 tests of [C8.4.53]; 6
tests of [C8.4.54]; 2 tests of [C8.4.55]; 2 tests of [C8.4.56]; 5 tests of [C8.4.57]; 10 tests of [C8.4.58]; 3 tests of
[C8.4.59]; 8 tests of [C8.4.60]; 17 tests of [C8.4.61]; 9 tests of [C8.4.62]; 6 tests of [C8.4.63]; 2 tests of [C8.4.64];
1 test of [C8.4.65]; 8 tests of [C8.4.66]; 2 tests of [C8.4.67]; 5 tests of [C8.4.14]; 11 tests of [C8.4.28]; 11 tests of
[C8.4.46].
Figure C8.4.9: Experimental-to-calculated punching resistance ratio for a total of 12 specimens representing
foundation slabs with shear reinforcement using approach (2) as defined in Table C8.4.2: 3 tests of [C8.4.13]; 5
tests of [C8.4.69]; 4 tests of [C8.4.31].
References [C8.4.1] Fédération internationale du béton, fib Model Code for Concrete Structures 2010, Ernst & Sohn,
Germany, 2013, 434 p.
[C8.4.2] Muttoni, A., and Schwartz, J., Behaviour of Beams and Punching in Slabs without Shear
Reinforcement, IABSE Colloquium, V.62, Zürich, Switzerland, 1991, pp. 703-708
[C8.4.3] Muttoni A., Shear and Punching resistance of Slabs Without Shear Reinforcement (In German:
Schubfestigkeit und Durchstanzen von Platten ohne Querkraftbewehrung, Beton- und Stahlbetonbau),
Beton-und Stahlbetonbau, V.98, 2003, pp. 74-84
[C8.4.4] Muttoni A., Punching shear strength of reinforced concrete slabs without transverse reinforcement, ACI
Struct J, Vol.105, No.2, 2008, pp. 163-172
[C8.4.5] Fernández Ruiz M., Muttoni A., Applications of the critical shear crack theory to punching of R/C slabs
with transverse reinforcement, ACI Struct J., V.106, No.4, 2009, pp. 485-494
[C8.4.6] Muttoni A., Fernández Ruiz M., Shear in slabs and beams: should they be treated in the same way?,
fib bulletin 57, Lausanne, Switzerland, 2010, pp. 105-128
[C8.4.7] Muttoni A., Fernández Ruiz M., Bentz E., Foster S., Sigrist V., Background to fib Model Code 2010
shear provisions–part II: punching shear, Struct Conc., V.14, No.3, 2013, pp. 204-214
[C8.4.8] Muttoni A., Fernández Ruiz M., Simões J.T., The theoretical principles of the Critical Shear Crack
Theory for punching shear failures and derivation of consistent closed-form design expression, Struct
Concr., V. 19, 2018, pp. 174-190, https://doi.org/10.1002/suco.201700088
[C8.4.9] Simões J.T., Fernández Ruiz M., Muttoni A., Validation of the Critical Shear Crack Theory for punching
of slabs without transverse reinforcement by means of a refined mechanical model, Struct Conc, V. 19,
2018, pp. 191–216. https://doi.org/10.1002/suco.201700280
[C8.4.10] Muttoni A, Fernández Ruiz M., Simões J. T., Recent improvements of the Critical Shear Crack Theory
for punching shear design and its simplification for code provisions, Proceedings of 5th fib Symposium,
Melbourne, 2018, 14 p.
[C8.4.11] Muttoni A., Simões J.T., Faria D.M.V., Fernández Ruiz M., A mechanical approach for the punching
shear provisions in the second generation of Eurocode 2, Hormigón y Acero, 2023, doi:
https://doi.org/10.33586/hya.2022.3091
[C8.4.12] Muttoni A., Simões J. T., “Shear and punching shear according to the Critical Shear Crack Theory:
background, recent developments and integration in codes,” Rev. IBRACON Estrut. Mater. (submitted
for publication)
[C8.4.13] Hegger J., Ricker M., Sherif A. G., Punching resistance of Reinforced Concrete Footings, ACI Struct
J., V.106, No.5, 2009, pp. 706-716
[C8.4.14] Vollum R.L., Abdel-Fattah T., Eder M., Elghazouli A.Y., Design of ACI-type punching shear
reinforcement to Eurocode 2, Mag conc Research, V.62, No.1, 2010, pp. 3-16
[C8.4.15] Siburg C., Ricker M., Hegger J., Punching shear design of footings: critical review of different code
provisions, Struct Conc., V.15, 2014, pp. 497-508
[C8.4.16] Ricker M., Siburg C., Punching shear strength of flat slabs – critical review of Eurocode 2 and fib Model
Code 2010 design provisions, Struct Conc., V.17, 2016, pp. 457-468
[C8.4.17] Einpaul, J., Bujnak, J., Fernández Ruiz, M., Muttoni, A., Study on influence of column size and slab
slenderness on punching resistance, ACI Struct J., V.113, No.1, 2016, pp. 135-145
[C8.4.18] Einpaul J., Brantschen F., Fernández Ruiz M., Muttoni A., Performance of Punching Shear
Reinforcement under Gravity Loading: Influence of Type and Detailing, ACI Struct J., V.113, No.4,
2016, pp. 827-838
[C8.4.19] Kueres D., Siburg C., Herbrand M., Classen M., Hegger J., Uniform Design Method for punching shear
in flat slabs and column bases, Eng Struct., V.136, 2017, pp. 149-164
[C8.4.20] Fernández Ruiz M., Muttoni A., Size effect in shear and punching shear failures: differences between
statically determinate members and redundant structures, Struct Concr., 2017, pp. 1-11, DOI:
10.1002/suco.201700059
[C8.4.21] Dönmez A., Bazant Z.P., Size effect on punching resistance of reinforced concrete slabs with and
without shear reinforcement, ACI Struct J., V.114, No.4, 2017, pp. 875-886
[C8.4.22] Kinnunen S., Nylander H., Punching of Concrete Slabs Without Shear Reinforcement, Transactions of
the Royal Institute of Technology, No.158, 1960, 112 p.
[C8.4.23] Guandalini S., Symmetric punching shear of reinforced concrete slabs (In French), Doctoral Thesis,
EPFL, Lausanne, Switzerland, No.3380, 2005, 289 p.
[C8.4.24] Einpaul J., Punching resistance of Continuous Flat Slabs, Doctoral Thesis, EPFL, Lausanne,
Switzerland, No.6928, 2016, 209 p.
[C8.4.25] Belletti B., Walraven J.C., Trapani F., Evaluation of compressive membrane action effects on punching
shear resistance of reinforced concrete slabs, Eng Struct., V.95, 2015a, pp. 25-39
[C8.4.26] Belletti B., Pimentel M., Scolari M., Walraven J.C., Safety assessment of punching shear failure
according to the level of approximation approach, Struct Conc., V.16, No.3, 2015b, pp. 366-380
[C8.4.27] Beutel R., Hegger J., The effect of anchorage on the effectiveness of the shear reinforcement in the
punching zone, Cement and Concrete Composites, V.24, 2002, pp. 539-549
[C8.4.28] Lips S., Fernández Ruiz M., Muttoni A., Experimental Investigation on Punching resistance and
Deformation Capacity of Shear-Reinforced Slabs, ACI Struct J., V.109, 2012, pp. 889-900
[C8.4.29] Brantschen F., Faria D. M. V., Fernández Ruiz M., Muttoni A., Bond Behaviour of Straight, Hooked, U-
Shaped and Headed Bars in Cracked Concrete, Struct Conc., V.17, No. 5, 2016, pp. 799-810
[C8.4.30] Hernández Fraile D., Simões J. T., Fernández Ruiz M., Muttoni A., A mechanical approach for the
maximum punching resistance of shear-reinforced slab-column connections, Proceedings of the fib
Symposium, Lisbon, 2021, pp. 1628-1639.
[C8.4.31] Simões J.T., Bujnak J., Fernández Ruiz M., Muttoni A., Punching shear on compact footings with
uniform soil pressure, Structural Concrete, Struct C., No.4 , 2016
[C8.4.32] Muttoni A., Fernández Ruiz M., Guandalini S., Poinçonnement des ponts-dalles, ASTRA, 2007.
[C8.4.33] Clément T., Influence de la précontrainte sur la résistance au poinçonnement de dalles en béton armé,
Doctoral Thesis, EPFL, Lausanne, Switzerland, 2012, No. 5516, 222 p.
[C8.4.34] Clément T., Pinho Ramos A., Fernández Ruiz M., Muttoni A, Design for punching of prestressed
concrete slabs, Struct Conc., V.14, No. 2, 2013, pp. 157-167
[C8.4.35] Clément T., Pinho Ramos A., Fernández Ruiz M., Muttoni A., Influence of prestressing on the punching
resistance of post-tensioned slabs, Eng Struct., V.72, 2014, pp. 59-69
[C8.4.36] SIA. Code 262 for Concrete Structures. Zürich, Switzerland: Swiss Society of Engineers and Architects;
2013, 102 p.
[C8.4.37] Muttoni A., Simões J.T., Hernández Fraile D., The punching shear provisions of the 2º generation of
Eurocode 2: a discussion on the derivation of the formulae, adopted simplifications and their
consequences, under preparation, 2023
[C8.4.38] Muttoni A., Fernández Ruiz M., The Critical Shear Crack Theory for punching design: from a
mechanical model to closed-form design expressions, ACI/fib International Punching Shear
Symposium, Bulletin fib 81, 2007, pp. 237-252
[C8.4.39] Cavagnis F. Fernández Ruiz M., Muttoni A., A mechanical model for failures in shear of members
without transverse reinforcement based on development of a critical shear crack, Eng Struct, V. 157,
2018, pp. 300-315
[C8.4.40] Abu-Salma D., Vollum R. L., Maccorini L., Design of biaxially loaded external slab column connections,
Engineering Structures, V. 249, 2021, 16 p.
[C8.4.41] Muttoni A., FprEN 1992-1-1:2023 – Background document to clause 4.3.3 and Annex A – Partial safety
factors for materials, CEN/TC250/SC2/WG1/TG6, 2023, 21 p.
[C8.4.42] Tassinari L., Non symmetric punching of reinforced concrete slabs (In French : Poinçonnement
symétrique des dalles en béton armé avec armature de poinçonnement), Doctoral Thesis, EPFL, No.
5030, Lausanne, Switzerland, 2011, 197 p.
[C8.4.43] Einpaul J., Ospina C. E., Fernández Ruiz M., Muttoni A., Punching Shear Capacity of Continuous
Slabs, ACI Struct J., V. 113, No. 4, 2016, pp. 861-872
[C8.4.44] Einpaul J., Fernández Ruiz M., Muttoni A., Influence of moment redistribution and compressive
membrane action on punching resistance of flat slabs, Eng Struct., V.86, 2015, pp. 43-57
[C8.4.45] Walkner R., Critical review of EC 2 regarding punching and improving the design approach (In German:
Kritische Analyse des Durchstanznachweises nach EC2 und Verbesserung des
Bemessungsansatzes), Doctoral Thesis, Innsbruck University, 2014.
[C8.4.46] Ferreira M. P., Melo G. S., Regan P. E., Vollum R. L., Punching of Reinforced Concrete Flat Slabs with
Double-Headed Shear Reinforcement, ACI Struct J., 2014, pp. 363-374
[C8.4.47] Andersson J.L., Punching of Concrete Slabs with Shear Reinforcement, Transactions of the Royal
Institute of Technology, Stockholm, Sweden, 1963
[C8.4.48] Marti P., Pralong J., Thürlimann B., Schubversuche an Stahlbeton-Platten, Institut für Baustatik und
Konstruktion, No.7305-2, Zürich, Switzerland, 1977, 123 p.
[C8.4.49] Seible F., Ghali A., Dilger W. H, Preassembled shear reinforcing units for flat plates, ACI J Proc., V.77,
No.1, 1980, pp. 28-35
[C8.4.50] Van der Voet A.F., Dilger W.H., Ghali A., Concrete Flat Plates with Well-Anchored Shear
Reinforcement Elements, Canadian Journal of Civil Engineering, V.9, No.1, 1982, pp. 107-114
[C8.4.51] Müller F.X., Muttoni A., Thürlimann B., Punching shear tests on slabs with openings (In German:
Durchstanzversuche an Flachdecken mit Aussparungen), Institut für Baustatik und Konstruktion,
No.7305-5, Zürich, Switzerland, 1984, 117 p.
[C8.4.52] Mokhtar A.S., Ghali A., Dilger W.H., Stud Shear Reinforcement for Flat Concrete Plates, ACI J., No.82-
60, 1985, pp. 676-683
[C8.4.53] Chana P.S., Desai S.B., Design of shear reinforcement against punching, The Structural Engineering,
V.70, 1992, pp. 159-164
[C8.4.54] Yamada T., Nanni A., Endo K., Punching Shear Resistance of Flat Slabs: Influence of Reinforcement
Type and Ratio, ACI Struct J., V.88, 1992, pp. 555-563
[C8.4.55] Chana P.S., A prefabricated shear reinforcement system for flat slabs, Proceedings of the Institution of
Civil Engineers: Structures & Buildings, V.99, 1993, pp. 345-358
[C8.4.56] Ladner M., Schaeidt W., Gut W., Durchstanzversuche an Flachdeckenausschnitten, EMPA Report No.
419, 1996, 38 p.
[C8.4.57] Lee S.C., Lee S.B., Teng S., Morley C.T., Punching shear tests on high strength concrete slabs,
Proceedings of the 5th international symposium on utilisation of high strength-high performance
concrete, 1999, pp. 401-410
[C8.4.58] Gomes R.B., Regan P.E., Punching Resistance of RC Flat Slabs with Shear Reinforcement, ASCE
Journal of Struct Eng., V.125, No.6, 1999, pp. 684-692
[C8.4.59] Oliveira D.R., Melo G.S., Regan P.E., Punching resistances of Flat Plates with Vertical or Inclined
Stirrups, ACI Struct J., V. 97, 2000, pp. 485-491
[C8.4.60] Regan P.E., Samadian F., Shear Reinforcement against punching in reinforced concrete flat slabs, The
Structural Engineer, 2001, pp. 24-31
[C8.4.61] Beutel R., Durchstanzen schubbewehrter Flachdecken im Bereich von Innenstützen, RWTH, Aachen,
Germany, 2002, 267 p.
[C8.4.62] Birkle G., Punching of Flat Slabs: The Influence of Slab Thickness and Stud Layout, Doctoral Thesis,
University of Calgary, Calgary, Canada, 217 p., 2004
[C8.4.63] Rojek R., Keller T., Punching Tests using Reinforcement with Enhanced Bond (In German), Structural
Analysis and Design Approaches, Beton-und Stahlbetonbau, V.102, No.8, 2007, pp. 584-556, 2007
[C8.4.64] Broms C. E., Ductility of Flat Plates: Comparison of Shear Reinforcement Systems, ACI Struct J.,
V.104, No.6, 2007, pp. 703-711
[C8.4.65] Hegger J., Häusler F., Ricker M., Zur maximalen Durchstanztragfähigkeit von Flachdecken, Beton -
und Stahlbetonbau, V.102, 2007, pp.770-777
[C8.4.66] Feix J., Schustereder C., Durchstanzen nach EN 1992-1-1, Durchstanzversuche zur Festlegung der
Nachweisgrenzen, Bauingenieur, V.82, 2007, pp. 135-142
[C8.4.67] Etter S., Heinzmann D., Jäger T., Marti P., Versuche zum Durchstanzverhalten von Stahlbetonplatten,
IBK Bericht, 64, Zürich, Switzerland, 2009
[C8.4.68] Lips S., Muttoni A., Fernández Ruiz M., Punching of flat slabs: Design example, EPFL, Lausanne,
Switzerland, 2011, 13 p.
[C8.4.69] Siburg C., Hegger J., Experimental Investigations on Punching Behaviour of Reinforced Concrete
Footings with structural dimensions, Struct C., V.15, 2014, pp. 331-339
For citations, copying and referencing see Note on Front page! page 358
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
8.5 Background to FprEN 1992-1-1:2023
Introduction This document has been prepared to clarify the background of the design provisions of FprEN 1992-1-1:2023
with respect to Section 8.5 (“Design with strut-and-tie models and stress fields”). These provisions inherit the
approach of former Section 6.5 of EN 1992-1-1:2004, but updates its content and design clauses based on the
evolutions of the last 20 years. Also, the section has been reworked to have an enhanced consistency with
Section 8.2.3 (“Members requiring design shear reinforcement”) and Annex G (“Design of membrane-, shell-
and slab elements”) of FprEN 1992-1-1:2023.
Reasons for change Current provision for strut-and-tie models in EN 1992-1-1:2004 (clause 6.5) gave rules for the design of D-
regions (discontinuity regions “where a non-linear strain distribution exists” as stated in the code) based on the
strut-and-tie method. These provisions were to a large extent consistent with the vision by Schlaich and co-
workers [C8.5.1], that represented the state-of-the-art in the 1980’s. Since then, many advances in the
knowledge have been consolidated in this field, leading to a more comprehensive understanding of several
topics as:
i. analogies of strut-and-tie models with the stress field method [C8.5.2]
ii. explicit consideration of the strain state in the calculation of efficiency factors [C8.5.3]
iii. integration of strut-and-tie models and stress fields within automated design procedures and
consistency with design of B-regions [C8.5.4]
Such advances have shown limitations and deficiencies of the EN1992-1-1:2004 approach, advising to revise it.
A detailed review of current state-of-the-art in this domain has recently been completed by a dedicated group of
fib (WP2.2.4), whose results have been condensed within a specific Bulletin [C8.5.5]. The revision of the
FprEN1992-1-1:2023 provisions are to a large extent based on this knowledge and such document shall be
consulted for extended clarifications. The fib Bulletin [C8.5.5] can thus be considered as a more
comprehensive background document for clause 8.5 of FprEN 1992-1-1:2023.
8.5.1 General This clause has been revised and extended to clarify the definitions used and the analogy between strut-and-tie
models and stress fields. In general, the strut-and-tie models can be considered as the equilibrium models
where the location of the struts and ties represent the resultant of compression fields and reinforcement
respectively, see Fig. C8.5.1.
Figure C8.5.1: Analogy between: (a) stress fields; and (b) strut-and-tie models
8.5.2 Struts and Other than an explicit definition of the stress in the strut, this clause incorporates, with respect to EN 1992-1-
compression fields 1:2004, a variable value of the efficiency factor for the concrete strength depending on the cracking state of
the compression field:
Constant values for a compression field developing at a given angle with a tie (stressed
reinforcement, §8.5.2(4)). This approach is particularly suitable for a preliminary design, where the
exact value of the angle between struts and ties is not known a priori, but its range can be easily
estimated
Variable value for a compression field developing at a given angle with a tie or tension field (Formula.
(8.119)). This approach allows for a tailored calculation of the resistance, particularly for a more
refined calculation in a final check or automated design procedure
Local value of the efficiency factor based on its principal tensile strain, Formula (8.121). This
approach is mostly addressed for an automated calculation (as a FE stress field analysis or for the
use of Annex G)
The different approaches are compared in Figure C8.5.2. As it can be noted, the constant values are aimed at a
simple and safe estimate of the efficiency factor while Formula (8.119) provides in general higher estimates of
the strength.
Figure C8.5.2: Comparison of different approaches to calcualate the concrete efficiency factor
It is interesting to note that Formula (8.119) can be consistently derived from the local formulation of (Formula
(8.121)), by estimating the principal strain (1) assuming that the compressive strain in concrete is 0,001 and
that the strain tensor is parallel to the stress field (applies when one assumes that cracks, which have the same
inclination as the compression field, open without sliding):
1= x +( x + 0.001)∙cot2 (C.11.3.1)
where the principal strain value is calculated for a given level of horizontal strain ( x) and an inclination of the
compression field equal to (refer to Fig. C8.5.3). As it can be noted, this assumption is consistent with the
procedure for shear design presented in § 8.2.3(7) and Formula (8.45). In this case, for calculation of Formula
(8.119), it is considered that x ≈ 0.001 (horizontal strain at mid-height of a compression field where one side
reaches the yield strength and the deformations on the other are neglected, refer to Formulae. (8.47) - (8.49)).
In addition to these aspects, several considerations have been incorporated covering the contribution of
compression reinforcement (§8.5.2(7)) and also local enhancement of the compressive strength due to
confinement (§8.5.2(6))
8.5.3 Ties The clause has been reworked to enhance clarity and to clarify how to account for prestressing units.
8.5.4 Nodes Design for nodes has been integrated as part of the general verification of the compression fields reaching the
nodes (see example of Fig. C8.5.2).
The provisions for the nodal regions are consistent with current draft of Model Code 2020 (based on the work by
fib’s WP 2.2.4 [C8.5.5]) with respect to the conditions for the strength reduction factors of cracked concrete. For
CCT nodes, a simplification has been adopted for the value of in cases with partial anchorage within the node.
This simplification consists of assuming a linear interpolation between the value for a fully smeared nodal region
(parallel compression field) and a concentrated node (full anchorage outside of the nodal region). This choice
(not explicit in current draft of MC2020) avoids requiring a detailed check of the nodal region.
With respect to the stress states of the nodes, they refer normally to pseudo-hydrostatic stress conditions (equal
stresses in two directions, with no out-of-plane stress). This ensures that no explicit check of the stress state in
the nodes has to be performed (refer to Kupfer’s response of biaxially-loaded concrete [C8.5.2]). For some
cases, as for the partial anchorage case shown in Figure 8.28c, a more complex stress state may be provided.
A detailed verification of such state is also not required provided that the values of the factors are respected
for the compression fields converging in the node.
For other cases, similar treatment can be followed, as for instance when multiple struts and ties converge in a
node (which can be subdivided into a number of smaller nodes, see Fig. C8.5.5)
Figure C8.5.5: Example of a node where multiple struts converge and detail of nodal region
8.5.5 Spreading of struts This clause is based on the design approach of previous EN1992-1-1:2004 for §8.5.5(1-2) but making the
design assumptions consistent between different cases. It is also allowed to apply alternative (tailored)
approaches based on specific knowledge.
A new clause 8.5.5(3) has been added. This clause refers to a reinforcement for local spreading of a load
applied near to an edge. This can be a typical situation of columns supported on short walls. The reinforcement
detailed in §8.5.5(3) prevents local cracking to develop in an uncontrolled manner near the loading region.
Respecting this condition does not however exclude to develop a complete load-carrying model for the element.
Alternatively, a detailed analysis of the D-region can be performed ensuring proper detailing.
References [C8.5.1] Schlaich, J.; Schäfer, K.; and Jennewein, M., “Toward a Consistent Design of Structural Concrete,”
PCI Journal, V. 32, No. 3, 1987, pp. 74-150.
[C8.5.2] Muttoni, A.; Schwartz, J.; and Thürlimann, B., Design of Concrete Structures with Stress Fields,
Birkhäuser, Basel-Boston-Berlin, 1997, 143 pp.
[C8.5.3] Vecchio, F. J., and Collins, M. P., “The Modified Compression-Field theory for Reinforced Concrete
Elements Subjected to Shear,” ACI Journal Proceedings, V. 83, No. 2, Mar.-Apr. 1986, pp. 219-231.
[C8.5.4] Fernández Ruiz, M., and Muttoni, A., “On Development of Suitable Stress Fields for Structural
Concrete,” ACI Structural Journal, V. 104, No. 4, July-Aug. 2007, pp. 495-502.
[C8.5.5] fib, WP2.2.4, The International Federation for Structural Concrete, “Design and assessment with
strut-and-tie models and stress fields: from simple calculations to detailed numerical analysis”,
Bulletin 100, fib, 2021.
Background document to
clauses 8.6 and I.8.6
Partially loaded areas
Introduction
This document provides the background to subsection 8.6 (Partially loaded areas).
This problem is solved in most modern design codes (EN-1992-1-1, MC-2010, ACI-318) by strut-and-tie
modelling using the model of Fig. 2, where a1 is the block dimension, and a0 is the dimension of the plate
parallel to a1.
a1
a0
a1 a0
N Sd
a1
a1/2
a1
a1/2
N Sd a1 a0 N Sd a1 a0
2 a1 2 a1
NSd Sd a1
The part of this force that does not need to be deflected N*Sd can be expressed as in Eq. (2).
*
NSd Sd a0
Therefore, the force that does need to be deflected can be determined as in Eq. (11).
NSd NSd
*
Sd a1 a0 NSd
a1 a0
a1
To determine the tensile force needed to achieve the diffusion of the force at the bearing, NSd can be divided
into two halves each located at a distance from the centre of the block of (a1-a0)/4. This force must be
transferred to a1/4 (measured horizontally) from the centre of the block. It is assumed that this is accomplished
in a distance equal to the width of the block, the centroid of the tension force being located at a1/2 (measured
vertically). More generally a certain diffusion angle mod can be assumed. The value of the tangent of this
angle, if the diffusion distance is taken as a1, as in the case represented in Fig. 2, would be tanmod =0,5.
More generally, the tensile force generated by the spreading of the compression stresses can be expressed
as in Eq. (4).
Nsd a a0 Nsd a
TSd tg mod 1 tg mod 1 0
2 a1 2 a1
For a three-dimensional problem, with a bearing plate of dimensions a0b0 and a concrete block of dimensions
a1b1 (b1 is defined later), this equation can be generalized as in Eq. (5) where the geometric mean of the
ratio between plate and block dimensions is used.
NSd a0 b0
TSd tg mod 1
2 a1b1
In existing structures, the presence of sufficient transverse reinforcement cannot always be guaranteed, so
that this tensile force may have to be taken by the tensile resistance of concrete. For this, it may be assumed1
1
Other more complex assumptions are possible, such as modelling the failure surface with a pyramidal shape as proposed by De Vries [C08-5]
in his “physical model”. However, due to the uncertainties regarding stress distribution in concrete within the failure surface, such an assumption
only leads to more complexity which is not rewarded by a better fit.
that the area of concrete resisting this force has a depth of one block dimension (a1) and an effective width,
b1, which will be dependent on the width of the bearing plate and, again, on the height of the resisting section,
a1, and can be expressed as in equation (6).
b1 bh a1 b
Where b is the maximum block dimension and b0 is the dimension of the plate parallel to b, and is another
model parameter (like mod). Finally, it must be considered that the distribution of tensile stresses within this
area will not be uniform as shown on the right-hand side of Fig. 1. Therefore, the mean admissible tensile
stress can be expressed as fctd, where is a stress distribution factor which must be smaller than 1.0 and
fctd is the design tensile strength of concrete. With these definitions, the resistance condition can be expressed
in terms of the ultimate load which produces side blowout failure, NRd,2, as in Eq. (7).
Nsd ab
Tsd TRd tg mod 1 0 0 fctd a1b1
2 a1b1
fctd a1b1
NRd ,2 Nsd NRd ,2 2
tg mod ab
1 0 0
a1b1
Eq. (7) provides an expression for the resistance of the concrete block to side blowout failure. The model
parameters values which are adopted are the following:
- tan mod will be taken as 0,5 as is normally done in code standards.
- will be taken a 2/3 assuming a parabolic distribution of tensile stresses.
- Finally, will be taken as 1.00, assuming a lateral diffusion angle within the resisting section also
equal to mod at failure. This value is consistent with the assumed diffusion angle in the
perpendicular direction.
With these values Eq. (7) simplifies to Eq. (8);
8 a1b1
NRd ,2 fctd
3 a0 b0
1
a1b1
NRd ,2 8 fctd a1b1
sd
'
a0 b0 3 a0 b0 ab
1 0 0
a1b1
Because the tensile strength is subjected in FprEN 1992-1-1:2023 to a sustained load coefficient (kct) which
is not to be applied here, as it would make the model overconservative, the tensile strength has, instead,
been expressed in terms of the square root of the compressive strength throughout the code (as, for instance
for shear resistance). Thus the expressions in FprEN 1992-1-1:2023 are expressed as in Eq. (9)
fck a1b1
NRd ,2
c ab
1 0 0
a1b1
NRd ,2 f ab 1
sd
'
ck 1 1
a0 b0 c a0 b0 ab
1 0 0
a1b1
Besides the tension blow-out failure, two types of compression failures need to be checked, as contemplated
in all code standards regarding concentrated loads. If the loading plate is very small, failure can occur in
triaxial compression at the plate surface. If the plate is large with respect to the concrete block and the force
too high, a uniaxial compression failure can occur. These failure modes are expressed in Eq. (10).
ab
NRd ,1 min 1 1 ;7,0 a0 b0fcd
a0 b0
N ab a b
sd
'
Rd ,1 min 1 1 ;7,0 0 0 fcd
As a b
0 0 As
The limit of 3,0fcd for triaxial compression currently in EN 1992-1-1:2004 was increased to 7,0 fcd in the initial
proposal, as shown in the previous equation. The reason for this is that the current value is extremely
conservative as shown by tests (see tests by Spieth [C08-6], or the Niyogi localized load on concrete cubes
[C08-7] data). Both show an increment in concrete strength up to 7 times the uniaxial concrete strength fcd
when specimens were subjected to high confinement. This can also be seen by applying concrete models for
triaxial stresses such as the model by Menetrey-Williams [C08-8]. The origin of this conservative limit can be
traced back to Leonhardt [C08-1], who, while working with the method of admissible stresses, proposed a
reduction factor of 3,0 for the admissible tri-axial concrete strength. This factor has not been accounted for
when translating the safety format from admissible stresses to the partial factor method. It should be now
time to account for it, the value of 7,0 probably still being conservative. Unfortunately, this value was again
reduced to the traditional value of 3,0 to achieve consensus withing SC2, with the allowance given in section
8.6 (“unless larger resistance can be justified based on refined analysis including tensile stresses due to load
or restraint, where relevant”). These considerations should be accounted for in the National Annexes.
The three failure criteria (tension, uniaxial compression and triaxial compression) are summarized in Fig. 3.
For very small plates, the dominant failure mode is triaxial compression, while for very large plates, with
respect to the concrete dimensions, the failure mode is uniaxial compression. For intermediate plate
dimensions, tension failure dominates.
a0/a1
Fig. 3 – Ultimate load represented against the relative size of the plate and dominating zones for each failure
criterion (adapted from ref [C08-4])
For uniaxial compression failure the actual block dimensions are taken since any reduction thereof
significantly reduces correlation to test data. Experimental observation by Niyogi [C08-7] confirms that
defining the effective unloaded area as a projection of the load plate tends to under-predict the measured
capacities of eccentric specimens.
Even though design codes recommend using a concrete resisting area having a similarity relationship to the
plate dimensions, the application of such criteria to the experimental data available would predict compression
failure in almost all cases whereas tension side blowout was the type of failure observed in all the tests carried
out by De Vries.
Another important point is that the model assumes uniform distribution of stresses on the plate and therefore
it should be ensured that the plate is rigid. This condition can be met by providing a plate depth greater than
half the distance between the bar edge and the plate edge.
The comparison of the model to experimental data is extensively reported in reference [C08-4] and will not
be repeated here.
Some amendments to account for cases where the 3-dimensional behaviour does not occur (cases where
the width of the plate is similar to the width of the concrete block or cases where the dimensions of the plate
are very different (i.e. long plates)), which were pointed out by the Swiss Mirror Group ([C08-9]) were
introduced during discussions.
To account for plate lengths that are similar in width to the dimension of the concrete block, the definition of
the effective dimensions of the concrete block (a1, and b1) are defined as shown in Eq. (11). Dimension a1 of
the contributing concrete area is taken as a which is the length of the load introduction block parallel to a0.
The perpendicular dimension, b1, is taken as the plate dimension in that direction, b0, plus the difference
between a1 and the plate dimension parallel to a1, a0, but not larger than the actual block dimension, b. In this
way if a0=a1 the behaviour is conservatively considered as unidirectional (see Fig. 4b) and when a0 is
significantly smaller than a1 3-dimensional effects are accounted for.
Ac1 1
A if max(a0 , b 0 ) 3d 0
c 0 1 Ac 0
f Ac1
Rd ,t ck
c d1 1
d if max(a0 , b 0 )>3d 0
d
0 1 0
d1
d 0 min(a0 , b 0 )
d1 block dimension parallel to d 0
With these amendments, tests carried out at the ETH for which the formulation of draft D5 provided unsafe
results, turn into conservative estimations of the ultimate capacity as shown in Fig. 5.
200
150
U,model [MPa]
100
50
0
0 50 100 150 200 250
u,exp [MPa]
Formulation in D5 Corrected Equations 1:1 slope
Fig. 5 Comparison of tests carried out at ETH with the formulation for partially loaded area of the D5 draft and
with the corrected model.
eb
b1
ea
a1
a0, F1 a0, F2
a0, F s
c2, F1 2 1
2 2 a s a0, F1 a0, F2 s a0, F1 a0 , F2
0 , F1
2 2 2 2 2
a0, F1 a0, F2
a0, F s a0 , F1 a0, F2 a0, F2 a0, F1
2 2 2 a
c2, F2 0, F 2 s s
2
2 2 2 2 2
Fig. 7 Definition of contributing concrete area for cases where two adjacent plates have different dimensions
References
[C08-1] F. Leonhartd. Vorlesunger über Massivbau. 1973. Springer Verlag, Berlin/Heidelberg.
[C08-2] Alejandro Pérez Caldentey, Francesco Marchetto, Hugo Corres Peiretti, Jesús Iglesias Villareal,
Plate-anchored reinforcement bars: A new simple and physical model for practical applications, In
Engineering Structures, Volume 52, 2013, Pages 168-178, ISSN 0141-0296,
https://doi.org/10.1016/j.engstruct.2013.02.019.
[C08-3] Marchetto, F., Pérez Caldentey, A. and Corres Peiretti, H. (2016), Structural performance of corner
joints subjected to a closing moment using mechanical anchorages: an experimental study. Structural
Concrete, 17: 987–1002. doi:10.1002/suco.201500222.
[C08-4] Marchetto, F. Use of headed reinforcement bars in construction. A theoretical approach to determine
the dimensions of anchorage plates and experimental tests on knee joints subjected to closing moment. PhD
Thesis. Technical University of Madrid, 2015.
[C08-5]DeVries, R.A., Anchorage of Headed Reinforcement in Concrete. 1996, The University of Texas at
Austin: Austin. p. 314.
[C08-6] Hans-Peter Spieth. Das Verhalten von Beton unter hoher örtlicher Pressung. Beton und
Stahlbetonbau, 56(11):257–263, November 1961.
[C08-7] S.K. Niyogi, Bearing Strength of Concrete - Geometric variations. Journal of structural division. July
1973.
[C08-8] P. Menetrey, K. J. William. Triaxial failure criterion for concrete and its generalization. ACI Structural
Journal, 1994, 92 (3), pp 311-318. 1995.
[C08-9] W. Kaufmann, T. M., Markić, T. and Morger, F. (2020). Comment on Section 8.6 Partially loaded
areas, addressing inconsistencies that would partly result in unsafe designs. CEN/TC 250/SC 2/WG 1 N 845,
16.
[C08-10] Meyerhof, G. G. (1963). Some Recent Research on the Bearing Capacity of Foundations. Canadian
Geotechnical Journal, Vol. 1, No. 1, 1963, pp. 16-26.
Introduction
This document provides the background to clause 9.2.3 (Refined control of cracking).
k2
1 2
(1)
21
where 1 and 2 are the greater and lesser tensile strains (2=0 if part of the section is compressed) in the
section. With this definition k2=1.0 for pure tension and 0.5 for pure flexure.
The rationale behind the above factor is that, in bending, the transfer length will be shorter because the
tension force (coloured area in Fig. 2 times the width of the section) that is needed to produce a new crack
will be half than that in tension because the strain of the least tensioned fibre is zero, therefore the transfer
length would be half. The current Eurocode 2 (EN 1992-1-1:2004) formulation, however, instead of providing
small crack spacings in flexure, is calibrated in such a way that the model results in notoriously exaggerated
crack spacings for tension members, as shown already in Fig. 1.
Note that the concept of an effective area is meant to account for the fact that the transfer length is a zone of
load introduction and therefore Navier’s hypothesis is not applicable. The consideration of an effective area
is meant as a way to account for this fact while still assuming the validity of Navier’s hypothesis. But there is
still a difference in the way the stresses are distributed in tension and bending, and this should be accounted
for.
EN 1992‐1‐1:2004
400
Flexure
y = 1.06x
350 R² = 0.94
Tension
y = 1.41x
300 R² = 0.95
200
150
100
50
0
0 50 100 150 200 250 300 350 400
sm,exp [mm]
Flexure Tension 45º line Linear (Flexure) Linear (Tension)
Fig. 1 Comparison of crack spacing predicted by EN 1992-1-1:2004 and the actual measured values
Fig. 2 Effect of distribution of stresses prior to cracking on transfer length according to EN 1992-1-1:2004
The above reasoning, as shown in Fig. 2, is not sound, because what matters is not the stress gradient within
the full tensile zone itself but rather the tensile gradient within the effective area around the bar. In a bending
member of significant size, the effective area can represent only a small part of the cross-section and the
approximation of a tie, as done in MC 2010 can be reasonable. However, in small members, the model of EN
1992-1-1:2004 would be better. Fig. 3 demonstrates that, for tension, the mean stress, mean, in the effective
area will be fctm when the next crack occurs. However, in the case of flexure, the mean stress can be
determined as shown in Eq. (2),
1 1 f ct ,ef
mean
2
f ct ,ef c,min,ef f ct ,ef
2 h xg
h xg hc,ef
1 h xg hc ,ef (2)
f ct ,ef 1
2 h xg
k fl
In the above equation xg is the depth of the neutral axis in the uncracked section. This expression results in
a value to 1.0 for pure tension (xg=) and a value of 0.5 in flexure if hc,ef is equal to (h-xg).
For a rectangular cross section and pure bending, the expression for kfl of Eq. (2) simplifies to Eq. (3):
h
1 h xg hc ,ef
k fl 1
1 1 2 h c , ef 1
2 2hc ,ef
2 h xg 2 h 2 h
(3)
2
h hc ,ef
h
Fig. 3 Consideration of the effective tensile area around a bar in flexural elements
Fig. 4 Cracking patterns of beam cast in ‘good’ casting position (top) and ‘poor’ casting position (bottom)
The effect of casting position has been confirmed by analysis of the cracking pattern of ties, where the face
cast in good casting position has a definite tendency to develop a cracking pattern with more closely spaced
cracks. Fig. 5 provides further illustration of this fact using results of tie specimen 16-20-T tested in 2017.
Fig. 5 Cracking pattern of a tie showing different crack spacing on different faces
Experimental evidence, reported in [C09-8], seems to show by comparison of the cracking patterns of two
side faces of the tie, one concreted in ‘poor’ casting position and one concreted in ‘good’ casting position,
that both cover and bar diameter influence the crack pattern. However as the experimental evidence is still
very limited, a very simple model to account for this effect was adopted to the revision of EC2, meant
essentially at recognizing casting position as an important factor in determining crack spacing, to be refined
in the future when sufficient experimental evidence is available.
the introduction of a coefficient to account for casting position is proposed, as follows:
Note that a more precise definition would lead to a different height of the effective area where two rows of
reinforcement are present than where there is only one layer. This difference is waived in FprEN1992-1-
1:2023 to improve ease-of-use with a small sacrifice in precision.
Fig. 7 Effective tension area of concrete around a group of bars, Ac,ef,group (to be applied when effective tension
areas of adjoining bars overlap)
A limit on the effective area around a bar is given as a linear function of the bar diameter. This condition
accounts for the fact that a bar can only control cracks within its proximity. With the definition of the effective
area given in MC 2010 and EN 1992-1-1, a single bar placed in the middle of a large rectangle of concrete
would have an effective area equal to the area of concrete, and the value of the effective area would increase
indefinitely with the dimensions of the cross section. This does not make sense and a limit is therefore
necessary for consistency.
hx
k1 (7)
r dx
where:
h is the section height
d is the effective depth, and
x is the depth of the neutral axis of the cracked section
This coefficient was included in the left-hand side comments to MC 2010, but not in the right-hand formulation
to avoid explicitly increasing crack width predictions. As an illustration, Fig. 8 shows two tests of the
experimental campaign described in reference [C09-8] with 32 mm cover (represented by circles) and 82 mm
cover (represented by squares). The black symbols show the crack width measured at the level of the bar,
while the white symbols show the crack measured at the most tensioned concrete fibre. The grey symbols
result from multiplying the values of crack width at the bar level by k1/r. As can be seen, the grey points land
almost on top of the white symbols (in some cases the white symbols are not visible anymore), thereby
showing that factor k1/r adequately addresses the issue of increase of crack opening due to curvature.
0.60
0.40
0.30
0.20
0.10
0.00
0 100 200 300 400 500 600
Theoretical steel stress ‐ State II [MPa]
16‐20‐F At bar level c=32 mm 16‐20‐F At most tensioned fibre of concrete c=32 mm
16‐20‐F Bar level * (h‐x)/(d‐x) 16‐70‐F At bar level c=82 mm
16‐70‐F AT most tensiones fibre of concrete c=82 mm 16‐70‐F Bar level * (h‐x)/(d‐x)
Fig. 8 Illustration of the performance of factor k1/r on two beam tests with different covers
0.70
0.60
wm, est=TS sm,exp s/Es [mm]
0.50
Flexure
sm,est = 0.58wm,exp
0.40 R² = 0.65
Tension
0.30 sm,est = 0.76wm,exp
R² = 0.73
0.20
0.10
0.00
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
wm,exp [mm]
Fig. 9 Comparison between measured crack width and estimated crack width without accounting for factor
k1/r.
By comparing the measured crack width against the estimated crack width (calculated by multiplying the
experimentally observed crack spacing by the strain in the reinforcement at the crack and by a tension
stiffening factor (TS), according to EN 1992-1-1:2004), see Fig. 9, two effects can be observed:
– The effect of shrinkage happening before the test. Shrinkage compresses the reinforcement before
the tests. This strain will then be released when cracking occurs and, as a result, the crack width
due to the load is overestimated if no correction is introduced to account for the pre-existing
compressive strain in steel. This explains why the estimated cracks widths fall below the measured
values
– The effect of curvature, since the underestimation of the crack width is much larger for bending
tests than for tension tests.
0.70
0.60
0.20
0.10
0.00
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
wm,exp [mm]
Fig. 10 Comparison between measured crack width and estimated crack width accounting for factor k1/r.
If coefficient k1/r is used to correct the estimated crack width, the skew between tension tests and flexure tests
shown in Fig. 9 is corrected, as is shown in Fig. 10. Even so, the slope of the trendline is still not satisfactory.
As mentioned above, the fact that the slope is still smaller than 1, is attributed to the shrinkage of concrete
occurring before the tests. These effects can, unfortunately, not be adequately corrected since the shrinkage
strains developing before the tests are not generally reported in the literature.
1 f
sm cm s kt ctm 1 e p , eff (10)
Es p , eff
where:
350
Flexure
300 sm,FprEN 1992 = 0.98sm,exp
sm, Fpr EN 1992‐1‐1:2023 [mm]
R² = 0.95
250 Tension
sm,FprEN1992 = 0.99sm,exp
R² = 0.97
200
150
100
50
0
0 50 100 150 200 250 300 350 400
sm, exp [mm]
Flexure Tension 45º line Linear (Flexure) Linear (Tension)
Fig. 11 shows the comparison between predicted values, srm,cal FprEN1992-1-1:2023, and experimental values of
crack spacing, sm,exp. The correlation lines show a high coefficient of determination of 0.95 in flexure and 0.97
in tension and a slope close to 1.00. By comparison of Fig. 1 and Fig. 11, it can be observed that scatter is
reduced as it will now be demonstrated.
Table 2 shows the statistical parameters referred to the 73 tests used for the calibration for crack spacing
(see Annex A of reference [C09-10]). The table includes the mean squared error, the minimum value of the
ratio of model prediction and experimental values (min), its maximum value (max), its mean value (), its
standard deviation () and its coefficient of variation (COV).
It can be seen that the proposed corrections improve the prediction quality in statistical terms, for all the
considered statistical parameters, both in flexure and in tension.
nº of tests 37 36 37 36
200
srm, FprEN 1992:2023 [mm]
150
100
50
0
0 50 100 150 200 250
sm, exp [mm]
Test Data Set ‐ Flexure
Fig. 12 Comparison between predicted and measured crack spacing for the independent test series (Annex
B)
Table 3 Statistical analysis of crack spacing for the independent data set
/N)
2 26.30 23.55
Model/Measured
max(sm,model/sm,exp)= 1.63 1.76
Summary
In this background document, reasons have been given for why the cracking model of EN 1992-1-1:2004 had
to be revised, the most important being the uneven behaviour regarding crack width in tension and flexure.
The main proposals to improve the cracking model are the following:
– A new coefficient has been proposed to account for the effects of non-uniform stress distribution
(kfl). The introduction of kfl by replacing k2 with a better value, does away with the gross
overestimation of crack spacing in tension and allows to delete the (h-x)/3 limit to the effective
height currently implemented in the EN 1992-1-1:2004 model.
– Recent experimental evidence demonstrates that casting position has a strong influence on crack
spacing, increasing it when tension bars are concreted in ‘poor’ casting position. This effect has
been accounted for by introducing factor kb.
– The definition of the effective area has been reformulated to eliminate some of the inconsistencies
of the current model.
– The calibration, made using the database used in the original calibration of MC 2010, has been
tested against an independent database and has proven to be robust.
– It has been observed that the model for the crack width, which derives the mean crack width from
the product of the mean crack spacing and the mean strain difference of steel and concrete
underestimates the measured crack width because of the generally unreported effect of shrinkage
of concrete occurring before testing. It is therefore not possible to make a meaningful comparison
between models and experimental databases in terms of crack width.
[C09-5] Corres Peiretti, H., Pérez Caldentey, A., Petschke, T. PrEN Chapter 7 – Serviceability Limit State.
References
Cracking. Supporting Document. May 2003.
[C09-6] fib Model Code for Concrete Structures 2010 (2013). Fédération internationale du béton (fib),
Lausanne, Switzerland. ISBN 978-3-433-03061-5
[C09-7] CEN-TC250 (2004), EN 1992-1-1. Eurocode 2. Design of concrete structures. General rules for
buildings.
[C09-8] García. R. Pérez Caldentey, A. (2018). Cracking of Reinforced concrete: Tension vs. Flexure: Report
on cracking tests of large ties. Technical University of Madrid (UPM).
[C09-9] fib Bulletin 92 (2019), Serviceability Limit State of Concrete Structures. Lausanne, Switzerland, ISSN:
1562-3610, pp 292. doi.org/10.35789/fib.BULL.0092.
[C09-10] Pérez Caldentey, A, García, R, Gribniak, V, Rimkus, A. Tension versus flexure: Reasons to modify
the formulation of MC 2010 for cracking. Structural Concrete. 2020; 21: 2101– 2123.
https://doi.org/10.1002/suco.202000279
[C09-11] Broms, Bengt B., (1964) Crack Width and Crack Spacing in Reinforced Concrete Members, JL62-
67 Part II.
[C09-12] Clark, A.P. (1956), Cracking in Reinforced Concrete Flexural Members. ACI Journal, Proc. vol. 27,
No. 8, Apr. 1956, pp.851–862.
[C09-13] CUR (1994) Dutch Centre for Civil Engineering, Research and Codes Report no. 37 (Out
of print).
[C09-14] Farra, B., Jaccoud, J.P., (1993), Influence du Béton et de l’armature sur la fissuration des structures
en Beton. Rapport des Essais de Tirants sous déformation imposée de courte durée. EPFL, Lausanne.
Département de Génie Civil, École Polytechnique Fédérale de Lausanne, pub. No. 140.
[C09-15] Farra, B., Jaccoud, J.P. (1996) Rapport des Essais de Tirants sous déformation imposée de courte
durée.
[C09-16] Hognestad (1962). Journal of PCI Research and Development Laboratories. 1962.
[C09-17] Rüsch, E.H., Rhem G. (1963-64), Versuche mit BetonformStählen. (1963), pt. II (1963), pt. III (1964).
Deutscher Ausschuss für Stahlbeton, No. 140.
[C09-18] Pérez Caldentey, A., Corres Peiretti, H., Peset Irribaren, J. and Giraldo Soto, A. (2013), Cracking of
RC members revisited: influence of cover, φ/ρs,ef and stirrup spacing – an experimental and theoretical study.
Structural Concrete 01/2013. Nr. 16, pp. 69-78.
[C09-19] Rimkus A. Effects of Bar Reinforcement Arrangement on Deformations and Cracking of Concrete
Elements. Doctoral Dissertation. Vilnius: Technika, 2017. 143 p
[C09-20] Jakubovskis, R.; Kaklauskas, G.; Gribniak, V.; Weber, A.; Juknys, M. 2014. Serviceability analysis
of concrete beams with different arrangement of GFRP bars in the tensile zone. ASCE Journal of Composites
for Construction, 18(5): 04014005-1–10. DOI: 10.1061/(ASCE)CC.1943-5614.0000465.
[C09-21] Krips, M. (1984), Rissbreitenbeschränkung im Stahlbeton und SpannBeton. 1984. Doctoral thesis
[C09-22] Viktor Gribniak, Alejandro Pérez Caldentey, Gintaris Kaklauskas, Arvydas Rimkus, Aleksandr
Sokolov (2016) Effect of arrangement of tensile reinforcement on flexural stiffness and cracking, Engineering
Structures, Volume 124, 2016, Pages 418-428, ISSN 0141-0296.
[C09-23] Rimkus, A.; Gribniak, V. 2017. Experimental Investigation of cracking and deformations of concrete
ties reinforced with multiple bars. Construction and Building Materials, 148: 49–61. DOI:
10.1016/j.conbuildmat.2017.05.029.
[C09-24] Steyl L, (2016) Plastic cracking of concrete and the effect of depth. PhD thesis. Stellenbosch
University.
[C09-25] Combrinck, R., Steyl, L., Boshoff, P. (2018), Interaction between settlement and shrinkage cracking
in plastic concrete. Construction and Building Materials 185:1-1
[C09-26] Borosnyói, A., Balázs, G. (2005), Models for flexural cracking in concrete: the state of the art.
Structural Concrete Vol. 6, Issue 2. pp. 1464-4177. Thomas Telford, fib.
[C09-27] Annette Beldholm Rasmussen. (2019). Modelling of reinforced concrete in the serviceability limit
state : A study of cracking, stiffness, and deflection in flexural members. PhD Thesis. Aarhus University.
[C09-28] Pérez Caldentey, A., Jones. T., Goodchild, C. Consideration of shrinkage strain in cracking
calculations. Background Document to Clause 9.2.4 of prEN 1992-1-1:2020. CEN TC 250/SC2/WG
[C09-29] Gribniak, V.; Jakubovskis, R.; Rimkus, A.; Ng, P.-L.; Hui, D. 2018. Experimental and numerical
analysis of strain gradient in tensile concrete prisms reinforced with multiple bars. Construction and Building
Materials, 187: 572–583. DOI: 10.1016/j.conbuildmat.2018.07.152.
[C09-30] R. I. Gilbert, S. Nejadi. An experimental study of flexural cracking in reinforced concrete members
under short term loads. Report. University of South Wales. 2015.
[C09-31] Calderón Bello, E. (2008). Estudio experimental de la fisuración en piezas de hormigón armado
sometidas a flexión pura. PhD Thesis. Technical University of Madrid.
[C09-32] Wu, M.H.Q (2010). Tension stiffening in reinforced concrete – instantaneous and time-dependent
behaviour. PhD Thesis. The University of Southern Wales.
[C09-33] Frosch, R. J., Blackmann D.T, Radabaugh, R.D. (2003). Investigation of bridge deck cracking in
various bridge superstructure systems. Report No. FHWA/IN/JTRP-2002/25. Purdue University West
Lafayette, Indiana 47907. February 2003.
[C09-34] Base, C.D., Beeby, A.W., Taylor, P.J. (1966). An investigation of the crack control characteristics of
various types of bar in reinforced concrete beams. Research report 18. Cement and Concrete Association.
[C09-35] Kaklauskas, G. (2019) Personal communications (unpublished matter).
03.01.2023
For citations, copying and referencing see Note on Front page! page 384
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
9.2.3 (4) Background to FprEN 1992-1-1:2023
Introduction
This document provides background to clause 9.2.3 Refined control of cracking and addresses under which
circumstances shrinkage strain increases crack width. The document centres on a tie and evaluates the
problem considering increasing degrees of complexity, as follows:
– Tie subjected to shrinkage but free to move
o linear elastic problem: the tie does not crack due to redistribution of stresses
o cracked analysis
– Tie subjected to shrinkage but with movement at the ends fully restrained. An external axial force
develops due to shrinkage
o linear elastic analysis
o cracked analysis
As a result of these analyses, it will be concluded that shrinkage would tend to increase the mean strain of
steel with respect to concrete by the shrinkage strain reduced by a tension stiffening factor. However, further
analysis will reveal that, for the crack formation stage, the increase in crack width will be very small due to
the reduction in the cracking force produced by the tensile stresses generated by shrinkage in the uncracked
concrete. Furthermore, it will be concluded that, for stabilised cracking, the development of shrinkage will
result in a reduction of the transfer length, and therefore, in the formation of a new crack between the most
widely spaced ones. For these reasons, FprEN 1992-1-1:2023 does not require the consideration of
shrinkage strains in crack width calculations.
L
s Es s
c c Ec
c cs c (1)
Ec Ec 1 c cs
Compatibility conditions dictate that the strain variation in reinforcement, s, should be equal to the strain
variation in concrete, c, and equal to the strain variation of the reinforced concrete section, (See Eq.(2))
s c (2)
Finally, equilibrium implies that the variation in axial force, N, due to the development of shrinkage should
be zero, since the element is free to move and, therefore, the reaction will be zero. The tension force that will
develop in concrete due to the fact that concrete will shorten less than the shrinkage strain, must be balanced
by an equal compressive force in steel (see Eq. (3))
N 0 Ac c As s (3)
By introducing the conditions of compatibility into the constitutive laws and entering these into the equilibrium
equation, the strain of the element can be determined. Then, by introducing this strain into the constritutive
laws, the stresses in the materials can be calculated (see Eq.(4)). This equation will result in tension in
concrete since ||<|cs |.
s Es
Ec
c cs
1
N Ac c As s 0
Ec Es 1
(4)
Ac As cs Ac 0
1 Ec
Ah ,
A cs cs
cs c
Ah , E 1 1 e ,
1 s
Ec
Therefore, if the shrinkage strain is high enough so that the stress in concrete exceeds fct,eff, the section
would crack. The next paragraph develops the analysis for this case.
Fig. 2 – Simply supported tie free to move longitudinally with crack in concrete
Therefore, at the crack, the expressions of Eq. (5) will govern behaviour. Immediately adjacent to the crack
there will be a difference between the strain in the steel and the strain in concrete equal to –cs (tension)
s 0
s c cs
c cs
(5)
s 0
c 0
One transfer length away from the crack, compatibility will be reestablished. Tension stiffening effects can be
accounted for as follows:
cs
sm s kt s sE 0 kt 0
1 e, s
cs (6)
cm cs kt cs
1 e, s
sm cm cs 1 kt
where:
sE is the strain in the reinforcement at the section where compatibility between concrete and steel is re-
established.
If kt is taken as 0,4, as in FprEN 1992-1-1:2023 and EN 1992-1-1:2004, only 60% of the shrinkage strain will
contribute to increase the relative mean strain of steel with respect to concrete.
cs N
s c 0
E 1 Ec Es 1
1 s Ac As
Ec
1 Ec
Ac Ah ,
cs
Ah ,
Ec Ac (7)
N cs
1
s E s s 0
Ec Ec
c 0 cs
1 1 cs
The next section deals with analysis after concrete cracks.
N cr
s N
Es As s c cr cs
Es As
c cs
(8)
N
s cr
Es As
c 0
The effect of tension stiffening for this case can be evaluated as done before, by adapting Eq. (6) as follows
(Eq. (9)). This equation is in fact general and also valid for cases when restraint is not rigid. For a fully
restrained member, the condition that the mean steel strain be nil, need not apply between every two cracks
but only at the structural level along the full length of the element. This condition would allow to determine the
value of the shrinkage strain that would produce a new crack. This value would increase with the reduction
of the stiffness due to the appearance of new cracks.
N N cs N cr
sm s kt s sE cr kt cr
Es As Es As 1 e , s Ec Ah ,
1
(9)
cs N cr
cm cs kt cs
1 e , s Ec Ah ,
1
N cr
sm cm cs 1 kt
Es As
Again, tension stiffening applies to shrinkage in the same way as it applies to loads. It is stressed that for
axially constrained members, shrinkage will produce both an axial force equal to the cracking force, thereby
contributing to the value of tensile strain in reinforcement at the crack (s) and a compressive strain in
concrete, thereby increasing by these two effects the relative steel strain with respect to concrete at the crack.
The analysis, however, is not yet concluded, since the cracking force in the tie will in fact not be constant as
shown in Fig. 5, but will show a tendency to decrease. This is because shrinkage induces tensile stresses in
concrete between cracks and this makes it easier to produce new cracks in the crack formation stage. The
external force will, therefore become smaller with time, as shown in Fig. 6.
Ncr
Fig. 6 Behaviour of a tie subjected to an imposed strain accounting for redistribution of forces due to shrinkage
In order to try to quantify this effect four fully embedded ties representing elements with high and low
reinforcement ratios and large and small heights have been analysed using step-by-step time-dependent
analysis, using the software described in (Pérez Caldentey, 1996). Table 1 shows the main characteristics
of the sections that were considered. All the sections were assumed to be cast with a C30/37 concrete. To
determine the shrinkage strain, 50% relative humidity was assumed.
Section 1 2 3 4
width [mm] 1000 1000 1000 1000
height [mm] 250 250 1000 1000
Reinf. 512 1025 10 1032
0.28% 2.46% 0.21% 0.85%
hc,ef [mm] 110 167 110 175
cs (t=100y)[mm/m]= 0.622 0.622 0.575 0.575
Fig. 7 shows the evolution of the force needed to produce a new crack in the tie as function of time. The
cracking force increases in the short term due to the increase in the concrete strength. However, this tendency
is reversed after a while by the tensile stresses generated by the development of shrinkage. The reduction
on the cracking force is much more pronounced for elements with higher reinforcement ratios than for
elements with small reinforcement rartios. This is logical since stresses in concrete develop because
reinforcement does not allow the free development of shrinkage. The larger the reinforcement ratio the more
restrained the shrinkage strain and therefore the larger the tensile stresses in concrete needed to fulfil
compatibility conditions.
1 000
900
800
700
600
Fcr [kN]
500
400
300
28 days
200
100
0
1 10 100 1000 10000
Time [days]
h=25 cm ‐ 10phi12 h=25 cm ‐ 10phi25 h=100 cm ‐ 10phi16 h=100 cm ‐ 10phi32
0.50
0.45
0.40
0.35
0.30
0.15
0.10
0.05
0.00
1 10 100 1000 10000
Time [days]
h=25 cm ‐ 10phi12 h=25 cm ‐ 10phi25 h=100 cm ‐ 10phi16 h=100 cm ‐ 10phi32 No shrinkage
Fig. 8 Crack width as a function of time, accounting for and ignoring shrinkage strain
Nonetheless, as shown in Table 2 and Fig. 9 the error, is rather small, no larger than 18%, with the error
being larger for the lower reinforcement ratios as was to be expected. It was therefore agreed within SC2/WG1
to neglect this effect given the large uncertainties surrounding crack width calculations.
Table 2. Comparison of crack width calculations performed considering and neglecting the effect of
0.450
0.400
0.350
0.300
wk,cal [mm]
0.250
0.200
0.150
0.100
0.050
0.000
25cm‐10phi12 25cm‐10phi25 100cm‐10phi16 100cm‐10phi32
Section
Shrinkage ignored Shrinkage considered
Fig. 9 Comparison of calculated crack width accounting for and neglecting shrinkage
If stabilised cracking is reached, then the effect of shrinkage can be more easily neglected. As it will be
demonstrated, in such a case, shrinkage would end up by producing an intermediate crack, whose effect
would more than compensate for the effect of shrinkage on the largest crack width by reducing the crack
spacing to half.
Fig. 10 Equilibrium of the bar between the section at the crack and the section at the end of the transfer length
Fig. 10 shows the equilibrium of a reinforcing bar belonging to a tie between the section of a crack and the
section at the end of the transfer length (ls) where a new crack would form. The equilibrium condition of the
bar leads to the expression of the transfer length given in Eq. (10), where bm is the mean bond strength. bm is
used for simplicity as a “mean value” of time dependent bond strengths over the length between the crack
and the end of the transfer length.
s SE (t )
ls (t ) (10)
4 bm (t )
Fig. 11 shows the equilibrium between the section at the crack where all the force is resisted by the
reinforcement only and the section at the end of the transfer length where part of the force is resisted by
concrete in tension. The tensile capacity of concrete (fctm×Ac,eff) is reduced by the tension induced between
cracks by shrinkage (c×Ac,eff). From this equilibrium condition it is possible to obtain the variation of stress
in the reinforcement bar (s2 -sE), as shown in Eq. (11), and by introducing this expression in Eq. (10), the
transfer length can be obtained.
Fig. 11 Sectional equilibrium between the section at the crack and the section at the end of the transfer length,
accounting for the effect of shrinkage
Ac ,eff
s SE (t ) f ctm (t ) c
As
Ec 1 Ac ,eff
s SE (t ) f ctm (t ) cs 1
1 1 A
e , ef s
(11)
Ec
SE (t )
1 f (t ) (t ) 1
ls (t ) s ctm bm cs 1
4 bm (t ) 4 bm (t ) 1 1
eff
e , eff
1.8
Eq. (11) shows that the transfer length is reduced by the development of shrinkage, in spite of the increase
in tensile strength, because this increase is compensated by an increase in the bond strength. This remark
leads to the conclusion, that as shrinkage develops, new cracks will form in the specimen, thereby reducing
the maximum crack width.
Conclusions
As a result of the above considerations, the following conclusions can be drawn:
– If the effects of shrinkage on restraining loads and its effect on the transfer length are neglected, it
is found that shrinkage increases the mean relative strain of steel with respect to concrete. This
increase can be mitigated by accounting for tension stiffening effects, applied to the shrinkage
strain.
– In the crack formation stage, the above effect is partially compensated by the reduction with time
of the force needed to crack the section due to the development of tensile stresses in the uncracked
concrete due to shrinkage. This means that subsequent cracks will form for a lower value of the
restraining axial force and will result in smaller stresses in the reinforcement. Although this effect
does not fully compensate for the increase in crack width due to shrinkage, the errors in the crack
width estimate are below 20%. Given the large uncertainties around crack width calculations it is
considered an acceptable simplification to neglect altogether the effect of shrinkage for this case.
– For the stabilised cracking stage, shrinkage leads to reduction of the transfer length. This means
that new cracks will form between the cracks that are most widely spaced resulting in the reduction
of the crack width at that location by a factor of 2. This justifies neglecting the effect of shrinkage in
the stabilised cracking stage.
– The above considerations allow to fully neglect the effect of cracking for crack width calculations.
References
Bibliography
[1] CEB-FIP. (1993). Model Code 1990. Design Code. Lausanne: Thomas Telford.
[2] CEN. (2004). EN 1992-1-1. Eurocode 2 Design of concrete structures - Part 1-1: General rules and rules
for buildings . Brussels: Comité Européen de Normalisation.
[3] Farra, J. J. (1996). Rapport des Essais de Tirants sous déformation imposée de courte durée. Lausanne:
EPFL.
[4] fib. (2013). Model Code for Concrete Structures 2010. Lausanne: Fédération Internationale du Béton.
[5] Pérez Caldentey, A. (1996). Comportamiento en servicio del hormigón estructural (PhD Thesis - Technical
University of Madrid). Madrid: ache.
[6] Scott, R. J. (2020). Influence of shrinkage strains on the cracking of reinforced concrete beams. The
Structural Engineer: journal of the Institution of Structural Engineer, págs. 25-29.
[7] Trost, H. (1991). Creep, Relaxation and Shrinkage of Structural Concrete. . lABSE Colloquium Structural
Concrete. Stuttgart: 1991.
04.01.2023
For citations, copying and referencing see Note on Front page! page 395
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
9.3.3 and 9.3.2 Background to FprEN 1992-1-1:2023
Introduction
This document provides the background to clauses 9.3.3 (Simplified calculation of deflections for reinforced concrete
building structures) and 9.3.2 (Simplified deflection control by span / depth-ratio for buildings).
9.3.3 Introduction
Expression (9.23) Deflection calculations are complex due to the non-linear behaviour of concrete which cracks, has tension stiffening
effects and has a complex time-dependent behaviour. To consider this behaviour, concrete codes, such as EN 1992-
1-1, propose to determine the deflection of RC elements by nonlinear interpolation between cracked and uncracked
states (-method). Creep is taken into consideration by an effective modulus of elasticity while a separate calculation
is needed to consider the effect of shrinkage. In its more rigorous form, this method allows the calculation of
curvatures in different sections while the deflections must be determined by double integration of curvatures. The
application of this method becomes very complex in practical applications due to complex load histories and the fact
that the calculation is nonlinear. Furthermore, the uncertainties involved in construction times contribute to question
whether the significant computational effort required for a rigorous application of the method leads to better deflection
predictions. Yet another shortcoming of this method is that it is very complex to apply to continuous structures, since
the complexity is enhanced by the complexity of the structure’s geometry (possibly variable spans, variable widths,
presence of voids, etc.).
As is well known by building designers, deflections determine the depth of concrete floors, which must be optimized
to maximize the available free space. It is very important to have good tools to determine for each case the required
minimum depth at an early stage of the project and this topic has been addressed in several papers ([C09-6][C09-
7][C09-8][C09-9]), and incorporated, with more or less fortune into major design codes ([C09-10][C09-11][C09-12]).
Calculation of deflections in reinforced concrete beams is, however, not an easy task due to the complex material
behaviour (cracking, tension stiffening, time-dependent effects, etc.) and the dependence of the results on the
construction sequence and time of application of loads. Many complex models have been developed to this end, as
for instance those of Marí et al. [C09-13], Ghali et al. [C09-14] or more recently Ulm et al. [C09-15]. The effect of
shear deformations on deflections has been studied by Debernardi et al. [C09-16], leading to the conclusion that this
effect is only important for very stocky members and therefore negligible for common flexural members and will not
be considered here.
Due to the complexity of these calculations, there has been a search for simplified methods. Most of them are based
on simplifications, such as those incorporated in MC 1990[C09-17] or EHE-08[C09-18]. The original basis of this last
method can be found in reference [C09-19] with corrections introduced later on and documented in reference [C09-
20].
kI LOADS kS cs
(1)
cs is determined by applying on the linear elastic model a curvature given by Expression (2):
1 E S
s cs s (2)
r
cs E c, ef Ig
where:
Ss is the first order moment of the provided tension and compression reinforcements with respect to the
centroid of the gross concrete cross section,
Ig is the inertia of the gross concrete cross section,
Es is the modulus of elasticity of steel and,
1, 05Ecm
Ec,ef is the effective modulus of elasticity of concrete: Ec ,eff ,where mean is the mean creep
1 mean
coefficient. If it is assumed that the dead load (qDL) is applied at age t0, the superimposed dead load (qSDL) is applied
at age t1, and the quasi-permanent live load (2qLL) is applied at age t2, mean can be estimated from Expression (3).
k k k
II 1 I 1
Ie I II II
(4)
II I Ig Ig
kI I 1 1
Ie I II Ie I cr
qL4
Where k is the expression for the deflection divided by the equivalent inertia (for example k ).
Ec ,eff
The ratio between the gross inertia and the cracked inertia can be analytically determined as follows, as a function of
the homogenised reinforcement ratio (e,ef), which accounts for creep effects:
1 3
e ,ef 1
2
3
I cr 3 0.6 d
Ig
3
2.7 e , ef
h (5)
1 h
12 d
where:
=x/d is the ratio of the cracked neutral axis to the effective depth of the section
=As/bd is the reinforcement ratio referred to the tension reinforcement
d/h is the ratio between the effective and the total height of the section
The approximation to the ratio between the fully cracked and gross inertia given in Eq. (5) is plotted in Fig. 1. The
approximation is quite close to the theoretical values.
2.50
2.00
1.50
Icr/Ig
1.00
0.50
0.00
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
e,ef
Icr/Ig, aprox Icr/Ig
Fig. 1 Ratio between the fully cracked and gross inertia and proposed approximation
The -factor can be determined assuming that the member is subjected to bending but no axial force, and that the
calculation is for long-term analysis as:
2
M
1 0.5 cr (6)
Mk
where:
Mcr is the moment that cracks the critical section
Mk is the characteristic moment at the critical section
1.00 if M k M cr
kI Ig
I 1 if M k M cr
(7)
cr
kS Se/Sg is aimed to estimate the ratio between the first order moment in the cracked section, accounting for tension-
stiffening effects, and the first order moment of reinforcement in the gross section. It has been fitted to a mean value
for concrete strengths classes C30, C60 and C100 (see Fig. 2) using an estimation of the equivalent first order
moment of the area of reinforcement with respect to the centroid of the section, Se, determined by interpolating
between cracked and uncracked states, using parameter defined in paragraph 7.4.3 of EN 1992-1-1:2004. The
results are not very sensitive to strength classes and this approach is deemed sufficient. The resulting formulation is
given in Expression (8). As can be seen, a distinction is made between reinforcement ratios, for which the section will
crack under the characteristic combination of actions, and situations for which it will not.
1,00 if M k M cr
kS (8)
455 35 1.6 if M k M cr
2
1.80
1.60
1.40
1.20
1.00
Se/Sg
0.80
0.60
0.40
0.20
0.00
0.00% 0.50% 1.00% 1.50% 2.00% 2.50% 3.00% 3.50% 4.00%
C30 C60 C100 Proposed adjustment kS
Verification
A verification of the formulation described above has been carried out by computing the deflection of simply
supported beams covering the full range of practical variables. For this, the following values were used:
Concrete classes C30, C60 and C100
Geometrical reinforcement ratios of 0,3%, 0,5%, 1%, 1,5%, 2% and 2,5%
Effective heights of 0,2; 0,3; 0,4; 0,5; 0,6; 0,7; 0,8; 0,9 and 1,00. Total heights were assumed to be equal to
effective heights plus 0,05 m
qLL/qTot ratios were taken as 0,5, 0,4 and 0,25
All possible combinations of the above values were considered so that calculations were made for 3693=486
cases. The rheological parameters were determined for each one. The creep coefficients were determined using
Expression (3) assuming that:
The self-weight is applied at 7 days
The superimposed dead load is 15% of the self-weight and is applied at 60 days
The quasi-permanent live load is applied at 365 days
The relative humidity is 50%
The deflection is determined for a design life of 100 years
Table 1 shows a selection of the rheological parameters used for the calibration
fck [MPa] d [m] h/d qLL/qTot mean cs [mm/m] (36500,7) (36500,60) (36500,365)
30 0.2 1.25 0.5 2.84 0.617 3.2 2.29 1.74
60 0.2 1.25 0.5 1.62 0.528 1.84 1.31 0.96
100 0.2 1.25 0.5 1.07 0.47 1.21 0.86 0.62
30 0.2 1.25 0.4 2.91 0.617 3.2 2.29 1.74
60 0.2 1.25 0.4 1.66 0.528 1.84 1.31 0.96
100 0.2 1.25 0.4 1.1 0.47 1.21 0.86 0.62
30 0.2 1.25 0.25 2.99 0.617 3.2 2.29 1.74
100 0.2 1.25 0.25 1.13 0.47 1.21 0.86 0.62
30 0.5 1.1 0.5 2.56 0.565 2.9 2.07 1.54
60 0.5 1.1 0.5 1.5 0.492 1.7 1.21 0.88
100 0.5 1.1 0.5 1 0.448 1.14 0.8 0.58
With the above formulation it was possible to obtain a quite good approximation to the -method currently proposed
by EN 1992-1-1:2004 and which is kept as the general method in FprEN 1992-1-1:2023. Table 2 shows the general
statistics of the ratio of the deflection predicted by the simplified model and that predicted by the -method, including
coefficient of variation (Cv), standard deviation (), the mean value (), as well as the minimum and maximum
values, and 5% and 95% quantiles assuming a normal distribution. These values are given for the three concrete
classes considered together and separately. The simplified formula provided consistent values.
Table 2 Statistics for the simplified method
Mcr determined with Ig
All tests C30 C60 C100
9.3.3 method/‐method
Fig. 3 shows the comparison between both methods. The regression line is close to the 45º line, and the coefficient
of determination (R2) is very high. Some of the results seem to be over-conservative. These correspond to low
reinforcement ratios as can be seen from Fig. 4. This is not surprising as for these reinforcement ratios the section is
on the limit of cracking or not, and, of course, this makes a very significant difference in the value of the estimated
deflection when the assumption is not correct.
Fig. 4 through Fig. 8 show the variation of the ratio of the deflection predicted by the simplified method and by the -
method with respect to the reinforcement ratio, the concrete strength, the beam height, the ratio of total height to
effective height and the live load to total load ratio. No significant bias is observed with respect to any of these
variables.
200
180
160
Deflection Simpl. Formula [mm]
140
120
y = 1.0587x
R² = 0.9937
100
80
60
40
20
0
0 20 40 60 80 100 120 140 160 180 200
Deflection ‐method [mm]
1.8
1.6
1.4
Simpl. Formula/‐method
1.2
0.8
0.2
0
0 0.005 0.01 0.015 0.02 0.025 0.03
Geometrical Reinforcement ratio,
1.8
1.6
1.4
Simpl. Formula/‐method
1.2
0.8
0.6
y = ‐0.0012x + 1.1304
0.4 R² = 0.2396
0.2
0
5 15 25 35 45 55 65 75 85 95 105
fck [MPa]
1.8
1.6
1.4
Simpl. Formula/‐method
1.2
1
y = 0.0067x + 1.0504
R² = 0.0006
0.8
0.6
0.4
0.2
0
0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00 1.10
Effective height [m]
1.8
1.6
1.4
Simpl. Formula/‐method
1.2
0.8
y = ‐0.0615x + 1.1225
R² = 0.0029
0.6
0.4
0.2
0
1.00 1.05 1.10 1.15 1.20 1.25 1.30
h/d
Fig. 7 – Results with respect to total height/effective height ratio – no significant bias
1.8
1.6
1.4
Simpl. Formula/‐method
1.2
y = ‐0.0053x + 1.0565
0.8
R² = 6E‐05
0.6
0.4
0.2
0
0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55
qLL/qTot
9.3.2 Converting the formula for deflection calculations into a formula for the slenderness limit
Expression (1) can easily be converted into a general formulation to determine the slenderness limit. In this
Simplified
paragraph it will be shown how to do this. Expression (9) shows the condition for the slenderness limit, i.e., that the
deflection control
deflection be limited to a fraction (1/a) of the span. It also shows the expressions of the deflections due to a uniformly
by span/depth ratio
distributed load and a constant curvature due to shrinkage.
L
kI LOADS kS cs
a
5 qqp L4
LOADS K
384 E 1 3
bh
c ,eff
12 (9)
h h
A d As h d 2
Es s 2 2 L
Kcs cs
cs
Ec ,eff 1 3 8
bh
12
where:
K is a factor that considers the support conditions for the deflection due to uniformly distributed loads and can
be determined from Expression (10) (for a detailed derivation see [C09-9]):
f simply,sup
K3 (10)
f real , sup,cond
where:
fsimply,sup is the linear elastic deflection of the simply supported member of arbitrary span subjected to a
uniformly distributed load, and
freal,sup,cond is the linear elastic deflection of the member with the actual support conditions with the same
arbitrary span and subjected to the same uniformly distributed load.
Kcs is a factor that considers the support conditions for the deflection due to shrinkage and can be determined
from Expression (11) (for a detailed derivation see [C09-9]):
fcs , simply,sup
Kcs (11)
fcs ,real ,sup,cond
where:
fcs,simply,sup is the linear elastic deflection of the simply supported member of arbitrary span subjected to a
constant curvature, and
fscs,real,sup,cond is the linear elastic deflection of the member with the actual support conditions with the same
arbitrary span and subjected to a constant curvature.
Assuming that the provided reinforcement is that strictly needed in ULS, the value of the quasi-permanent load can
be determined from the ultimate bending resistance as shown in the last equation of Expression (12).
L2 Af
M Rd qRd As f yd d 0,5 s yd
8 bfcd
As f yd 8
qRd As f yd d 1 0,5
bdfcd L2
LL
1
TL (12)
LL 2 LL LL
qqp G 2Q 1 2
TL TL TL
kDL
qRd GG QQ LL LL LL
1 G 1
Q
G
TL TL TL
Q
LL
TL
8
qqp kDL As f yd d 1 0,5 2
L
In expression (12) it is assumed that there is no need for compression reinforcement in ULS.
Introducing Expression (12) into Expression (9), and developing, the slenderness limit can be obtained as shown in
Expression (13). The final expression proposed in Eq. (14) assumes that there is no compression reinforcement
since the effect of compression reinforcement on deflections is limited, and the increase in precision for this case is
not worth the complication.
L2
qqp kDL qRd kDL As f yd d 1 0,5
8
As' 0
8
5 kDL As f yd d L2 1 0,5 L
4
K 1 3
384 Ec ,eff bh
12 L
kI
h ' h a
As d As h d 2 (13)
k K Es 2 2 L
S cs cs Ec ,eff 1 3 8
bh
12
3
L Ec ,eff h 1
d 12akI d 5 E h h h
KkDL f cd 1 0,5 kS Kcs cs s 1 ' 1
48 8 2d 2d d
The final expression for the slenderness limit, assuming (’=0) is shown in (14).
3
L Ec ,ef h 1
5 E h
(14)
d kI 12a d
KkDL f cd 1 0,5 kS Kcs cs s 1
48 8 2d
The slenderness limits shown in Table 9.3 of FprEN 1992-1-1:2022 can be derived from the above expression.
[C09-6] Branson, D.E.: Deformations of concrete structures. McGraw-Hill, New York (1977)
References
[C09-7] Gilbert, R.I. Deflection Calculation for Reinforced Concrete Structures—Why We Sometimes Get It Wrong.
ACI Structural Journal. Vol 96, No. 6, pp 1027-1032, Nov. 1999.
[C09-8] Bischoff, P.H., Scanlon, A. Effective Moment of Inertia for Calculating Deflections of Concrete Members
Containing Steel Reinforcement and Fiber-Reinforced Polymer Reinforcement. ACI Structural Journal.
Vol. 104, No. 1, pp. 68-75, Jan-Feb 2007.
[C09-9] Pérez Caldentey, A., Mendoza Cembranos, J., Corres Peiretti, H.: Slenderness Limits for deflection
control: a new formulation for flexural RC elements. Structural Concrete (2016). DOI:
10.1002/suco.201600062
[C09-10] ACI Committee 318: Building Code Requirements for Structural Concrete (ACI 318-08) and commentary
(318R-08), American Concrete Institute, Farmington Hills, Mich., (2008), 430 p.
[C09-11] CEN: EN-1992-1-1. Eurocode 2. Design of concrete structures – Part 1-1. General rules and rules for
buildings, (2004).
[C09-12] fib: fib Model Code for Concrete Structures 2010. Ernst & Sohn. ISBN 978-2-88394-105-2 and ISBN 978-
2-88394-106-9.
[C09-13] Marí A.R.: Nonlinear geometric, material and time dependent analysis of three dimensional reinforced and
prestressed concrete frames. Report UCB-SESM-84/12. Berkeley, USA: Univ. of California; 1984
[C09-14] Ghali A. and Elbadry M.: User's manual and computer program CPF: Cracked plane frames in prestressed
concrete. Res. report CE85-2. Alberta, Canada: Dept. of Civil Eng., Univ. of Calgary; 1985
[C09-15] Ulm F.J., Clement J.L. and Guggenberger J.: Recent advances in 3-D nonlinear FE-analysis of R/C and
P/C beam structures. In: Proc. ASCE Struc. Cong. XII. 1994. p. 427-1433
[C09-16] Debernardi, G., Guglia, M. and Taliano, M.: Parametric analysis of the influence of shear on the deflection
of RC beams in service. Magazine of Concrete Research Volume 64, Issue 4, April 2012, pp 335-350.
[C09-17] CEB-FIP: Model Code 1990. Thomas Telford, 1993.
[C09-18] Ministerio de Fomento: EHE-08. Instrucción de Hormigón Estructural. EHE-08. (5ª edición) (2011).
[C09-19] Murcia Vela, J.: Cálculo práctico de flechas diferidas en estructuras de hormigón armado. Hormigón y
acero nº2015 (2000).
[C09-20] Mari, A., Bairán, J., Duarte, N.: Long-term deflections in cracked reinforced concrete flexural members.
Engineering structures.32 – 3, pp. 829 – 842.03/2010. ISSN 0141-0296
24.01.2023
For citations, copying and referencing see Note on Front page! page 406
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
10 and Annex E.4 Background to FprEN 1992-1-1:2023
Introduction
This document provides the background to clause 10 (Fatigue) and Annex E (Additional rules for fatigue
verification).
Chapter 10 of EC2-1-1
fatigue verification 10.2 Combinations of actions
6.8.3 Combinations of actions 10.3 Internal forces and stresses for fatigue
6.8.4 Verification procedure for reinforcing verification
and prestressing steel
- Steel, level 3 Simplified verifications (level 1)
10.4 Steel
6.8.5 Verification using damage equivalent 10.5 Concrete Compression
stress range 10.6 Concrete Shear
- Steel, level 2 10.7 Shear Interfaces
6.8.6 Other verifications
- Steel, level 1 E.1 General
6.8.7 Verification of concrete under damage equivalent stress range (level 2)
compression or shear E.2.1 General
- Concrete, compression, level 2
Annex E
E.2.2 Steel
- Concrete, compression, level 1 E.2.3 Concrete Compression
- Concrete, shear, level 1
explicit verification of damage strength (level 3)
EC 2-2 6.8
E.3.1 Verification conditions
- Concrete, level 3
E.3.2 Steel
6.2.5 (5) Shear at interface E.3.3 Concrete Compression
A s Ap
η
A s Ap ξ( s / p )
(C10.1)
The new equation (Eq. C10.2) is derived in agreement with clause 9.2.2(3) by decreasing the activated
prestressing steel area and calculating the stress range considering different inner lever arms of prestressing and
reinforcing steel.
s
Ae Ap ξ (C10.2)
p
Stress gradient
Clause 10.3(3) was added to take into account a redistribution of concrete stresses under compression for a given
stress gradient. Figure C10.2 shows the definition of the stress gradient and the design concrete stress c for
fatigue. The limit of c 2/3 c2 given in Figure C10.2 applies for fck = 55 MPa and increases to c 0,9 c2 for fck
= 100 MPa.
ck
𝑓cd,fat 𝛽cc 𝑡 ⋅ ⋅ 𝜂cc,fat (C10.4)
c
where
𝜂 , 𝑚𝑖𝑛 0,85 ⋅ 𝜂 ; 0,8 (C10.5)
⁄
𝜂 1 (C10.6)
Figure C10.3 depicts the development of fcd and fcd,fat according to the former and new versions of EN 1992-1-1.
90
fcd,fat fcd ‐ EN 1992‐1‐1:2004
fcd ‐ FprEN 1992‐1‐1
80
fcd,fat ‐ EN 1992‐1‐1:2004
fcd ‐ FprEN 1992‐1‐1
70 fcd,fat ‐ FprEN 1992‐1‐1
fcd ‐ EN 1992‐1‐1:2004
60
50
40
30
fcd,fat ‐ FprEN 1992‐1‐1
20
fcd,fat ‐ EN 1992‐1‐1:2004
10
fck
0
0 20 40 60 80 100 120
Figure C10.3: Development of fcd and fcd,fat according to FprEN 1992-1-1:2023 and EN 1992-1-1:2004
[C10.1] European committee for standardization (CEN): EN 1992-1-1, Eurocode 2: Design of concrete
References
structures Part 1-1, 2004.
[C10.2] Deutsches Institut für Bautechnik (DIBt): Richtlinie für Zulassungs- und Überwachungsprüfungen für
Spannstähle (guideline for approval and monitoring tests for prestressing steel), 2004. (in German)
[C10.3] Deutsches Institut für Bautechnik (DIBt): Ergänzung zur Richtlinie (supplement to guideline), 2014. (in
German)
[C10.4] Breedijk, Theo; Dieteren, Gerrie: Background document to the Dutch comments NL
118/119/126/127/128 on prEN 1992-1-1 draft D4-6 / D5-11 section 10.4, table C.1 and table E.1:
Fatigue strength reinforcing steel, 24.07.2020.
For citations, copying and referencing see Note on Front page! page 410
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
10.4 and C.4 and E.4 Background to FprEN 1992-1-1:2023
Introduction
This document provides the background to FprEN 1992-1-1:2023, clause 10.4, Annex C.4 (table C.1) and
Annex E.4 (table E.1).
Background
For above mentioned reasons over 500 fatigue tests have been performed on mainly B500B diameter 12 and
16 mm de-coiled bars from different mechanical straightening processes and cross-welded bars by resistance
welding and CO2 tack welding. The results show that the differences in straightening processes and cross
welding methods are limited and are being overruled by the scatter in the fatigue test results.
The fatigue tests are performed with constant stress amplitude with upper stress limit 0,6*fyk, load cycle
frequency 50 - 100 Hz and mostly stopped at N = 10 million load cycles (run-outs). The results are statistically
evaluated by 5 % linear regression from the failures (the k1 region) as well as by 5 % probabilistic approach with
Bayesian estimation from the failures + run-outs (the k1 region up to the horizontal CAFL) which provides better
insight on the knee-point N* (CAFL = constant amplitude fatigue limit). Further background is given in [1] and
[2].
As in EN1992-1-1:2004 the S-N curves are based on a fixed relation between k1 and k2 (Haibach – 1970): k2 =
2*k1 -1. For not welded bars there was a tendency for k1 = 4 but this would result in k2 = 7 which is too critical.
Therefore, for not welded bars k1 = 5 and k2 = 9 is maintained as in EN1992-1-1:2004. For welded bars, like in
EN1992-1-1:2004, k1 = 3 and k2 = 5 is maintained since the results for k1 did not show a need for adjustment.
The test results (see below) on the straightened bars show better fit with the results for N* = 2 x 106 than for N*
= 106 cycles, which therefore should be respected. On the welded bars the better fit for N* is somewhat higher
than 2 x 106. For the good sake of most testing in practice up to N = 2 x 106 and to prevent misunderstanding
between N* from the S – N curve and the N from the declared value by testing, it was proposed to apply N* = 2
x 106 for not welded and welded reinforcing steel. Also, already several countries are (were) used to specify
∆σRsk at N = 2 x 106 (UK at N = 5 x 106).
As expected, the test results on the mechanically straightened bars as well on the welded bars show for
diameter 12 mm relevant higher fatigue strength than for 16 mm which should be used in practice. For 16 mm
the test results on the mechanically straightened bars are (from the effect of de-coiling) somewhat lower and for
the welded bars about the same as the fatigue strength in EN 1992-1-1:2004 taken at N ≥ 2 x 106. Therefore, for
bars including de-coiled bars (are now ruling) 3 groups of bar sizes are proposed in order to use the best
possible properties. However the group of bigger bar sizes including de-coiled bar may interfere with the straight
produced bars which are not suffering from the mechanical straightening process. Therefore, arbitrary d = 20
mm is taken as the max. diameter for the mechanical straightening. The few possible straightening of bigger
sizes (up to 25 mm) should be addressed with special certification.
Conclusions
Based on the performed tests and evaluation of the results adjustment of the design parameters for S-N curves
is proposed as presented in Table C10.4.1.
Table C10.4.1 : Design parameters for S-N curves for carbon reinforcing steel (Table E.1 in FprEN1992-1-
1:2023)
By making the table a NDP different fatigue properties can be set in the National Annex if higher values of the
fatigue stress range and/or the number of cycles are confirmed by testing in accordance with EN 10080. It
should be noted that effect of de-coiling should be investigated if de-coiled bars are being applied.
Note 3 is added to table C.10.4.1 since the influence of welding is not present in the whole bar and in practice
this has been discussed many times.
To ensure the design parameters as presented in table C.10.4.1 the fatigue requirements in table C.1 of
FprEN1992-1-1:2023 are as follows:
Fatigue stress range 2σa [MPa] at N≥ 2 x 106 160 for bars and de‐coiled bars ϕ ≤ 12 mm
cycles based on testing with an upper stress limit 140 for bars and de‐coiled bars ϕ ≤ 16 mm
130 for bars and de‐coiled bars ϕ > 16 ≤ 20 mm
of 0,6*fyk (10% quantile)
130 for straight bars ϕ > 20 mm
100 for wire fabrics ϕ ≤ 12 mm
80 for wire fabrics ϕ > 12 mm
The adjustment of the design parameters for S-N curves also affects the simplified verification method
presented in 10.4. Required adjustments are as follows:
“a) reinforcing steel bars (for bent bars footnote a) of table E.1(NDP) should be applied):
- Δσsd ≤ 90 MPa unwelded reinforcing bars ø ≤ 12 mm
- Δσsd ≤ 73 MPa unwelded reinforcing bars ø > 12 mm
- Δσsd ≤ 40 MPa butt and tack welded reinforcing bars ø ≤ 12 mm
- Δσsd ≤ 30 MPa butt and tack welded reinforcing bars ø > 12 mm
- Δσsd ≤ 19 MPa couplers
Note: For practical reasons the split between diameter groups is limited at 12 mm.
Test results
Figure C10.4.1 Test results for 12 mm B500B mechanically straigthened bars tested with σmax= 0,6*fyk with
proposal FprEN1992-1-1:2023 (Proposal 160/2 – 5/9) and comparison with EN1992-1-1 :2004.
Figure C10.4.2 Test results for 16 mm B500B mechanical straigthened bars tested with σmax= 0,6*fyk with
proposal FprEN1992-1-1:2023 (Proposal 140/2 – 5/9) and comparison with EN1992-1-1 :2004.
Figure C10.4.3 Test results for 12 mm B500B welded bars tested with σmax.= 0,6*fyk with proposal
FprEN1992-1-1:2023 (Proposal 100/2 – 3/5) and comparison with EN1992-1-1 :2004.
Figure C10.4.1 Test results for 16 mm B500B welded bars tested with σmax= 0,6*fyk with proposal FprEN1992-1-
1:2023 (Proposal 80/2 – 3/5) and comparisson with EN1992-1-1 :2004.
[1] D.L. Allaix, F.B.J. Gijsbers Bayesian estimation of characteristics S-N curves for reinforcement bars
References
and proposal for the national annex of NEN-EN 1992-1-1, TNO-report TNO 2016 R10819, 21 June
2016
[2] T. Breedijk, New standard requirements on fatigue strength of reinforcing steel, Cement 5, 2020,
page 44 – 49 + annex page 1 – 5 (in Dutch).
For citations, copying and referencing see Note on Front page! page 416
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
11.3 Background to FprEN 1992-1-1:2023
Introduction This document has been prepared to clarify the background of the design provisions of FprEN 1992-1-1:2023
with respect to clause 11.3 concerning the permissible mandrel diameters for bent bars. These provisions update
current Section 8.3 of EN 1992-1-1:2004.
Reasons for change The current provisions for permissible mandrel diameters for bent bars of EN 1992-1-1:2004 (Section 8.3)
include a number of rules to suitably bend the reinforcement without damaging it or crushing the concrete inside
the bend. These rules have been identified to have two shortcomings:
1. Formula (8.1) of EN 1992-1-1:2004 relating the mandrel diameter and maximum stresses developed
in the reinforcement has been observed to have a poor performance to estimate the actual response
of bent details [C.11.3.1] (see also next paragraph).
2. Adapting the mandrel diameter according to the types of bends requires the use of different mandrels
for the same bar diameter. In addition, bending for large radii typically requires the use of machines
(based usually on the three-roller principle) different than those for normal mandrel diameters (where
the bars are bent around mandrels with Ømand = 4Ø or Ømand = 7Ø according to 11.3(2)). To avoid
these shortcomings and to increase the efficiency in bending the reinforcement, bends with larger
radii can be replaced by multiple bends with normal mandrels spaced by sufficiently long straight
segments. This requires a rational model to derive consistent rules easy to use.
Permissible mandrel Following these observations, a revision of the subsection has been performed. A detailed description of the
diameter, general case, theoretical principles and its implementation can be consulted in Monney et al. [C.11.3.1]. In this background
11.3(4) document, only the main assumptions and results will be highlighted.
The provisions of EN 1992-1-1:2004 provide a formula to evaluate the maximum force that can be developed in
a bar without damaging the concrete accounting for the bar diameter (Ø), the concrete strength (fcd), the concrete
cover (cd) and the mandrel diameter (Ømand). The origin of that formula is not completely clear, probably adapted
from a former British Standard [C.11.3.2] (see [C.11.3.1]). Although the basic parameters considered are
relevant, the performance of the expression is relatively poor, particularly for low Ømand /Ø ratios. Also, the
generality of the formula is penalized by the fact that it does not consider the bending angle of the detail. A
comparison of the performance of EN 1992-1-1:2004 with relevant test results is shown in Figure C11.3.1.
Figure C11.3.1: Performance of Formula (8.1) of EN 1992-1-1:2004 compared to tests (the first graph refers
to tests conducted by the authors, the second one shows tests found in the literature; the ratio
on the ordinates refers to the steel stress related to concrete crushing or spalling measured
in tests and the value calculated according to the code provision with average material
strengths and without partial safety factor), figure from Monney et al. [C.11.3.1].
In order to propose a more consistent formula for design, FprEN 1992-1-1:2023 has been based on a
mechanical model accounting for potential spalling failures. These failures, Fig. C11.3.2, occur due to the thrust
of a wedge of concrete forming at the bend region (blue region in Fig. C11.3.2). Such wedge is confined, thus
increasing significantly its resistance with respect to uniaxial conditions. The confinement pressures
(perpendicular to the bending plane) are in equilibrium with concrete tensile stresses developing in a confining
region (red region in Fig. C11.3.2a). As shown by detailed measurements (see [C.11.3.1]), a splitting crack
develops between the confined and the confining zone (white area between red and blue areas in Fig. C11.3.2).
Nevertheless, the load can be further increased and finally, failure occurs when the splitting crack develops in
an unstable manner leading to spalling of the concrete cover (Fig. C11.3.2).
The extent of the different areas has been observed by detailed measurement and analysis of test results
[C.11.3.1] to depend on a number of parameters:
Ratio between bar diameter and concrete cover
Figure C11.3.2: Mechanical model for FprEN 1992-1-1:2023: (a) plan view with confined (blue), confining
(red) and spalled areas (white); (b) detail of stresses and forces; and (c) detail of the concrete
wedge with the confinement stresses. (figure from [C.11.3.1])
This expression is the basis of Formula (11.1) of FprEN 1992-1-1:2023, where 2/ has been rounded to 0.65:
1/3
Ø f d c 1 Ø
σ sd 0, 65 f cd mand ck dg kbend 0.7 mand
(11.1)
Ø γC Ø
Ø 2 Ø
where k 45
bend 32
α
The performance of this Formula is compared in Fig. C11.3.3 to the same tests as those of Fig. C11.3.1. The
results show consistent agreement, suitably reproducing all design situations and a low scatter.
Figure C11.3.3: Performance of Eq. (C.11.3.1) (basis of FprEN 1992-1-1:2023) compared to test results,
figure from Monney et al. [C.11.3.1] (the first graph refers to tests conducted by the authors, the second one
shows tests found in the literature).
The linear Formula (11.1) can also be solved to determine directly the required mandrel diameter:
Ømand σ sd kbend σc,conf ,d (C.11.3.2)
Ø 0, 65 fcd 0, 7 σ c,conf , d
1/3
f d cd 1
where σ
c ,conf , d ck dg
γC Ø Ø 2
Simplified detailing rules, To enhance ease of use, simplified detailing rules are given in 11.3(3) for the cases where fyd ≤ 25ꞏfcd, covering
11.3(3) most practical cases. The verification of the concrete inside the bend may be omitted for all bends with an angle
bend ≤ 45° at a clear distance cx ≥ 2,5Ø from a free edge parallel to the bending plane and at a clear distance
between parallel bars cs ≥ 5Ø (equal to twice cx). As shown in Fig. C11.3.4 on the basis of Formula (11.1) of
FprEN 1992-1-1:2023, the limit cx ≥ 2,5Ø covers all practical cases. For diameter not larger than 32mm, this limit
could even be reduced to cx ≥ 2Ø.
Figure C11.3.4: Results of Formula (11.1) of FprEN 1992-1-1:2023 compared to the simplified detailing rule
cx ≥ 2,5Ø in 11.3(4).
Rules for multiple bends, Following the implications of Formula (11.1), it can be observed that the deviation angle of the bend (parameter
11.3(4) bend) plays a significant role. Small deviations allow in fact to develop large stresses, potentially ensuring
yielding of the reinforcement without cover spalling.
Based on this idea, bending of a bar can be performed as a succession of multiple smaller bends with the same
radius, see Figure C11.3.5. This concept allows to use the same mandrel as for normal bends (Ømand = 4Ø or
Ømand = 7Ø) and the same machine, allowing to bend the bars for all situations in a more efficient manner.
Figure C11.3.5: Performance of a detail with multiple bends, figure from Monney et al. [C.11.3.1].
In order to ensure suitable performance of the detail, two criteria shall nevertheless be respected:
Ensuring no damage of the reinforcement while bending of the bare bar (fulfilling the requirement of
11.3(2), namely Ømand = 4Ø for Ø ≤ 16 mm and Ømand = 7Ø for Ø >16 mm).
A minimum distance between the bends is provided (lmand in Fig. C11.3.5)
A general expression to calculate minimum values of lmand based on the mechanical model of Formula (C.11.3.1)
is derived in [C.11.3.1] and plotted in Fig. C11.3.6 for variable bending angle bend, bar diameter and concrete
cover for the cases where fyd ≤ 25ꞏfcd, covering most practical situations. Such expression has been simplified
into FprEN 1992-1-1:2023 where the distance between the bends should not be shorter than 4Ø.
Figure C11.3.6: Minimum distance between bends lmand as a function of: (a) the bending angle; and (b) the
concrete cover with the detailing rule of FprEN 1992-1-1:2023.
Simplified rules for According to EN 1992-1-1:2004, the verification of the concrete strength inside the bend of hook and bend
standard hook and bend anchorages may be omitted if (i) the anchorage of the bar does not require a length more than 5Ø after the end
anchorages, of the bend and (ii) the bar is not positioned at the edge (plane of bend close to concrete face). Based on
theoretical considerations using Formula (C.11.3.1) confirmed by tests (see Figure C11.3.7, [C.11.3.1]), these
11.3(3) requirements can be mitigated by defining a minimum clear distance cx ≥ 1,5Ø between the bar and a free edge
parallel to the bend. The reduction of cx compared to normal bends can be explained theoretically by the fact
that the tensile stress is lower at the start of the bent because a part of the tensile force is transferred by bond.
Figure C11.3.7: Maximum tensile stress at failure for different hook and bend anchorages (tensile stress
measured at the start of the bent, see definition of lb).
Enhancement in case of If transverse bars are placed within the bend, a portion of the deviation forces can be transferred to the
transverse bars placed transverse bars as shown in Figure C11.3.8 below. In this case, the limit to the steel stress sd in Formula (11.1)
within the bend may be increased by the factor defined in Formula (11.2) (see [C.11.3.3] for its development).
11.3(5)
Figure C11.3.8: Effect of transverse bars placed inside the bend (figure from Monney [C.11.3.3]
References [C.11.3.1] Monney, F.; Fernández Ruiz, M.; Muttoni, A., “Design against splitting failures in reinforced concrete
due to concentrated forces and minimum bend diameter of reinforcement”, Engineering Structures,
245 (2021) 112902, p. 21
[C.11.3.2] BS 8110, Structural use of Concrete - Code of Practice for Design and Construction, British
Standards Institution, Part. 1, London, UK, 1985.
[C.11.3.3] Monney F., “Revisiting Detailing Rules for Bending Bars, Anchorage and Minimum Shear
Reinforcement in Concrete Structures”, PhD Dissertation, École Polytechnique Fédérale de
Lausanne (EPFL), Switzerland, 2022, p. 184.
For citations, copying and referencing see Note on Front page! page 422
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
11.4 and 11.5 Background to FprEN 1992-1-1:2023
Introduction
This document provides the background to clauses 11.4 (“Anchorage of reinforcing steel in tension and
compression”) and 11.5. (“Laps of reinforcing steel in tension and compression and mechanical couplers”).
However, mechanical couplers (Clause 11.5.6) are not covered in this paper.
Please note that background to the following clauses is provided in separate background documents:
- 11.4.7 “Anchorage of headed bars”;
- 11.5.4 “Laps using U-bar loops”:
- 11.5.5 Laps using headed bars;
- 11.6.3 “Minimum radius of curvature and straight length of tendons adjacent to anchorages”;
- 11.7 “Deviation forces due to curved tensile and compressive chords”..
1 11 1 1
fc 4 lb 20 25 mm 5 cd 4
sd 54 (C11.1)
25 MPa
Where cd is defined in Figure C11.1a.
Figure C11.1: (a) definition of nominal cover cd; (b) non-uniform distribution of bond stresses and (c) nonlinear
distribution of reinforcement force in the development length compared to the usual assumption of a constant
bond stress along the development length (dotted blue lines)
It can be noted that the influence of the anchorage length ℓb is not linear as it accounts for the non-uniform
distribution of bond stresses along the development length as well as a strain effect (decreased bond strength
as a function of reinforcement strain [C11.2, C11.3], see figures C11.1b and c).
Equation (C11.1) can be rewritten to express directly the anchorage length as a function of the stress in the
reinforcement s and other relevant parameters:
11 1 1 1
lb 20 25 MPa 4 s 5 4
f (C11.2)
c 54 MPa 25 mm cd
or
5 20 4 5
25 MPa 11 s 11 11 11
lb (C11.3)
fc 54 MPa 25 mm cd
As the original equation has a semi-empirical origin, it is reasonable to round the exponents of the design
equation to simpler fractions. In addition, reference values for sd, and cd are adapted to correspond to the limit
values for the simple table (clause (2)). Finally, when switching from mean test values to design level, design
stresses and characteristic concrete strength are used and a calibration parameter klb is added to arrive at the
format used in Formula (11.3) in FprEN 1992-1-1:2022 (Note: coefficient for casting position kcp is not shown in
(C11.4), see text to Clause 11.4.2(4) below):
1 3 1 1
25 MPa 2 sd 2 3 1.5 2
lbd klb (C11.4)
fck 435 MPa 20 mm cd
Some of the terms in this equation have to be limited due to insufficient validation by testing (lack of experimental
data) or because of a mechanically justified threshold. According to [C11.1], the following limits should be
defined for calculation purposes:
- 15 MPa ≤ fcm ≤ 110 MPa: An upper limit has been introduced for fcm in FprEN 1992-1-1:2022 with the
ratio (25/fck) 0,3 since Formula (11.3) tends to underestimate anchorage length for concrete
strengths fcm 90 MPa. The lower limit has been neglected as it applies mostly to existing structures
(potentially to be accounted for in old existing structures);
- 0.5 ≤ cd/ ≤ 3.5: the lower limit is neglected because it is already covered by clause 6.5.2.3; the upper
limit is rounded to 3.75 (1.5 / cd ≥ 0.4);
- ≥ 12.5 mm: according to [C11.1], this limit reflects experimental evidence (probably reflecting a low
relative rib area for small bars). Although, based on Figure 3-12 in [C11.1], this limit could probably
be removed, it has been fixed to /20 mm ≥ 0.6 in FprEN 1992-1-1:2022.
As shown in the figure C11.2 below, the rounding of the exponents of Eq. (C11.4) has only a low to negligible
influence on the results. The same diagrams show also the trends when compared to EN 1992-1-1:2004
(EC2:2004). It has to be observed that only the trends are compared, not the absolute values of the anchorage
lengths, as parameter klb in Expression C11.4 is subject to calibration against test data for the relevant design
situation, see below.
a) b)
c) d)
Figure C11.2: Influence of a) concrete strength, b) bar diameter, c) developed tensile steel stress and d)
concrete cover on the anchorage length (note: only the influence is compared, not the absolute values as
parameter klb in Expression C11.4 is subject to calibration)
Figure C11.2b) shows that according to fib MC2010 (and FprEN 1992-1-1:2022), the influence of bar diameter
on the anchorage length is more than linearly proportional (influence of the size effect on the bond strength).
With respect to the strain effect and the influence of non-uniform bond stress (neglected in EC2:2004), Figure
C11.2c) shows that according to fib MC2010 and FprEN 1992-1-1:2022, relative to the requirement for a bar
stress of 435MPa, the anchorage length can be substantially reduced for moderate and low stress levels. This
could explain why only few problems related to insufficient anchorage length have been reported so far
(anchorages and laps are in practice only occasionally subjected to high stress levels).
Comparisons
Figure C11.3 shows a comparison of the required anchorage length according to EC2:2004 and FprEN 1992-1-
1:2022 with klb = 50 (assumed parameters are defined in the figure caption). The required anchorage length at
reinforcing steel stresses sd 350 MPa is comparable. For higher reinforcing steel stresses, FprEN 1992-1-
1:2022 requires longer anchorage lengths. For sd = 435 MPa, the same required anchorage length as given in
EC2:2004 would be obtained for these parameters with klb = 39.
Characteristic strength of ribbed bars was around 400MPa when MC78, which forms the basis of current EC2
design rules, was drafted. The main steel Grade in current practice is now 500MPa, an increase of 25%. To
maintain a consistent level of safety with that when these rules were originally compiled, anchorages for Grade
500 bars should now be (500/400)1.5 times longer than for Grade 400 bars, according to Expression [C11.3],
corresponding to a 40% increase in anchorage and lap length. The increase in bond length required by current
EC2 provisions has been only (500/400)1.0, i.e. 1.25 times longer. With the introduction of higher Grade 500 bars
a further 12% increase in anchorage and lap lengths over current values should have been introduced to
maintain the level of safety intended when current provisions for bond lengths were formulated. Based on
[C11.1], the required further increase would be 20%. This should be kept in mind when interpreting Figure C11.3.
Also shown in Figure C11.3 is the required anchorage length according to prEN 1992-1-1:2021. The required
anchorage length is identical to the values of FprEN 1992-1-1:2022 for reinforcing steel stresses sd 435 MPa.
However, below that stress the required anchorage length was longer because an exponent on the reinforcing
steel stress of n = 1,0 was used in prEN 1992-1-1:2021.
For the actual calibration of klb versus test results as a function of the reliability factor , see specific paragraph
“Test results and calibration of klb against test data” below.
Figure C11.3: Comparison of the required anchorage length according to EN 1992-1-1:2004 and FprEN 1992-
1-1:2022 (fck = 30 MPa, = 16 mm, cd/ =1.5, αct=1.0,C=1.5, good bond conditions and no confinement)
Design example
The anchorage length of previous example (fck = 30 MPa, = 16 mm, cd/ =1.5, αct=1.0, C=1.5, good bond
conditions and no confinement) is designed for sd = 435 MPa:
According to EN 1992-1-1:2004 According to FprEN 1992-1-1:2022
This example shows that FprEN 1992-1-1:2022 is significantly easier to use. Already the first calculation (clause
(2)) allows for a fast and accurate result. If it were important, refinements using Clause (3) or Clauses (3) and
(5) allow for a reduction of the required anchorage length under certain conditions.
2 1 2 1
l 3 fck 3 20 mm 9 cd 3
sR 435 MPa b (C11.5)
klb 25 MPa 1.5
The concept of bond strength fbd has been used in the past in order to dimension the anchorage length on the
basis of equilibrium and assuming a constant value of the bond stress throughout the development length:
2
fbd lb s (C11.6)
4
Conversely, Eq. (C11.6) allows also determining an average bond strength from Eq. (C11.5):
MPa mm
𝑓 ⋅ (C11.7)
⋅ MPa .
Or as a function of the steel stress sd:
MPa mm
𝑓 ⋅ (C11.8)
35 MPa MPa .
The physical meaning of this average strength is represented in Figures C11.1b and c (dotted blue lines). As
shown in Figure C11.4, the average bond strength is very sensitive to the anchorage length or to the maximum
steel stress. This confirms that this value cannot be used for dimensioning the anchorage length, or only
assuming a sufficiently conservative value which can be developed for long anchorages.
Figure C11.4: Average bond strength as a function of (a) the anchorage length and (b) the maximum steel
stress (klb = 40)
Figure C11.4a also confirms that high average bond stresses may be developed in short anchorages. Assuming
for instance lb=5ꞏ (typical bond tests on short anchorages), the average bond strength becomes for klb = 40:
1 2 1
f ck 3 20 mm 9 cd 3
f bd 5.4 MPa (C11.9)
25 MPa 1.5
which gives higher values than generally assumed (dot in Figure C11.4a).
During the first hours after casting, fresh concrete behaves like a saturated soil [C11.4] and consolidates. Water
11.4.2(4)
in excess migrates from the bottom fresh concrete towards the top (so-called “concrete bleeding”) and the
Good and poor bond concrete settles (between 0 and 3 mm for a fresh concrete depth of 1 m). As a consequence of concrete bleeding
conditions – effect of and concrete settlement, porous concrete develops under the reinforcement bar and a void can even form in
casting position case the reinforcement bar is restrained to settle with the concrete [C11.5]. The water /binder ratio also increases
near the top of the fresh concrete and reduces tensile strength of the concrete cover. These phenomena, which
are more pronounced in the upper part of the concrete members, are associated to a loss of the bond properties
(effect known as “top bar effect” or “poor bond condition”).
EC2:2004 accounts for this effect with parameter 1 which applies to the design bond strength (1 =1.0 for good
bond conditions, 1 =0.7 for poor bond conditions). ACI 318 applies a similar factor (1.3), but to the development
length. A recent literature review [C11.6] has confirmed this effect, but some findings are still to be verified.
Recent tests have shown a significant reduction of the anchorage strength already for moderate depths [C11.7]
which is potentially related to high consistency class. The influence of the concrete consistency has been verified
by further tests, see clause “Test results” below. In addition, it has to be noted that the governing failure mode
observed in the [C11.7] tests is pull-out. Work is ongoing on the verification whether the effect of poor bond
conditions in case of splitting or spalling failures is the same.
For these reasons, in FprEN 1992-1-1:2022, the concept of EN 1992-1-1:2004 is kept, but its parameters have
been adjusted and recalibrated based on recent testing [C11.9]. A coefficient kcp has been introduced in Formula
(11.3) of FprEN 1992-1-1:2022 to account for good and poor bond conditions directly in the formula. The
coefficient takes the following values:
kcp = 1,0 for good bond conditions
kcp = 1,2 for poor bond conditions and for all bars used in slipform construction (unless it is shown that the vertical
bars cannot move during casting)
kcp = 1,4 for bars executed under bentonite or similar slurries (unless data is available for the specific slurry
used).
The factor kcp = 1.20 for ‘poor’ bond (accounting for the effects of casting on bond) is based on work by Suryanto,
Aitken and Cairns [C11.9], where it was concluded that for a member with overall depth greater than 250 mm a
top cast factor of 1/0.85 could be considered for concrete with a consistency S2 and 1/0.75 for S5 concrete. A
factor of 1.20 (=1/0.83) aligns with usual design assumption that S2 consistence (50 – 90 mm slump) concrete
will be used.
Slipform construction is aligned with ‘poor’ bond conditions. There are perceived issues of long unsupported
vertical bars and dragging of concrete over horizontal bars during construction. No test data is known.
The provision for bentonite slurry walls is adapted from MC2010 and Jones and Holt [C11.19]. This paper
advocates a factor of 1.4 - the contemporary factor pertaining to ‘poor’ bond.
Further recent testing [C11.10] has confirmed that indented bars show similar difference to ribbed bars between
good and poor bond conditions, see clause “Test results” below.
crack parallel to the free surface as shown in Fig. C11.5a becomes governing and the transverse reinforcement
cannot be activated. To account for this effect, the potential splitting cracks are shown in the code figure. The
parameter nc (number of legs of reinforcement providing reinforcement and crossing the potential splitting crack)
as well as parameters for transverse reinforcement in the design equation are defined in a different manner from
fib MC2010 and EN 1992-1-1:2004. An active pressure acting on the free surface (Figure C11.5c) can also
provide confinement and can be accounted for in a similar manner.
Figure C11.5: Definition of cases where the anchorage length may be reduced due to confinement provided
by (a) confinement reinforcement (nc=1, nb=2); (b) transverse reinforcement and (c) external pressure
fib MC2010 accounts for confinement by increasing the design bond strength (Eq. 6.1-21 of fib MC2010 with
slightly different symbols and definitions, influence of cmax neglected and influence of minimum transverse
reinforcement neglected as it is not to be considered in the calibration of klb):
⁄ .
𝑓 𝑘 𝜌 𝑓 , 2𝜎 2𝑓 , 0,4𝜎 𝑓 (C11.10)
Where ρconf is the ratio of the reinforcement providing confinement referred to the diameter of the bar to be
anchored or spliced:
∑ ,
𝜌 (C11.11)
⋅ ⋅ℓ ⋅ ⋅
2𝑡 𝑐𝑠
𝑐 , min 𝑐𝑥 ; 𝑐𝑦 25 ;
𝑠𝑡 2
; 3.75 𝑐𝑑 6 (C11.12)
where 𝑐 70𝜌 12 𝜎 ⁄ 𝑓 ⋅
Parameters 70 and 12 in cd have been determined based on calibration against test results. The limit of cd,conf
6ϕ for cases with confinement reinforcement or with compression stress perpendicular to the potential splitting
failure was confirmed based on the same test results. Formula (C11.12) is considered much simpler to use with
little loss of accuracy. Figure C11.6 shows the influence of confinement reinforcement on the required anchorage
length for an example.
Figure C11.6: Required anchorage length as a function of the amount of confinement and transverse
reinforcement (fck = 25 MPa, bar to be anchored = 20 mm, cd = 30 mm, klb = 50, good bond conditions and
designed for sd = 435 MPa, nb = 1, nc = 1 for confinement reinforcement)
Note: nb=1 and nt=1 was used to determine conf = (ntt2/4nbst) for transverse reinforcement; and nb=1 and
nc=1 was used to determine conf = (ncc2/4nbsc) for confinement reinforcement
For bars in compression, a reduction of the anchorage length by 15 may be taken into account if the end face
11.4.2(6)
of the bar is at a distance of at least 5 from the free concrete surface in the direction perpendicular to the end
Anchorage of bars in face of bar to avoid punching of the cover in front of the bar end face. However, an absolute minimum anchorage
compression length of 10 applies in FprEN 1992-1-1:2022.
These provisions for transverse and confinement reinforcement for the anchorage of straight bars 32 mm
11.4.2(7)
correspond to the provisions in EN 1992-1-1:2004, Clause 8.8(6). However, the reinforcing steel area As has
Transverse and been replaced by (2/4).
confinement
reinforcement for
anchorage of bars 32
mm
These provisions for anchorage of bundles are consistent with EN 1992-1-1:2004, Clause 8.9.1 and 8.9.2.
11.4.3
However, while EC2:2004, 11.4.3, implicitly assumed all bars within a bundle were of the same size. FprEN
Anchorage of bundles 1992-1-1:2022 makes provision for bundles comprising bars of mixed size.
Deductions for standard bends and hooks in tension are in accordance with fib MC2010 Cl 6.1.3.4 and 5.
11.4.4(1), (3)
Where cover is less than 3.5ϕ, the effect of bends and hooks on the anchorage capacity of bars in compression
Anchorage of bars with
should generally be neglected because of the risk of concrete spalling due to transverse deflections of the bar
bends and hooks in
tension and compression close to a bend or hook in compression. FprEN 1992-1-1:2022 permits an exception when all free concrete
surfaces, perpendicular and parallel to the bar, have a distance of at least 3,5 from the bar surface and end
face. If those distances are satisfied spalling is prevented and the same reduction of the anchorage length by
15 applies as for straight bars in compression, however, to the straight projected length to the bend or hook,
with the same absolute minimum anchorage length of 10.
Anchorage of bars with welded transverse reinforcement in current EN 1992-1-1:2004 has the same anchorage
11.4.5
length reduction factor as given for standard hooks. Hence, in FprEN 1992-1-1:2022 the same reduction of
Anchorage of bars with the anchorage length by 15 is permitted as for bars with hooks under the specified conditions for the size and
welded transverse number of transverse bars, however, with an absolute minimum anchorage length of 5.
reinforcement
The minimum requirements for the strength of the weld are given in Annex C.4.1(3).
For U-bar loops subject to pure tension, i.e. both legs having the same tension force, anchorage may be
11.4.6
assumed to be provided at the end of the bends if the loop details (bends) comply with Clause 11.3 of FprEN
Anchorage of U-bar loops 1992-1-1:2022 for permissible mandrel diameters.
Alternatively, or for the case where the legs of the loop have different forces, the anchorage length may be taken
as the one of straight bars according to (11.3) but reduced by 20 but with an absolute minimum anchorage
length of 10. The bends have to comply with minimum permissible mandrel diameter.
If, according to clause (4), laps are located away from sections where inelastic deformations in the reinforcing
11.5.2(4), (5)
steel (plastic hinges) are expected to occur, laps may be detailed with up to 100% of bars lapped at any section
Laps and they may be designed for the maximum value of the reinforcement design stress sd. The same applies for
laps with bars in compression.
In cases where the previous requirement is difficult to satisfy, clause (5) allows design of laps which are located
near sections where inelastic deformations in the reinforcing steel are expected to occur. In these cases, the
designer has basically three options:
i. a confinement reinforcement is arranged according to 11.5.2(8); or
ii. the laps are staggered; or
iii. the laps are designed for a design stress 1.2 times the design value of the reinforcement stress.
Exceptions i. and ii. are justified by the fact that confinement reinforcement and staggering the laps reduce the
brittleness of lap failure and better maintain residual strength after peak load [C11.8], [C11.15]. The percentage
of lapped bars allowed depends on the member type (35% for linear members and 50% for planar members).
This is justified by the increased redundancy of planar members.
The factor 1.2 applied in exception iii. is justified by the fact that plastic deformations of the reinforcement lead
to a reduction of the lap resistance. Alternatively, an increase by a factor of 1.2 results in reinforcement stresses
near the tensile strength.
The required lap length in FprEN 1992-1-1:2022 is now equal to the required anchorage length multiplied by a
11.5.2(2) and Table 11.3
factor kls=1,2 (NDP), i.e. a calibration coefficient of (klb x kls). This is different to fib MC2010 and is based on
Required lap length for recent individual, separate calibration of tests on anchorages and tests on laps. This recent calibration work also
bars in tension and uses a single exponent n = 1,5 for the entire range of reinforcement stresses sd in Formula (11.3). Since the
compression value of coefficient klb depends on the choice of n in Formula (11.3) both values klb and n are now NDPs.
Provisions of EN 1992-1-1:2004 have been simplified addressing several Systematic Review Comments (SRC).
11.5.2(8) and (9)
For compression laps, at least one transverse bar or confinement link should be placed at each end of the lap,
Reinforcement providing within a maximum distance of 50mm or 2, from the ends of the lap. This is required to control potential cracking
confinement in laps of due to the introduction of compression forces by end bearing from the bar into the concrete.
straight bars
Figure 11.12 has been modified showing the difference between transverse reinforcement (Figures 11.12a and
11.12b) and confinement reinforcement (Figures 11.12c and 11.12d). In addition, two options are given
(reinforcement providing confinement distributed over the lap length (a) and c)) or concentrated at the ends of
the spliced bars (b) and d)).
Design example
The tension lap length for the beam represented below with fck = 30 MPa, = 25 mm, cnom = 25 mm, αct=1.0,
C=1.5, good bond conditions is designed for sd = 435 MPa assuming that the lap is located in an area with
potentially inelastic deformations in the reinforcing steel. It is assumed, that the laps are not staggered.
In a first step, the lap length is calculated neglecting the confinement provided by the stirrups:
According to EN 1992-1-1:2004 According to FprEN 1992-1-1:2022
In a second step, the confinement effect provided by the stirrups is accounted for. The calculation according to
FprEN 1992-1-1:2022 is performed designing two options: first considering the given stirrups ϕ12@150 and as
a second option, by arranging sufficient confinement reinforcement according to 11.5.2(8) so that the lap may
be designed without increasing the design stress (1st option according to 11.5.2(5)).
For cases where bars 20 mm are lapped or the percentage of lapped bars is 25%, a minimum amount of
11.5.2(10)
transverse or confinement reinforcement should be provided which corresponds to the value specified in the
current EN 1992-1-1:2004.
The provisions for lapping bundles of bars of EN 1992-1-1:2004 are maintained with two modifications:
11.5.3
- there is provision for bundles of mixed diameter;
Laps of bundles
- there is no distinction anymore between laps with bars smaller and laps with bars equal or larger than
32mm diameter consistent with the remainder of FprEN 1992-1-1:2022;
- lap length of bundles with 2 or 3 bars may be calculated based on individual bar diameter either when
laps are staggered or when an additional “lap-bar” is added, [C11.16].
Calibrations were performed for FprEN 1992-1-1:2022 Formulae (11.3) and (11.4) at RWTH Aachen, [C11.22]
Test results and
and [C11.23], and at EPFL, [C11.24], each considering anchorages and laps separately. These calibrations
calibration of klb against
were based on test data compiled by ACI 408 supplemented by additional results added by TG2.5 [C11.17],
test data
from RWTH Aachen and EPFL for comparison with the strength predicted by FprEN 1992-1-1:2022 Formulae
(11.3) and (11.4). For these calibrations, the reliability-based methodology of Mancini et al [C11.13], was used
to analyse the data.
Calibrations at EPFL and at RWTH Aachen performed for anchorages found a mean value of coefficient klb,m =
22 and recommended for design klb = 50 and klb = 55, respectively, for a target reliability factor of = 3,8, see
Figure C11.7. A coefficient of klb = 50 was retained as recommended values in FprEN 1992-1-1:2022.
Figure C11.7: Test results for anchorages compared against theory mean values and design values
Figure C11.8 which is based on data from [C11.11] illustrates that the modifications in FprEN 1992-1-1:2022
Formula (11.4) for the effect of confinement reinforcement and compression stress in cd may be used up to
cd,conf = 6.
Figure C11.8: Test results for the effect of confinement reinforcement and compression stress on anchorage
Calibrations at EPFL and at RWTH Aachen performed for laps found mean values of coefficient klb,m = 32 and
klb,m = 33 respectively, and recommended for design klb = 60 and klb = 70, respectively, for a target reliability
factor of about = 3,8, see Figure C11.9. A coefficient of klb = 60 was retained as recommended values in FprEN
1992-1-1:2022. Note: the coefficient for laps was renamed in FprEN 1992-1-1:2022 to the product (klb kls).
Figure C11.9: Test results for laps compared against theory mean values and design values
Table C11.1 summarises the results of the above referenced calibrations for anchorages and laps in terms of
coefficient klb and (klb kls), respectively, as a function of the reliability index .
Mean Characteristic = 3.8
Anchorages klb 22 35 50
Table C11.1: Coefficients klb and (klb kls) as a function of reliability index
The effect of good/poor bond conditions (top cast factor) was recently investigated again, [C11.9] and [C11.12].
Effect of casting position
Main conclusions of this research were that the overall definition of good/poor bond conditions according to EN
1992-1-1:2004, maintained in FprEN 1992-1-1:2022, is reasonable. The coefficient kcp = 1,2 is reasonable or
slightly conservative for low slump concretes, however, slightly non-conservative for high slump concretes.
However, for practical reasons, a single value for kcp is maintained in FprEN 1992-1-1:2022, see Figure C11.10.
Further research by Cairns [C11.14] has demonstrated the top cast effect diminishes as lap length increases.
Results in Fig C11.10 were obtained from relatively short (lb/ϕ = 5) anchorages and are therefore conservative
for practical construction where a minimum lap length of 10ϕ applies.
Top cast factor in low slump concrete (C30, S2) Top cast factor in high slump concrete (C30, S5)
Figure C11.10: Top cast factor as a function of distance from top surface and concrete slump
Note: A comparison of top cast factors between ribbed and indented reinforcing steel bars is given in [C11.10].
The effect of fatigue loading on bond performance has been investigated e.g. in [C11.20]. It has been found that
Effect of fatigue loading the fatigue strength of bond as a function of the number of load cycles follows closely the fatigue strength of
concentrically loaded concrete under compression, see Figure C11.11. This means that no fatigue bond failure
will occur during several million load reversals if for the usual anchorage lengths required for reinforcing steel
bars the upper load (reinforcing steel stress) is smaller than about 50 percent of the static ultimate load of the
specific anchorage length. A previously applied cyclic load does not negatively affect the deformation behaviour
of the anchorage near failure or the ultimate load compared with monotonic load increase in a static bond test
providing failure does not occur during the fatigue test.
Figure C11.11: Fatigue strength of bond as a function of the number of load cycles, [C11.20]
Recent fatigue testing of bond [C11.21] has confirmed the above results and conclusions. Repeated loading
does not negatively affect the ultimate strength, provided that the anchorage does not fail under fatigue loading.
During fatigue loading the peak bond stresses move towards the unloaded end of the bar with increasing number
of load cycles, see Figure C11.12.
Figure C11.12: Distribution of steel stresses and strains over bond length at upper and lower load in fatigue
tests as a function of load cycles acc. to [C11.21].
Based on this favourable behaviour of anchorages, verification of anchorages and laps under fatigue loading
was not specifically covered in FprEN 1992-1-1:2022.
Comparison of lbd for A general trend for the comparison of anchorage lengths between EN 1992-1-1:2004, prEN 1992-1-1:2021 and
anchorages according to FprEN 1992-1-1:2022 can be observed already in Figure C11.3.
EN 1992-1-1:2004, prEN Figures C11.13 and C11.14 show detailed comparisons of design values of anchorage length, lbd, calculated
1992-1-1:2021 and FprEN according to EN 1992-1-1:2004, prEN 1992-1-1:2021 and FprEN 1992-1-1:2022 for klb = 50 in poor and good
1992-1-1:2022 bond conditions, respectively, for two values of cover cd and three different bar diameters.
= 16 mm
= 25 mm
Figure C11.13: Comparison of anchorage length lbd between EN 1992-1-1:2004, prEN 1992-1-1:2020-11 and
FprEN 1992-1-1:2022 with coefficient klb = 50 for cd = 1,5 in good and poor bond conditions
= 16 mm
= 25 mm
Figure C11.14: Comparison of anchorage length lbd between EN 1992-1-1:2004, prEN 1992-1-1:2020-11 and
FprEN 1992-1-1:2022 with coefficient klb = 50 for cd = 3,75 in good and poor bond conditions
FprEN 1992-1-1:2022 results in larger anchorage lengths compared with EN 1992-1-1:2004 for large diameter
bars and steel stresses in the order of 435 MPa or above, in particular in good bond conditions. Comparable or
smaller anchorage lengths result for steel stresses of sd 300 MPa. Significantly smaller anchorage lengths
are obtained with FprEN 1992-1-1:2022 for small bar diameters, small steel stresses and poor bond conditions
combined with large cover.
A general trend for the comparison of lap lengths between prEN 1992-1-1:2021 and FprEN 1992-1-1:2022 can
Comparison of lsd for
be observed in Figure C11.15: Lap lengths for steel stresses below 300 MPa are smaller due to the effect of the
laps according to EN
1992-1-1:2004, prEN unique exponent n = 1,5, above that value the lap lengths are larger due to the coefficient kls = 1,2. At a steel
1992-1-1:2021 and FprEN stress of 300 MPa the lap lengths of prEN 1992-1-1:2021 and FprEN 1992-1-1:2022 are identical.
1992-1-1:2022
Figure C11.15: Comparison of lap lengths lsd between prEN 1992-1-1:2020-11 and FprEN 1992-1-1:2022
Figures C11.16 shows detailed comparisons of design values of lap length, lsd, calculated according to EN 1992-
1-1:2004, prEN 1992-1-1:2021 and FprEN 1992-1-1:2022 for (klb kls) = 60 in poor and good bond conditions,
respectively, for a cover cd = 1,5 and three bar diameters.
= 12 mm
= 16 mm
= 25 mm
Figure C11.16: Comparison of lap length lsd between EN 1992-1-1:2004, prEN 1992-1-1:2020-11 and FprEN
1992-1-1:2022 with coefficient (klb x kls) = 60 for cd = 1,5 in good and poor bond conditions
Note: Plots are shown for 6 = 1.15 where up to 1/3rd of bars are lapped at a section and for 6 = 1.5 where 50%
or more of bars are lapped at a section
Except for large bar diameters with high steel stresses sd 435 MPa, lap lengths according to FprEN 1992-1-
1:2022 are smaller than those according to EN 1992-1-1:2004 with 6 = 1,5. For steel stresses sd 300 MPa,
lap lengths according to FprEN 1992-1-1:2022 are generally smaller or comparable to EN 1992-1-1:2004 even
for values 6 < 1,5.
[C11.1] fib, Bond and anchorage of embedded reinforcement: Background to the fib Model Code for
References
Concrete Structures 2010, fib Bulletin 72, 2014, 161 p.
[C11.2] Fernández Ruiz M., Muttoni A., Gambarova P., Analytical modelling of the pre and post-yield
behaviour of bond in reinforced concrete, ASCE Journal of Structural Engineering, Vol. 133, N° 10,
2007, pp. 1364-1372.
[C11.3] Fernández Ruiz M., Muttoni A., Gambarova P., A re-evaluation of test data on bond in R/C by
means of FEM modelling, Studies and researches, V. 27, 2007, pp. 113-134.
[C11.4] Clear C.A., Bonner D. G., Settlement of fresh concrete – an effective stress model, Magazine of
Concrete, 1988, pp. 3-12
[C11.5] Castel A., Vidal T., Viriyametanont K., François R., Effect of Reinforcing Bar Orientation and
Location on Bond with Self-Consolidating Concrete, ACI Structural Journal, 2006, pp. 559-567
[C11.6] Cairns J., Influence of casting position on bond strength, Joint meeting PT.SC2.T1 and AhG
Detailing, Lausanne, 10.1.2017
[C11.7] Metelli G., Marchina E., Plizzari G.A., Effect of casting position on bond, 4th Meeting of fib TG 2.5,
Stuttgart, 25.9.2017.
[C11.8] Metelli G., Cairns J., Plizzari G., The influence of percentage of bars lapped on performance of
splices, Materials and Structures, 2015, pp 2983–2996
[C11.9] Suryanto B., Aitken M., Cairns J., Influence of casting position on anchorage strength of steel
reinforcement, Heriot-Watt University, April 2020
[C11.10] Suryanto, Aitken & Cairns, Influence of Casting Position on Anchorage Strength of Indented Steel
Reinforcement by Heriot-Watt University, December 2020
[C11.11] Amin, Results from database compiled for anchorage at end supports, South Bank University,
London, 2009
[C11.12] Moccia F., Fernandez-Ruiz M., Metelli G., Muttoni A., Plizzari G., Casting position effects on bond
performance of reinforcement bars, fib Structural Concrete, Vol. 22, January 2021, 21 p.
[C11.13] Mancini G., Carbone V.I., Bertagnoli G., Gino D., Reliability-based evaluation of bond strength for
tensed lapped joints and anchorages in new and existing reinforced concrete structures, fib Structural
Concrete, Vol. 19, June 2018, pp. 904-917
[C11.14] Cairns J., Top cast effect: the influence of bond length Structural Concrete 2021;1–14 DOI:
10.1002/suco.202100376
[C11.15] Cairns J. Staggered lap joints of tension reinforcement. Structural Concrete, Vol 15 no. 1. March
2014. DOI 10.1002/suco.201300041
[C11.16] Cairns J., Lap Splices of Bars in Bundles. ACI Structural Journal. Vol 110, no. 2. Mar 2013
[C11.17] fib Task Group 4.5 "Bond models", Splice test database obtained from
http://fibtg45.dii.unile.it/files%20scaricabili/Database_splicetest%20Stuttgart %20sept%202005.xls;
2005
[C11.18] Micallef M, Vollum RL. The behaviour of long reinforcement laps in tension. Mag Concrete Research
2018; 70(14):739–55. https://doi.org/10.1680/jmacr.17.00285
[C11.19] Jones A, Holt D, Design of laps for deformed bars in concrete under bentonite and polymer drilling
fluids, the Structural Engineer, 2004, 82, No. 18, 32-38
[C11.20] Rehm G., Eligehausen R., Bond of ribbed bars under high cycle repeated loads, ACI Journal,
February 1979, pp. 297-309
[C11.21] Lemcherreq Y., Kaufmann W., Vogel Th., Fatigue of bond: Experimental investigation using pull-out
tests with distributed fibre optical sensors, Conference Paper https://doi.org/10.3929/ethz-b-
000465183, Institute of Structural Engineering, ETH Zurich, Switzerland, 2020-11
[C11.22] Hegger J., Camps B., Überprüfung der Bemessungsregeln für Übergreifungsstösse und
Verankerungen (gerade Stäbe, Abbiegungen und Haken) nach dem Entwurf der zweiten Generation
des Eurocode 2, Abschlussbericht, Institutsbericht Nr. 431/2020, IMB, RWTH Aachen, 22.12.2020.
See CEN-TC250-SC2-WG1_N1260_Mr Hegger PRB-PG2_0553
Hegger_Abschlussbericht_PRB_Uebergreifungsstoss-Teil_1_(202... and CEN-TC250-SC2-
WG1_N1258_Mr Hegger 2021-11-04_Presentation Professor Hegger_Anchorages and Laps_final
[C11.23] Camps B., Schmidt M., Hegger J., Anchorages and laps according to prEC2 – Review and
comparison to current EC2, fib Structural Concrete, publication under preparation.
[C11.24] CEN-TC250-SC2-WG1_N1284_Mr Ganz-Mr Muttoni - Latest technical changes - Clauses 11.4 & 11.5
- Joint WG1-SC2 meeting 2022-06-27-rev
12.1.2017, A.M. 2nd draft (after discussion with AhG Detailing, 10.1.2017)
13.1.2017, A.M. Average bond strength implemented
30.11.2017, A.M., New template, completed with new provisions in prEN 1992-1-1:2017-10
09.02.2021, HRG Text updated to prEN 1992-1-1:2020-11
12.04.2021, HRG Implemented comments and background information of Ch. Goodchild and J. Cairns, added fatigue, amended Figure C11.6,
corrected anchorage length according to EN 1992-1-1:2004, added comparison of lap length, prEN 1992-1-1:2020-11
15.04.2021, HRG Replaced hand-written examples with printed version, prEN 1992-1-1:2020-11
05.10.2022, HRG Updated to FprEN 1992-1-1:2022
25.01.2023, HRG Implemented comments from J. Cairns
02.03.2023, HRG Added references [C11.23] and [C11.24]
27.03.2023 Peer review
Page 1/500
For citations, copying and referencing see Note on Front page! page 439
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
11.4.7 Background to FprEN 1992-1-1:2023
Introduction This document describes the development of design formulae for the calculation of the anchorage resistance of
headed bars (clause 11.4.7) in FprEN 1992-1-1:2023.
Reasons for change New provisions on the anchorage of headed bars consistent with those for partially loaded areas have been
drafted in prEN 1992-1-1:2021, but several inconsistencies have been found with respect to the provisions of
EN 1992-4:
- the influence of compressive concrete strength on the anchorage strength is different (providing too
high resistance for high values of fck);
- the group effect (effect of spacing) is different;
- the effect of the distance to edges and corners is different.
For these reasons, a new proposal has been developed based on the same mechanical model which has been
used to draft the provisions for the minimum mandrel diameter ([C11.4.7.2], see also presentations and
discussions within CEN/TC250/SC2/WG1 in 2022). This choice is related to the fact that the two cases are
similar in terms of failure modes and parameters involved. This document describes the developed model as
well as the proposed provisions.
Similarities with the Figure C.11.4.7.1 shows failures of a headed bar near to the free surface and a similar failure of a bent
model developed for reinforcement due to the deviation forces. One can observe that in both cases, failure occurs due to spalling of
verifying the mandrel the concrete cover (so-called concrete blow-out failure).
diameter and
development of a
design formula for
headed bars
a) b)
Figure C11.4.7.1: (a) concrete spalling observed at the anchorage of headed bards (so-called blowout, tests by
Furche and Eligehausen, [C11.4.7.1]); (b) spalling failures due to an insufficient mandrel
diameter of bent bars near to a free edge (tests by Monney et al. [C11.4.7.2]).
This comparison shows that the mechanical model developed for the minimum mandrel diameter (see details in
[C11.4.7.2]) can be adapted to calculate the resistance of anchorages with headed bars. The main ideas and
assumptions are:
- The steel stress which can be activated in the anchored reinforcement bar without accounting for a
bond component corresponds to the triaxial concrete compressive strength fc,3 which can be activated
under the head multiplied by the ratio between the net area of the head and the cross-sectional area
of the reinforcement bar As :
2
A As h
kh, A h 1 (11.9)
As
where Ah is the total Area of the head;
- The concrete compressive stress under the head can exceed the uniaxial compressive concrete
strength fc (which should be reduced by the strength reduction factor cc, see clause 5.1.6(1)), due to
the confinement effect of the surrounding concrete. It can be calculated as [C. 11.4.7.3]:
fc ,3 cc fc 4 c ,1 (C11.4.7.1)
- The confinement concrete compressive stress c,1 in the confined area Ac,c parallel to the free surface
is in equilibrium with a confining tensile stress fct.,eff acting on the confining area Ac,t (see [C11.4.7.2])
and can be calculated as:
Act
c ,1 f ct , eff (C11.4.7.2)
Acc
- The confined area Ac,c under the head (see blue area in Figure C11.4.7.3c) is assumed to be
proportional to the net head area Ah – As;
- The confining area Act where the effective tensile strength fct.,eff can act (see red area in Figure
C11.4.7.3c) is assumed to be proportional to the distance ad between the bar axis and the free surface
times the nominal head diameter h;
- Similarly to [C11.4.7.2], the effective concrete tensile strength fct.,eff is assumed to be proportional to
the square root of the compressive concrete strength and is reduced with a size effect factor which
depends on the aggregate size parameter ddg (defined in 8.2.1(4)) and the nominal size √∙h with an
exponent of 1/3:
13
d dg
f ct , eff k1 fc (C11.4.7.3)
h
where k1 is a calibration factor.
Considering that the force in the bar is equal to the force that can be transmitted to concrete through the head
area before failure ( s
'
2 c ,1 Ah As ), the stress which can be activated in the reinforcement
4
becomes thus:
13
d dg a A As
kh , A cc f c 4 k1
'
fc k2 d h h
s
A h As
2
h
4 (C11.4.7.4)
56 13
a d h d dg
kh , A cc f c k3 f c
where k3 is the calibration factor replacing 16∙k1∙k2/ which can be determined, for the case of single bars near
to a free surface, on the basis of a comparison with laboratory tests ([C.11.4.7.1], [C.11.4.7.4], [C.11.4.7.5] and
[C.11.4.7.6]). As shown in the comparison of Figure C11.4.7.2 and Table C11.4.7.1, excellent results can be
obtained assuming k3=11,0. Only for the test series by Lee and Breen [C.11.4.7.6], a larger scatter can be
observed. This is due to the fact that for these tests, in some cases, the headed bars were anchored near to a
confinement reinforcement (spiral reinforcement in a pile) or adjacent to very dense longitudinal reinforcement
(both reinforcements not considered in the comparison).
Table C11.4.7.1: Statistical values of the ratio measured / calculated resistance (Rexp / Rcalc ) according to
Formula (C11.4.7.4).
2.00
1.80
1.60
1.40
Rexp / Rcalc
1.20
1.00
0.80
0.60
De Vries, 1996
0.40 Furche and Eligehausen, 1991
0.20 Hasselwander et al., 1977
Lee and Breen, 1966
0.00
1.00 1.50 2.00 2.50 3.00 3.50 4.00
a)
h /
2.00
1.80
1.60
1.40
Rexp / Rcalc
1.20
1.00
0.80
0.60
De Vries, 1996
0.40 Furche and Eligehausen, 1991
0.20 Hasselwander et al., 1977
Lee and Breen, 1966
0.00
0 10 20 30 40 50
b) fc [MPa]
Figure C11.4.7.2: Comparison of Formulae (C11.4.7.4) and (C11.4.7.5) with tests series in ([C.11.4.7.1],
[C.11.4.7.4], [C.11.4.7.5] and [C.11.4.7.6]); (a) influence of head size and (b) influence of
concrete strength.
As shown in Figure C11.4.7.2a, excellent results are obtained also for very small heads despite the fact that
very high concrete stresses are activated under the head (about 15fc in the tests with h=1.24by Furche and
Eligehausen [C.11.4.7.1]).
It has to be noted that formula (C11.4.7.4) covers the failure mode related to spalling of the concrete cover (also
called “blow-out” failure). For large covers, very high stresses can be reached and local failure can occur due to
compression of the matrix in the concrete and collapse of pores [C.11.4.7.7]. For this reason, the concrete stress
fc,3 has to be limited to the maximum value:
Based on the tests presented in [C.11.4.7.7], a value part of about 15 can be assumed for uncracked concrete
in the region of the head. Nevertheless, for the sake of simplicity (and despite the fact that the calibration factors
k3 of Formula (C11.4.7.4) and part of Formula (C11.4.7.5) refer to different failure modes and different
calibrations), the same value of part = 11.0 is assumed. In a similar manner as according to EN 1992-4:2018,
coefficient part shall be reduced for concrete cracked in the region of the head (part = 8.0 can be assumed for
this case).
Design value and The design value accounting for material, geometrical and model uncertainties can be determined based on the
design formula (11.8) approach of Annex A of FprEN 1992-1-1:2023 (see [C.11.4.7.8] for details). With respect to the first term in
based on reliability Formula (C11.4.7.4) linearly related to the uniaxial compressive concrete strength, the same partial safety factor
analysis as for uniaxial compression can be used (C = 1.5). For this reason, in the design formula, the expression cc∙fc
can be replaced by the design value fcd according to 5.1.6(1) which contains implicitly the sustained loading
effect. For the other term, which is associated to the concrete tensile strength, the uncertainties defined in Table
C11.4.7.2 below can be considered (see also Table A.2 of FprEN 1992-1-1:2023 and [C.11.4.7.8]). The
coefficient of variation and the bias factor of the resistance can be calculated with following formulae adapted
from Formulae (A.4) and (A.7) of FprEN 1992-1-1:2023:
2 2
V fc V ηis
VR V d2 V θ2 (C11.4.7.6)
2
2
1
f 2
R cm ηis d θ (C11.4.7.7)
fck
Table C11.4.7.2: Statistical values of the uncertainties assumed to calibrate the partial safety factor
Coefficient of
Uncertainty Bias factor
variation
The partial safety factor can then be calibrated according to Formula (A.1) of FprEN 1992-1-1:2023 with the
sensitivity factor R =0.8 and the target value for the 50-year reliability index tgt = 3.8:
Bars near to a corner As described above, formula (11.8) has been derived for the case of a single headed bar parallel to a free
or group of bars surface (position defined by the distance ad = ay between the bar axis and the free surface) and sufficiently
distant from a corner or another bar. For this case, the failure mode related to spalling (blow-out) is shown in
Figures C11.4.7.3a-b whereas the confined and the confining areas are shown in Figure C11.4.7.3c.
For the case of a bar too near to a corner (Figure C11.4.7.3d), the confining area can be limited by the presence
of the corner, reducing therefore the resistance. An additional detrimental effect is related to the fact that the
failure mechanism is characterized by a possible movement of the spalled concrete portion (red vector toward
the corner in Figure C11.4.7.3d) which has a further reduction of the confining zone as a consequence. To
account for this case, in FprEN 1992-1-1:2023, the nominal value of the distance ad is defined in a different
manner (see Table C11.4.7.3). In addition, for elongated head plates near to a corner, failure can occur first
under the head portion near to the corner. To account for this case, the distance ax shall be defined according
to Figure 11.9c of FprEN 1992-1-1:2023. It has to be noted that for circular and square heads, the formula in
Table C11.4.7.3 refers to the case where ay is the smallest distance to an edge (ax≥ay , otherwise, ax shall be
replaced by ay and vice versa). For the case with an elongated head, both combinations should be verified and
for the case of headed bars near to more than one edge, all edges should be verified separately.
A similar reduction applies also for a group of bars along an edge with a distance sx smaller than 4ay (see Figure
C11.4.7.3e). For this case, the nominal distance ad is determined with Formula (11.10) (see Table C11.4.7.3)
which has been derived assuming a linear transition between the two limits: ad= ay for the upper limit with sx=4∙ay
(no interaction between bars) and the lower limit for sx=x. For the latter case, a safe value of ad can be calculated
assuming that the heads are in contact and act in a similar manner as a strip load parallel to edge (for this case,
since no tests with a large number of headed bars have been found in the literature, the factor k3 in Formula
(C11.4.7.4) has been calibrated on the basis of tests with strip loads parallel to an edge found in the literature).
For the less detrimental case of a group with a limited number of bars, the resistance may be calculated
according to EN 1992-4 which provides a more favourable reduction factor for two and three bars.
Please note that Formula (11.10) in FprEN 1992-1-1:2023 contains an editorial error: coefficient 2,3 should read
0,23 as stated in Table C11.4.7.3 (using a value of 2,3, ad would become larger that ay which is physically not
possible).
Figure C11.4.7.3: Failure mechanisms, confined areas and confining areas for: (a-c) single bar; (d) corner bar
and (e) group of bars
Group of h sx 4 a y s x h 4a y s x 1
ad a y 0, 23 a y h 4a 1 (11.10)
bars along
an edge 4a y h 2 y h h
2
Headed bars with a In case of headed bars consisting of ribbed or indented reinforcement, a combination of the force carried by the
bonded length, head and the bond strength developed along the anchorage length may be accounted for according to 11.4.7(3).
11.4.7(3) In this case, the required anchorage length lbd to carry the remaining stress sd-’sd defined in Figure 11.9a may
be calculated according to Formula (11.11):
lbd 1,1 lbd sd lbd sd' (11.11)
The effect of adjacent In some practical cases, external compressive forces (e.g. from supports) or forces due to opposed active
plates and load headed bars (e.g. corner joint, see for instance Figure C11.4.7.4) can act in the zone of the head providing a
introduction points, beneficial confinement of the head and preventing blow-out failures. In the cases reported in [C.11.4.9] and
11.4.7(5) [C.11.4.7.10] for instance, a significant increase of the resistance has been attained. Since additional research
needs to be conducted to develop general provisions and detailing rules are still missing (position of the external
load with respect to the head, relative positions between heads, etc.), the provision according to 11.4.7(5) opens
the possibility to design these cases using a more refined model without giving detailed design rules.
Figure C11.4.7.4: Confinement effect by opposite plates in case of the reinforcement of a frame corner with a
closing moment anchored with heads (from Marchetto et al. [C. 11.4.7.10])
References [C. 11.4.7.1] Furche, J., Eligehausen, R., “Lateral Blow-out Failure of Headed Studs Near a Free Edge”;
Anchors in Concrete-Design and Behavior, ACI SP-130 81991), p. 235–252.
[C. 11.4.7.2] Monney, F.; Fernández Ruiz, M.; Muttoni, A., “Design against splitting failures in reinforced
concrete due to concentrated forces and minimum bend diameter of reinforcement”,
Engineering Structures, 245 (2021) , p. 21
[C. 11.4.7.3] Nielsen M.P, Hoang L.C., “Limit Analysis and Concrete Plasticity”, CRC Print, 2011, p. 816.
[C. 11.4.7.4] De Vries R.A., “Anchorage of Headed Reinforcement in Concrete”, Dissertation, The University
of Texas at Austin, 1996, p. 314
[C. 11.4.7.5] Hasselwander G.B., Jirsa J.O., Been J.E., Lo K., “Strength and Behavior of Anchor Bolts
embedded near edges of concrete piers”, Research report 29-2F, Center for Highway Reearch,
The University of Texas at Austin, 1977, p. 142
[C. 11.4.7.6] Lee D.W., Been J.E., “Factors Affecting Anchor Bolt Development”, Research report 88-1F,
Center for Highway Reearch, The University of Texas at Austin, 1966, p. 66
[C.11.4.7.7] Lieberum K.H., Reinhardt H.W., “Strength of Concrete on a extremely small bearing area”, ACI
Structural Journal, 1989, p. 67-75
[C.11.4.7.8] Muttoni A., “Background document to clause 4.3.3 and Annex A, Partial safety factors for
materials”, Background Document to FprEN 1992-1-1:2023, CEN/TC 250/SC 2, 23.02.2023, p.
21
[C. 11.4.7.9] Marchetto, F. Use of headed reinforcement bars in construction. A theoretical approach to
determine the dimensions of anchorage plates and experimental tests on knee joints subjected
to closing moment. PhD Thesis. Technical University of Madrid, 2015.
[C. 11.4.7.10] Marchetto, F., Pérez Caldentey, A. and Corres Peiretti, H. (2016), Structural performance of
corner joints subjected to a closing moment using mechanical anchorages: an experimental
study. Structural Concrete, 17: 987–1002. doi:10.1002/suco.201500222.
05.05.2021
For citations, copying and referencing see Note on Front page! page 446
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
11.5.4 Background to FprEN 1992-1-1: 2023
Introduction
This document describes the theoretical and experimental background of the simple equation proposed for the
next generation of EC2 [C11.2.1] for calculation of the capacity of U-bar loop connections subjected to
combinations of normal forces and bending moments. The proposal is based on an analytical upper bound
plasticity model [C11.2.4, C11.2.5, C11.2.6] which has been verified with a large number of available test
results.
U-bar loop connections (Figure C11.2.1) are frequently used to establish structural continuity between precast
concrete bridge decks. Here, U-bars protruding from each side of the deck elements to be connected are
overlapped in the connection, such that they confine a flattened cylindrical concrete core. After placement of the
precast decks, transverse reinforcement bars are placed within the overlapping U-bars and the connection is
subsequently grouted with concrete or mortar. In many applications, U-bar loop connections will often be
subjected to pure tension or tension in combination with a bending moment.
Figure C11.2.1: Illustration of a U-bar loop connection loaded in combined tension and bending; Plane
illustrations (a, b) and cross sectional illustration (c).
General
The model is based on the upper bound theorem of limit analysis (theory of plasticity). The original formulation
by Joergensen [C11.2.5] was developed following a two-step procedure. The first step involves a method to
design the connection in such a way that premature failure in the concrete at the joint is not governing. The
second step is necessary to ensure that the capacity related to failure in the concrete of the joint is not entirely
depending on the tensile strength of concrete. This eventually leads to the formulation of a procedure to
determine the minimum transverse reinforcement that is required to prevent brittle failure of the concrete. Based
on the analytical results obtained for the case of pure tension, it is possible to extent the model to cover the
case with combination of tension and bending moments.
2T
1
Arcsin (C11.2.1)
2
a
1
H
Here, = Arctan(a/H), where the parameters a and H are shown in Figure C11.2.2(a). Further, in Eq. (C11.2.1),
is the effectiveness factor (see Eq. (C11.2.4) below) andT is the mechanical ratio of transverse
reinforcement:
AsT f yT
T (C11.2.2)
Ac fc
where fyT and AsT are the yield strength and the total cross sectional area of the transverse reinforcement,
respectively. Additionally, fc refers to the uniaxial compressive strength of the joint concrete while Ac refers to the
cross sectional area of the concrete core confined by the overlapping U-bars. Further discussions of Ac will be
provided in the next sub-section.
With given by Eq. (C11.2.1), the optimal upper bound solution, Nc, appears as shown in Eq. (C11.2.3a), see
[C11.2.5]. It can be noted that the angle determined from Eq. (C11.2.1) is only a valid solution if it is not smaller
than as well as the internal angle of friction for concrete, = Arctan(3/4). Otherwise, must be taken as the
larger of and . This restriction has been discussed in details in [C11.2.5]. When inserting the two restrictions,
solution (b) and (c) in Eq. (C11.2.3) is derived. The subscript “c” indicates a capacity related to concrete failure:
2
4 T 1 T a a ;if and (a)
H H
2
2 T 3 4 a a 1
H H
Nc a 3
nL ; if and (b)
f c Ac 43
a H 4
H
a
2
a a 3
H 1 H ;if and (c)
H 4
(C11.2.3)
Here, nL is the number of U-bars available for tension transfer. The above mentioned effectiveness factor, ,
appears in Eq. (C11.2.3) as a reduction factor on fc. This is necessary in order to account for the fact that
concrete is not a perfectly plastic material [C11.2.8]. For the problem treated here, may be calculated as
follows, [C11.2.5]:
0.88 1
1 (f c in MPa , H in m) (C11.2.4)
fc H
According to Eq. (C11.2.1), = + /2 for T = 0. This means that Eq. (C11.2.3a) is valid when T = 0.
Consequently, the model predicts zero capacity when a loop connection does not contain transverse
reinforcement. This result is naturally conservative with respect to reality as, even without transverse bars, loop
connections are capable to carry some level of tension forces.
The tensile capacity of connections without transverse reinforcement is treated in section 2.3.
Figure C11.2.2: Failure mechanism in loop connection loaded in pure tension; yield lines (a) and rigid body
displacements (b).
Non-symmetric connections
U-bar loop connections may also have non-symmetric layout as shown in Fig. C11.2.1b, where the spacing
between overlapping U-bars is not constant but alternates between a1 and a2. It has been shown [C11.2.5] that
the capacity of non-symmetric connections can be determined on the basis of the mechanism illustrated in Fig.
C11.2.3, which is very similar to the one in Fig. C11.2.2. The analysis shows that the capacity of non-symmetric
connections is always larger than the capacity calculated on the basis of the solution for symmetric connections
(i.e. Eq. (C11.2.3)), provided that a (a1 a2 ) / 2 is inserted. Hence, use of Eq. (C11.2.3) with
a (a1 a2 ) / 2 will give conservative estimates for the capacity of non-symmetric connections.
Figure C11.2.3: Failure mechanism in loop connection loaded in pure tension; yield lines (a) and rigid body
displacements (b).
Figure C11.2.4: Failure mechanism in connection without transverse reinforcement; rigid body displacements
(a), relative displacement in yield lines (b) and concrete areas included in the model (c).
Based on the above discussion, it appears that a solution for connections without transverse reinforcement
should be developed by taking into account the tensile strength of concrete. As for the case of the cracking
moment of a cross section, the solution to be developed in this section may be regarded as a lower limit for the
tensile capacity of loop connections. This result can then be used to calculate the minimum transverse
reinforcement.
Figure C11.2.4 shows a failure mechanism that is similar to the one in Figure C11.2.2. This time, however, there
is no transverse reinforcement. Based on the same displacement field as assumed previously, the internally
dissipated energy and the external work may be calculated, see [C11.2.6]. However, in this calculation, the
separation yield lines which develop parallel to the n-axis should also be considered. The relative displacement
in the yield line between segments ② and ① is u2-1 = u② ‐ u① = (0, 2ut). For the yield line, which coincides
with the line of symmetry, the relative displacement, u3-3, is also (0, 2ut). This is justified because the two
segments ③ both move downwards by un but away from each other by ut. In this case, it is apparent that the
angle between the yield line and the relative displacement vector is /2. These yield lines are therefore called
separation yield lines, while the diagonal yield lines experience a combination of separation and sliding.
Lastly, there are the two yield lines running from the looped end of the outermost U-bars towards the nearest
boundary of the loop connection. However, there are no geometrical restrictions for the inclination of these two
yield lines. Consequently, the yield lines can always be oriented in such a way that they appear as separation
yield lines with a minimum of energy dissipation. The contribution to the load carrying capacity from these two
yield lines is a boundary effect, which can be neglected for loop connections in practice, where the number of
overlapping U-bars is high. Therefore, keeping in mind that the developed solution should be suitable for a
generalization (i.e. the result can be expressed as a capacity per U-bar) the effect of the two boundary yield
lines will not be taken into account in the following.
Now, on the basis of the assumed failure mechanism, the external work and the internal dissipated energy
(including the contribution from the separation yield lines) may be calculated, see [C11.2.6]. This leads to an
upper bound solution for the load carrying capacity. The optimal upper bound solution is found when the relative
displacement in each diagonal yield line forms the following angle with the yield line:
2nL 1 Ac* m m 1 (C11.2.5)
Arcsin l 1 l
L n Ac a
2
1
H
The parameters l and m are defined as follows:
t ft sin t ft 1
l 1 2 , m 1 2 (C11.2.6)
fc 1 sin fc 1 sin
Here, ft is the uniaxial tensile strength of concrete and t denotes the effectiveness factor for concrete stressed
in tension. The purpose of this factor is similar to the purpose of as mentioned in the above. The following
formulas, which also have been used in other plastic analyses (see [C11.2.8]), are adopted:
0.3
H
ft 0.26 fc2/3 , t 0.6 , (fc in MPa, H in meters) (C11.2.7)
0.1
The same restrictions as mentioned for the previous solution apply also to the angle determined from Eq.
(C11.2.5) (i.e. valid only if not smaller than as well as . With these restrictions and the angle in Eq.
(C11.2.5), the optimal solution for loop connections without transverse reinforcement may now be written as
follows (the subscript “ c,T = 0 ” indicates a tensile capacity related to concrete core failure when T = 0):
2 2
a 2nL 1 Ac* m m a m
l
H 1 1 ; if and a
nL Ac l l H l
2n 1 a m Ac * 2
a m
l L 3 4 1 1 5 3
N c , T 0 nL H l Ac
H
l
a 3
nL ; if and b
f c Ac 43
a H 4
H
2
a m a a 3
l 1 H l H ; if and c
H 4
(C11.2.8)
Where Ac/cos is the surface area of one diagonal yield line and Ac is the surface area of one separation yield
*
line. Strictly speaking, Ac* should be taken as ½(b H)t while Ac should be taken as H.t in order to make the
mechanism geometrically possible. However, as discussed in [C11.2.5] on the basis of the observed failure
pattern, it is reasonable for practical purposes to disregard the contribution of the cover. Therefore, as for the
previous solution (Eq. (C11.2.3)), the concrete cover will be neglected when determining the surface areas of
the yield lines. This means that Ac and Ac* are taken as the areas illustrated in Figure C11.2.4(c), i.e.:
D 2L H D 2L D 2L ,
2
Ac (a)
4
(C11.2.9)
1
Ac* b H D 2L , (b)
2
Here, the formula for Ac is only correct if H D 2L , where L is the cross sectional diameter of the U-bar.
This assumption is in accordance with all available experimental data (see [C11.2.5]). On the basis of the
empirical background, the solution in this document is therefore limited to connections with H D 2L .
Figure C11.2.5 shows the two solutions as a function of the transverse mechanical reinforcement ratio. From
Figure C11.2.5, it appears that the strength is governed by Eq. (C11.2.8) when the mechanical ratio of
transverse reinforcement is low. On the other hand Eq. (C11.2.3) is governing for large mechanical ratios of
transverse reinforcement. Hence, to ensure that the strength is not governed by the tensile strength of concrete,
the mechanical ratio of the transverse reinforcement must be larger than the ratio corresponding to the
intersection of the two curves. Figure C11.2.6 shows the mechanical ratio of the transverse reinforcement that
corresponds to the intersection point of the two curves. The calculations have been carried out by varying a
number of important design parameters.
Figure C11.2.5: Tensile capacity versus the effective mechanical ratio of transverse reinforcement. Parameters
*
used in example: a / H 0.5 , Ac / Ac 0.15 , m / l 0.80 , l 0.77 , nL 2
a
2
nL f c Ac 4 T T a 1
1
H ;if T (a)
H 2
Nc
2
a T
a 1
nL f c Ac H 1 ;if (b)
H 2
(C11.2.10)
Figure C11.2.7 shows the difference between Eq. (C11.2.3) and the simplified solution, Eq. (C11.2.10). The
difference between the two solutions comes from the contribution to the internal work of the transverse
reinforcement when the mechanical ratio of transverse reinforcement is high. It can be noted that, in practice,
the mechanical degree of transverse reinforcement is seldom so high that the difference between Eq. (C11.2.3)
and Eq. (C11.2.10) becomes significant.
Figure C11.2.7: Tensile capacity of loop connection versus effective mechanical ratio of transverse
reinforcement, T / .
It appears from the solution (and Figure C11.2.7) that the ratio a/H is very important for the tensile capacity of a
U-bar loop connection. The largest ratio in all tested specimens found in the literature is a/H = 100/170 ≈ 0.6
(see [C11.2.5]). It is therefore recommended to use only the solution in this document for connections with
a / H 0.6 .
Eq. (C11.2.10) can be reformulated into the following format:
2
a
a
N c nL f c Ac r
H H
(C11.2.11)
where r may be called the resistance factor of the transverse reinforcement and given by:
4 T T T 1
1 ;if 2 (a)
r
1 1
;if T (b)
2 (C11.2.12)
Based on the simplifications introduced, the complete solution for the tensile capacity, Nu, may be formulated as
follows:
N c (a)
N u min
N y nL AsL f yL (b)
(C11.2.13)
Where:
Nc is the tensile capacity when concrete failure is critical, see Eq. (C11.2.11)
(C11.2.14)
Where the effective tensile strength ftef = tft is determined by Eq. (C11.2.7). Fig. C11.2.8 shows the comparison
of the minimum reinforcement found by the iterative procedure and found by use of Eq. (C11.2.14) for a range
of parameters that are often used in practice.
Figure C11.2.8: Comparisons of approximate and accurate calculations of the minimum amount of transverse
reinforcement. Parameters used in calculation (when not varied): a 100 mm, H 170 mm, D 110 mm,
Figure C11.2.9: Loop connection subjected to combined tension and bending; sectional forces (a and b) and
sandwich model for distribution of internal forces (c).
Figures C11.2.9(a) and C11.2.9(b) illustrate a loop connection subjected to a tension force, N, and a bending
moment, M (the figure shows a connection with nL = 2). The set of sectional forces, (N, M), is assumed to be
statically equivalent with the distribution of internal forces in the sandwich model shown in Figure C11.2.9(c).
The sandwich model consists of two layers. The first layer has thickness x and carries a uniaxial compression
stress field with stress magnitude fc. The stress resultant of this stress field is denoted C = fcLx in the figure. The
second layer embeds the U-bars and has the thickness (D+2L). This layer transfers a tensile force, F, which, as
shown in the figure, is assumed to be:
F C Nu , [0;1] (C11.2.15)
where Nu is the pure tensile strength of the connection as determined by Eq. (C11.2.13). Based on the results
for the case of pure tension, it is possible to state that F can be transferred if F ≤ Nu. Remember that in the pure
tension model, the concrete cover was neglected. The cover can therefore here be used as the layer that
carries the uniaxial compression field. Keeping the picture of a sandwich model in mind, it is clear that the
thickness x of the compression layer cannot be larger than the cover c. Hence, there is an upper limit for the
compression force, namely C ≤ fcLc. The two mentioned conditions, i.e. C ≤ fcLc and F ≤ Nu, can be condensed
into the following condition, when F is replaced by the right hand side of Eq. (C11.2.15):
N u (1- )
C Cmax min (C11.2.16)
f c Lc
Now, by requiring static equivalence between (N, M) and (C, F), the following is found:
N N u , 0;1 (a)
(C11.2.17)
tx
M Cz ,z (b )
2
From Eq. (C11.2.17b) it is seen that maximum bending moment may be obtained when Cmax (see Eq.
(C.11.2.16)) is inserted. To carry Cmax, the compression layer must have a thickness, x, which can be
determined as follows:
N
N u (1- N )
u
x min (C11.2.18)
f c L
c
Cmax and x according to Eq. (C11.2.18) can now be inserted into Eq. (C11.2.17b) to calculate the maximum
bending moment as a function of the tension force, i.e. M(N). The result can be expressed as follows:
N
Nu 1
1 Nu 1 N t N u
(a)
M (N ) 2 f Lc N c f Lc
min c u
c
(C11.2.19)
f c Lc 2
1 t 1 (b )
2 c
Equation (C11.2.19) describes a N-M interaction diagram for U-bar loop connections. The solution is general
and can be used for connections with any number of overlapping U-bars. As a special case, Eq. (C11.2.19) also
provides a solution for the bending capacity of connections loaded in pure bending (when N = 0 is inserted). The
pure bending capacity will in the following be referred to as Mu, i.e. Mu = M(N =0).
Figure C.11.2.10 shows two examples of the calculated N-M interaction diagram. The horizontal cut-off, i.e. Eq.
(C11.2.19b), is only relevant in a small part of the diagram when the concrete core is made sufficiently strong by
increasing the ratio of transverse reinforcement, T. This is illustrated by the diagram with T/= 0.5. For low
values of T and in cases where Nu is governed by yielding of the U-bars (i.e. Eq. (C11.2.13b)), where the U-bar
diameter and spacing correspond a lightly/normally reinforced cross section, the N-M interaction diagram is
entirely described by Eq. (C11.2.19a). In such cases, the N-M diagram is almost linear as shown for T/= 0.2.
This is similar to the N-M diagram for a symmetric reinforced beam/slab cross section subjected tension and
bending.
Figure C11.2.10: N-M interaction diagram. Parameters used in calc.: a/H = 0.50, A*c /Ac = 0.15, nL =2, = 0.55,
fc =27MPa, Ac = 26000 mm2, L = 600 mm, c = 25 mm.
D c
M u 2Tu L Tu D 2L c (C11.2.20)
2 2
Now, suppose that the sandwich model is not available and suppose that the only information available is the
capacity of each the U-bar legs, Tu. In that case, a cross sectional analysis may be performed to estimate
bending moment that can be transferred from one side of the connection to the other side. As well known, it is
normal practice to neglect reinforcement bars near the compression zone when calculating the bending capacity
of a cross section. Hence, the internal forces Tu and C shown in Figure C11.2.11c would form the basis of
calculation. Since C = Tu in this case, the compression zone is reduced to x = ½c. Again, by requirering static
equivalent, the sectional analysis provide the following bending capacity:
3 3
M u Tu D L c (C11.2.21)
2 4
It can be seen by comparing Eqs. (C11.2.20) and (C11.2.21) that the bending capacity obtained from the
principle of a cross sectional analysis is smaller than the result of the sandwich model. For typical values of D,
L and c, the results are within app. 10%. As an example where D = 125 mm, L = 16 mm and c = 35 mm, Eq.
(C11.2.21) gives 9% smaller capacity than Eq. (C11.2.20).
The above result shows, that in practice, a sectional analysis can replace the sandwich model in cases with
combined tension and bending. And as the tension force becomes dominant, the results of the two approaches
will converge and be identical for the case of pure tension.
(a)
(b)
(c)
Figure C11.2.11: Internal forces (in U-bars) equivalent to pure tension (a); and internal forces equivalent to pure
bending according to the sandwich model (b) and to a sectional analysis (c)
The average value of N u ,test / N u ,calc for all tests is 1.01 with a standard deviation of 0.13.
Pure Tension Tests that satisfy the requirement of minimum transverse reinforcement
Tests used in comparison (Fig. C11.2.12):
- Hao (2004 [C11.2.3]): # of tests: 4*
- Gordon (2006 [C11.2.2]): # of tests: 8*
- Ma et al (2012 [C11.2.7]): # of tests: 4
- Joergensen (2014 [C11.2.5]): # of tests: 28
* Where the transverse reinforcement was not anchored at the ends, the capacity of the transverse
reinforcement is reduced according to a linear development curve calculated on the basis of the Eurocode 2
anchorage length; see detailed calculation in [C11.2.5].
The average value of N u ,test / N u ,calc is 1.02 with a standard deviation of 0.12.
Figure C11.2.12: Ratio N u ,test / N u ,calc as a function of different parameters varied in the tests for connections
with min
Pure Tension Tests that do not satisfy the requirement of minimum transverse reinforcement
Tests used in comparison (Fig. C11.2.13):
- Hao (2004 [C11.2.3]): # of tests: 11*
- Joergensen (2014 [C11.2.5]): # of tests: 10
The average value of N u ,test / N u ,calc is 1.01 with a standard deviation of 0.15.
Figure C11.2.13: Ratio N u ,test / N u ,calc as a function of different parameters varied in the tests for connections
with min
Pure tension tests versus theory using alternative formulation of the effectiveness factor
To keep a consistent formulation of the effectiveness factor in the entire FprEN1992-1-1:2023, the -formula as
stated in Eq. (C11.2.4) will be replaced by the formula below when implemented in FprEN1992-1-1:2023:
EC 2 fc k3
(C11.2.22)
where:
1
(C11.2.23)
fc min 40 MPa 1/3
fc
1/3
d dg
k3 0.85
H (C11.2.24)
It should be noted that fc accounts for the brittleness of concrete in a similar way as the factor does.
Likewise, k3 as well as the factor in Eq. (C11.2.4) account for the size effect. The parameter ddg is
defined in FprEN1992-1-1:2023.
Figure C11.2.14 shows plots of the ratio between tested and calculated capacity based on the modified
effectiveness factor (Eq. C11.2.22). The average value of N u ,test / N u ,calc is 1.05 and the standard deviation of
0.12. The plots can be directly compared to those in Figure C11.2.12, where Eq. (C11.2.4) was used to
calculate . It can be seen that the results obtained with these modifications are very similar.
Figure C11.2.14: Ratio N u ,test / N u ,calc as a function of different parameters varied in the tests for connections
with min
Figure C11.2.15: Comparison of tests with N-M diagram determined by the sandwich model (tests taken from
ref. [C11.2.5])
Proposal in FprEN1992-1-1:2023
The simplified Equation (C11.2.11) is the basis for subsection 11.5.4 (Laps using U-bar loops) of FprEN1992-1-
1:2023.
In addition to change of notation for consistency with FprEN1992-1-1:2023 (which will not be introduced or
explained below) some minor changes have been introduced to make the provision more general (i.e. covering
both pure tension and tension combined with bending) and more consistent with the rest of the code revisions.
These changes are explained below.
2
a
1 a
N c f c Ac r
2 H H
(C11.2.25)
Figure C11.2.15: Number of yield lines formed in U-bar connection with nL = 1 (a) and in a unit of U-bar lap (b).
2
a
1 a
Tc f c Ac r (C11.2.26)
4 H H
In FprEN1992-1-1:2023, the design compression strength is defined as fcd = fcfck/c. This means that a part of
the -formula in Eq. (C.11.2.22) will be embedded in the definition of fcd. Therefore, when introducing
(C.11.2.22) and fcd into (C.11.2.26) and into the definition of r, the following expression for the design strength of
each U-bar leg in the lap is obtained (except for the difference in notations, the expression has the same
meaning as the Equation of clause 11.5.4(2) in FprEN 1992-1-1:2023):
d dg a
1/3 2
a
Tcd 0.2 f cd Ac r (C11.2.26)
H H H
where
4 T T 1
1 ;if T (a)
2
r
1 1
;if T (b)
2 (C11.2.27)
and
0.85 Ac f cd
H
By use of Eq. (C11.2.26), crushing of the concrete within the U bar lap can be prevented when:
where T1 and T2 are the tension forces in the U-bar legs calculated from a sectional analysis when the lap
transfers combinations of normal forces and bending moments, see Figure C11.2.16. It should though here be
noted that according to the sandwich model, both legs will carry the same tension force and therefore, when
designing U-bar laps using Eq. (C11.2.26), both legs of the U bar must be anchored outside the connection for
the larger of the forces T1 and T2.
Figure C11.2.16: Tension forces in U-bar legs obtained from sectional analysis when lap transfers
combinations of normal forces and bending moments
min f yT 0.5 f c
(C11.2.31)
Equation (C11.2.31) is the one that is adopted in FprEN1992-1-1:2023. Comparison of Equations (C11.2.30)
and (C11.2.31) can be seen in Figure C11.2.17.
Figure C11.2.17: Comparison of Eqs. (C11.2.30) and (C11.2.31) for estimation of minimum transverse
reinforcement.
References
[C11.2.1] CEN European Committee for Standardization, Eurocode 2. Design of concrete structures – general
rules and rules for buildings, EN 1992-1-1, Brussels, Belgium, 2004. 225p.
[C11.2.2] Gordon, S. R., Joints for Precast Decks in Steel Concrete Composite Bridges. PhD thesis, School of
the Built Environment, Heriot Watt University, Edinburgh, UK, 2006
[C11.2.3] Hao, J. B., Structural Behaviour of Precast Component Joints with Loop Connection. PhD thesis,
Department of Civil Engineering, National University of Singapore, Singapore, 2004
[C11.2.4] Joergensen H. B., Hoang L. C., Tests and limit analysis of loop connections between precast
concrete elements loaded in tension, Engineering Structures, Vol. 52, pp. 558-569, 2013
[C11.2.5] Joergensen, H. B. Strength of Loop Connections between Precast Concrete Elements – Part I: U-bar
connections loaded in combined tension and bending, Part II: Wire loop connections loaded in shear.
PhD thesis, Department of Technology and Innovation, University of Southern Denmark, Odense,
Denmark, 2014.
[C11.2.6] Joergensen H. B., Hoang L. C., Strength of Loop Connections between Precast Bridge Decks
Loaded in Combined Tension and Bending, Structural Engineering International, Vol. 25, No 1, pp.
71-80, 2015
[C11.2.7] Ma, Z. J., Lewis, S., Cao, Q., He, Z., Burdette, E. G., and French, C. E. W., Transverse Joint Details
with Tight Bend Diameter U-bars for Accelerated Bridge Construction. Journal of Structural
Engineering, 138(6): 697-707, 2012
[C11.2.8] Nielsen M. P., Hoang L. C., Limit analysis and concrete plasticity. CRC Press, third edition, 2011
05.05.2021
For citations, copying and referencing see Note on Front page! page 464
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
11.5.5 Background to FprEN 1992-1-1: 2023
Introduction
This document describes the background to the equation proposed for calculation of the tension capacity of short
headed bar laps in the next generation of EC2 [1]. The proposal is based on the analytical upper bound plasticity
model of Joergensen and Hoang [2-4] used to calculate the tensile strength of laps using U-bars in FprEN 1992-
1-1: 2023 [1].
Short headed bar laps can for example be used to establish structural continuity between precast concrete slabs.
Headed bars protruding from each side of the precast elements to be connected are overlapped in the connection
as shown in Figure 1. After placement of the precast units, transverse reinforcement bars are placed over the
overlapping headed bars and the connection is subsequently infilled with concrete. The strength and ductility of
headed bar tension laps is increased by the provision of through thickness tie down reinforcement which provides
confinement. The design recommendations relate to headed bar laps loaded in either pure tension or flexure
induced tension.
General
The design approach for headed bar tension laps is based on the upper bound plasticity model proposed by
Joergensen and Hoang [2-4] for U-bar loop splices. This approach is described in detail in the background
document to C11.5.4 of FprEN 1992-1-1:2023 [5] for laps using U-bar loops. This document summarises the key
assumptions implicit in the design method pertinent to headed bar tension laps as well as the validation of the
method with experimental data. A change to the model of Joergensen and Hoang [2-4] for laps of both U-bar loops
and headed bars is that plane stress is considered instead of plane strain. This gives lower and more realistic
strength predictions for tested headed bar joints and is justified by the observation that the anchor plates are
suggestive of a strut and tie system in plane stress. The member depth is assumed to equal to the effective head
depth. Concrete and reinforcement are assumed to behave in a rigid plastic fashion and dowel action is neglected.
Proposal in FprEN1992-1-1:2023
The model applies to short laps of headed bars with head size large enough to anchor the bar force without
additional contribution from bond along the bar. Heads complying with 11.4.7 (1) of FprEN 1992-1-1:2023 [1]
satisfy this requirement for fyd ≤ 435 MPa. The maximum tensile force developed in each bar should be limited by
11.4.7(2).
Overlapping headed bars may be single (Figure 2a)) or multiple (Figure 2b)) and should be anchored for the
design force outside the connection.
The proposed design equation for the resistance of headed bar laps is derived from Eq. (C11.2.25) of the
background document for laps using U-bar loops [5], which is formulated in a manner that suits the problem of a
standard unit of U-bar lap, where only one yield line can develop as shown for a unit of headed bar lap in Figure
2a). The Equation is repeated here:
2
a
1 a
N c f c Ac r
2 H H
(1)
For headed bar laps, the dimensions a and H are defined as shown in Figure 2a.
Figure 2: Number of yield lines formed in (a) a unit of headed bar lap and (b) with ns= 3 headed bars.
EC 2 fc k3
(2)
where:
1
(3)
fc min 40 MPa 1/3
f c
and
1/3
d dg
k3 1.3
lsd 2 (4)
in which H in Eq. (1) is replaced by lsd 2 , where lsd is the lap length measured to the outside of the bar heads
and is the headed bar diameter which is assumed to equal the head thickness.
In the design expression, the term fc is absorbed into fcd which is defined in FprEN 1992-1-1:2023 as
f ck (5)
f cd fc
c
For the assumed range of model applicability 0 ≤ cs ≤ 0.5lsd, the resistance of a single headed bar lap TRd,c may
be calculated as:
/
𝑇 , 0,6𝑓 𝐴 𝑘 (6)
∅ ∅ ∅
where
Ac is the effective concrete area within the heads of the over lapping bars (Fig. 3a)):
Ac = (lsd-2)bh1 (7)
lsd is the overlapping length measured to the outside of the heads which are assumed to be of thickness where
is the headed bar diameter (Fig. 3a));
ddg is a coefficient that takes into account the concrete type and its aggregate properties according to 8.2.1(4) [1];
bh1 is effective width of the head perpendicular to the plane of the lap (for circular head with diameter 𝜙 :
𝑏 0.5𝜙 √π) (8)
cs is the clear spacing of headed bars (see Fig, 3b));
kst is the resistance factor of the transverse reinforcement, which may be taken equal to kst = 1 when ≥ 0,5 and
to kst = 4(1 – ) for lower values of 𝜔 / (9)
. ∅
Ast is the total area of the fully anchored transverse reinforcement positioned within Ac.
a) elevation view
b) plan view of single lap
c) plan view of multiple lap
In case of multiple headed bars as shown in Figure 3c), the resistance of Eq. (6) may be calculated using an
average net spacing cs = 0,5(cs1 + cs2) (see Figure 3c)) and multiplied by (ns – 1) where ns is the total number of
headed bars (ns = 5 in the example of Figure 3c)). Multiple overlaps may also be treated as an assembly of single
overlaps where cs is taken as the smaller of cs1 and cs2.
where
is the maximum diameter of the lapped headed bars;
To enhance ductility of laps designed for reinforcement yield, tie down reinforcement with total area Astd per single
lap should be provided perpendicular to the plane of the headed bars within Ac (Figure 3a). The tie down
reinforcement should be fully anchored outside Ac and placed symmetrically between the headed bars. The tie
down reinforcement may be provided in the form of either double headed shear studs or links. For lsd ≤ 200 mm,
Astd may be provided by a single double headed shear stud (Figure 3c). The minimum area of tie down
reinforcement per single lap equals:
𝐴std 0,12𝜙 (11)
obtained by multiplying the coefficient of 0.5 in Eq. C11.2.31 of [5] for U-bar loops by 0.6/(2×0.2) = 1.5. The
multiple of 0.6/(2×0.2) results from the increased concrete efficiency factor for headed bars in Eq. (6) compared
to Eq. C11.2.26 of [5] for U-bar loops for which gives the resistance of a single leg of the U-bar. Eq. (10) introduces
an empirical lower limit to the minimum area of transverse reinforcement of 𝐴st 𝜋𝜙 /8 since the theoretical
minimum area of reinforcement is significantly less than that provided in any of the tests within the experimental
database used to calibrate the model. Consequently, based on the minimum area of transverse reinforcement
provided in the tension tests of Vella et al. [6], the minimum area of transverse reinforcement is increased to 𝐴st
𝑚𝑎𝑥𝑖𝑚𝑢𝑚 0,75 𝑓ck ⋅ , 2/8), where is the headed bar diameter.
yk
can in some cases be significantly underestimated as shown in Figure 6 since the method is calibrated for
specimens having the TT’ transverse reinforcement arrangement and no through thickness shear studs.
For the 8 specimens in group 2 with no shear studs and transverse bars in the TT’ arrangement the mean and
standard deviation of Ttest/Tcalc are 1.19 and 0.10 respectively. For the 15 specimens with shear studs and
transverse bars in the TT’ arrangement the mean and standard deviation of Ttest/Tcalc are 1.43 and 0.11
respectively.
Figure 5: Typical flexural test specimen and section through specimen B2-26-2H20-S-10
Ptest
Transverse
sh Hhb (kN)
Test ID fc,cyl (MPa) ft,split (MPa) ∅tr (mm) bar positions
(mm) (mm) {failure
(see Fig. 4)
mode}
Note: J depicts joint failure, Y depicts headed bar yield, S depicts strut failure, T depicts tie failure.
ft,split spit cylinder tensile strength; ∅tr diameter of transverse reinforcement bars; H=lsd-2 (lap between inside face
of heads)
Table 2: Comparison of flexural resistance obtained in flexural tests of Vella et al. [6] and flexural resistances
calculated with tensile resistance obtained in corresponding tension test
Specimen M
Test ID
type (kNm)
2,5
2
Ttest/Tcalc
1,5
1
Studs TT' No studs TT'
0,5
StudsT'B' Studs TT'BB'
0
15 20 25 30 35 40 45 50 55 60
fc MPa
b) Influence of /min
3
2,5
2
Ptest/Pcalc
1,5
/min
c) Influence of
2,5
Ptest/Pcalc
1,5
1
Studs TT' No studs TT'
0,5
Studs T'B' Studs TT'BB'
0
0 0,5 1 1,5 2 2,5 3 3,5 4
1,8
1,6
1,4
1,2
Ttest/Tcalc
1
0,8
0,6
0,4
G5 studs
0,2
0
0 25 50 75 100 125 150 175 200 225 250
H mm
e) Influence of cs/H
1,8
1,6
1,4
1,2
Ttest/Tcalc
1
0,8
0,6
0,4
Studs TT'
0,2
0
0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,9 1
cs/H
Figure 6: Influence on Ttest/Tcalc of a) concrete strength, b) min, c) d) lap length H and d) cs/H
1,6
1,4
1,2
1
Ttest/Tcalc
0,8
0,6
0,4
0,2
0
0 20 40 60 80 100 120 140 160
1,6
1,4
1,2
1
Ttest/Tcalc
0,8
0,6
0,4
0,2
0
0,0 0,1 0,2 0,3 0,4 0,5 0,6
cs/(H =lsd-2 )
b)
Figure 7: Influence on Ttest/Tcalc for specimens of Li et al. [8] of a) H =lsd-2 and b) cs/H
In the absence of transverse reinforcement, the resistance was calculated assuming that the area of transverse
reinforcement equalled 𝐴 0.75 𝑓 since this is the theoretical area at which Eq. 6 gives the same
resistance as calculated in a yield line analysis without transverse reinforcement (see background document for
lapped U-bar loops [5]). The ratio Ttest/Tcalc is plotted against cs/H, kst and the tie down ratio respectively in
Figures 9a to 9c. Interestingly, the proposed design method gives safe results for all the test specimens of
Thomson despite the head size being insufficient to develop the full bar strength. Figure 9b shows that the degree
of conservatism of the design method reduces with the coefficient kst in Eq. 6 which depends on (Eq. 9). It should
be noted that kst was calculated with the theoretical minimum area of transverse reinforcement, which corresponds
to concrete tensile failure, since no transverse reinforcement was provided in these specimens. Figure 9c shows
that the ratio Ttest/Tcalc is almost independent of the tie down ratio. However, the provision of confining tie down
reinforcement increased the deformation at failure in the Thompson tests [9,10] with the transverse tie down detail
in Figure 8 being most effective. In this detail, the area of transverse reinforcement was 28% of the headed bar
cross sectional area and the tie down ratio was 0.32. These ratios compare with a minimum ratio of transverse
bar area to headed bar of 0.46 in the tests of Vella et al. [6,7] and a tie down ratio provided by shear studs of 0.11
per lapped bar (2×78.5/(3×491) = 0.11) or 0.16 per single lap. The presence of the shear studs increased lap
strength in the tests of Vella et al. as well as the deformation at failure in tests where headed bars yielded. The
increase in strength due to the confining studs is given by the ratio of strengths of corresponding specimens in
Group 1 (with studs) and Group 2 (without studs). This ratio varies between 1.12 and 1.28 with an average of
1.18.
The test specimens of Vella et al. [6,7] with only transverse reinforcement exhibited more ductility than those of
Thompson without tie down reinforcement. Additionally, a ductile response was also achieved in the tests of Li et
al. [8] with 152 mm laps despite the specimens having no tie down reinforcement. Consequently, it is proposed
that tie down reinforcement is only necessary when laps are designed for bar yield.
Figure 9d shows Ttest/Tcalc plotted against the minimum area of transverse reinforcement calculated as 𝐴
0.75 𝑓 . The resulting areas of reinforcement are seen to be very small compared with the cross-sectional
area of the 1’’ diameter lapped headed bars. Hence, for practical reasons a minimum area of transverse
reinforcement equal to half the headed bar cross sectional area is proposed. However, Figure 10 shows that the
measured strengths increase with kt in a similar fashion to the calculated strengths with 𝐴 0.75 𝑓 . The
tests of Vella et al. and Thompson suggest that added ductility can be provided with through thickness confining
reinforcement giving a tie down ratio of 0.15 per single lap or more. This leads to Eq. 11.19 for tie down
reinforcement in which the coefficient of 0.12 is given by 0.15 0.25𝜋𝜙 0.12𝜙 .
3,5
2,5
Ttest/Tcalc
1,5
0,5
0
0 0,05 0,1 0,15 0,2 0,25 0,3 0,35 0,4 0,45
cs/H
a)
3,5
2,5
Ttest/Tcalc
1,5
0,5
0
0,5 0,6 0,7 0,8 0,9 1
kst
b)
4
3,5
3
2,5
Ttest/Tcalc
2
1,5
1
0,5
0
0 0,2 0,4 0,6 0,8 1 1,2
Tie down ratio
c)
4
3,5
3
2,5
Ttest/Tcalc
2 provided
1,5
1
0,5
0
0 20 40 60 80 100 120 140 160
Astmin mm2
d)
Figure 9: Influence on Ttest/Tcalc for specimens of Thompson [9] a) cs/H, b) kst, c) tie down ratio and d) 𝐴
𝑚𝑎𝑥𝑖𝑚𝑢𝑚 0.75 𝑓 𝑡ℎ𝑒𝑜𝑟𝑒𝑡𝑖𝑐𝑎𝑙 ; 𝑝𝑟𝑜𝑣𝑖𝑑𝑒𝑑 𝑠𝑡𝑒𝑒𝑙 )
250
200
T per bar kN
150
100
50 Calculated
Test
0
0,5 0,6 0,7 0,8 0,9 1
kst
Figure 10: Influence of kst (calculated with 𝐴 0.75 𝑓 ) on strength of Thompson specimens [9]
Parametric study
The practical implications of the proposed design rules are illustrated in Figure 11. The headed bar arrangement
was assumed to consist of bars spaced at 200 mm centres on each side of the lap. The lap strengths were
calculated assuming 2 single laps per headed bar with 25 mm cover to the bar head which was assumed to be
square with area equal to 10 times the headed bar cross section as in the tests of Vella et al. [6,7]. Partial factors
of 1.5 and 1.15 were assumed for concrete and steel respectively with reinforcement yield strength equal to 500
MPa. The maximum required areas of transverse reinforcement required for the full strength laps shown in
Figure 11 were 145 mm2 for H16 bars, 209 mm2 for H20 bars and 293 mm2 for H25 bars. The cover to the head
was chosen to ensure that attainment of bar yield was not prevented by side blowout according to 11.4.7(2) in [4].
A single 10 mm diameter double headed stud positioned between each headed bar, and midway along the lap,
would be sufficient to satisfy the tie down requirements of Eq. (11).
0,8
TRd,c/Tyd
0,6
0,4
0,2
0
100 125 150 175 200 225
Lap length H mm
0,8
TRd,c/Tyd
0,6
0,4
0,2
0
30 40 50 60 70 80
Concrete strength fck MPa
b)
Figure 11: Influence of a) lap length and b) concrete strength on design lap strength according to proposed design
method
References
[1] FprEN 1992-1-1: 2023 Eurocode 2: Design of concrete structures– Part 1-1: General rules – Rules
for buildings, bridges and civil engineering structures
[2] Joergensen H. B., Hoang L. C., Tests and limit analysis of loop connections between precast
concrete elements loaded in tension, Engineering Structures, Vol. 52, pp. 558-569, 2013
[3] Joergensen, H. B. Strength of Loop Connections between Precast Concrete Elements – Part I: U-
bar connections loaded in combined tension and bending, Part II: Wire loop connections loaded in
shear. PhD thesis, Department of Technology and Innovation, University of Southern Denmark,
Odense, Denmark, 2014.
[4] Joergensen H. B., Hoang L. C., Strength of Loop Connections between Precast Bridge Decks
Loaded in Combined Tension and Bending, Structural Engineering International, Vol. 25, No 1, pp.
71-80, 2015
[5] Joergensen H. B., Hoang L. C., Background document to subsection 11.5.4: Laps using U-bar loops.
CEN/TC250/SC2/WG1/TG4, May 2021.
[6] Vella J. P., Vollum R. L., Jackson A., Investigation of headed bar joints between precast concrete
panels, Engineering Structures, Vol: 138, pp. 351-366, 2017.
[7] Vella J. P., Vollum R.L., Jackson A., Flexural Behaviour of Headed Bar Connections between Precast
Concrete Panels, Construction and Building Materials, 154, pp. 236-250, 2017.
[8] Li, L., Ma, Z., Griffey, M. E., and Oesterle, R. G., Improved Longitudinal Joint Details in Decked Bulb
Tees for Accelerated Bridge Construction: Concept Development. Journal of Bridge Engineering,
15(3): pp. 327-336, 2010.
[9] Thompson M.K., The Anchorage Behavior of Headed Reinforcement in CCT Nodes and Lap Splices,
Phd thesis, The University of Texas at Austin, 2002.
[10] Thompson MK, Ledesma A, Jirsa J, Breen J, Lap splices anchored by headed bars, ACI Structural
Journal, 103 (2): p. 271-279, 2006.
For citations, copying and referencing see Note on Front page! page 479
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
11.6.3 Background to FprEN 1992-1-1:2022
Introduction
This document has been prepared in order to provide background for the provisions in Clause 11.6.3 on minimum
radius of curvature of tendons. The provisions are mainly based on common practice in post-tensioning industry
and provisions in the European Assessment Document EAD 160004 for prestressing systems, [C11.6-1].
pR,max is the recommended maximum allowable pressure which in the absence of national regulations may be
taken as either 130, 150 or 230 kN/m
In the prestressing industry the following simple formula has been used since long for the minimum radius of
tendon curvature both for strands and wires which is also given in fib Model Code, [C11.6-2]:
Rmin = 2,8 (fpk Ap) 2,5m (C11.6-3)
where:
fpk is the specified tensile strength of the prestressing steel; (fpk Ap) expressed in [MN] for Rmin [m]
Ap is the cross sectional area of prestressing steel
Figure C11.6-1 compares the minimum radii of tendon curvature according to Formulae (C11.6-1), (C11.6-2) and
(C11.6-3) as a function of Ap. For the comparison, prestressing steel strand with 15,7mm nominal diameter,
150mm2 cross-section, and 1860 MPa specified tensile strength is considered. It is further assumed that the initial
stress is 0,80 fpk and that Ap represents about 40% of the duct cross section, as typically found in practice. As it
appears from Figure C11.6-1, minimum radii of tendon curvature as permitted in [C11.6-1] and [C11.6-2] are
obtained if the upper limit for the maximum transverse bearing stress in FprEN 1992-1-1:2022, Table 11.4 (NDP),
is set to pRd = 15MPa.
(a) Minimum radii for fck = 20MPa, pRd 12MPa (b) Minimum radii for fck = 30MPa, pRd 12MPa
Figure C11.6-2 compares the minimum radii of tendon curvature according to Formulae (C11.6-1) and (C11.6-4)
with the values in Table C11.6-1 as a function of Ap. For the comparison, prestressing steel strand with 15,7mm
nominal diameter, 150mm2 cross-section, and 1860 MPa specified tensile strength is considered. It is further
assumed that the initial stress is 0,80 fpk. As it appears from Figure C11.6-2, minimum radii of tendon curvature
as permitted in [C11.6-1] and [C11.6-2] are obtained if the upper limit for the maximum transverse bearing stress
in FprEN 1992-1-1:2022, Table 11.4 (NDP), is set to pRd = 30MPa.
(a) Minimum radii for pRd = 24MPa (b) Minimum radii for pRd = 30MPa
Figure C11.6-2: Comparison of minimum radii of curvature for external tendons
Figure C11.6-3 compares the minimum radii of tendon curvature according to Formulae (C11.6-1) and (C11.6-5)
as a function of Ap. For the comparison, prestressing steel strand with 15,7mm nominal diameter, 150mm2 cross-
section, and 1860 MPa specified tensile strength is considered. It is further assumed that the initial stress is 0,80
fpk. As it appears from Figure C11.6-3, minimum radii of tendon curvature as permitted in [C11.6-2] are obtained
if the upper limit for the maximum transverse bearing stress is set to pRd = 70MPa.
(a) Minimum radii for pRd = 55MPa (b) Minimum radii for pRd = 70MPa
Figure C11.6-3: Comparison of minimum radii of curvature for loop in internal tendons
It should be noted that the compressive stress on the concrete inside the loop curvature needs to be checked.
For the parameters of Figure C11.6-3(a) a design concrete strength of fcd = 30MPa, and for C11.6-3(b) of fcd =
40MPa is required. Hence, a sufficiently high concrete strength should be specified or the concrete should be
confined and verified to FprEN 1992-1-1:2022, Clause 8.1.4, or alternatively, a stiff, pre-bent steel tube with
oversized pipe diameter may be chosen.
Figure C11.6.3-4 illustrates a typical detailing for looped tendons used in practice to provide reinforcement
transverse to the plane of the tendon loop and to tie back half of the deviation forces in the plane of the loop to
the concrete on the outside of the loop, in the form of U-shaped bars.
Conclusions
‐ The provisions of FprEN 1992-1-1:2022, Clause 11.6.3, may be used to determine minimum radii of
tendon curvature for internal tendons, external tendons and for loops in internal tendons.
‐ In order to arrive at similar radii of curvature according to FprEN 1992-1-1:2022, Clause 11.6.3, as used
in practice and specified in [C11.6-1], [C11.6-2] and [C11.6-3] as relevant, the maximum transverse
bearing stresses of pRd = 15MPa, 30MPa and 70MPa for internal tendons, external tendons and loops
in internal tendons, respectively, as specified in FprEN 1992-1-1:2022, Table 11.4 (NDP) should be
used. However, if NSBs wish to apply more conservative values for tendon curvatures, i.e. larger radii
of tendon curvature, the values of Table 11.4 (NDP) can be reduced.
‐ Whatever values of pRd are chosen in FprEN 1992-1-1:2022, Clause 11.6.3, the actual radii of curvature
should not be smaller than the values specified in the post-tensioning system documentation of the
actual system to be installed in the structure unless verified by testing.
[C11.6-1] EAD 160004-00-0301, European Assessment Document “Post-tensioning kits for prestressing of
References
structures”, European Organisation for Technical Assessement (EOTA), Brussels, September 2016.
[C11.6-2] fib Model Code for concrete structures 2010, International Federation for Structural Concrete (fib),
Lausanne, 2013.
[C11.6-3] External Prestressing, Service d’Etudes Techniques des Routes et Autoroutes (SETRA), Bagneux,
France, February 1990.
For citations, copying and referencing see Note on Front page! page 484
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
11.7 Background to FprEN 1992-1-1:2023
Introduction This document has been prepared to clarify the background of the design provisions of FprEN 1992-1-1:2023
with respect to Section 11.7 (“Deviation forces due to curved tensile and compressive chords”). These
provisions are new as the topic was not specifically covered by EN 1992-1-1:2004.
Reasons for new Deviation forces due to curved tensile or compressive chords is a common design situation for many members
provisions such as vaulted cut-and-cover tunnels, arches, pipes, shells and similar members (see Figure C11.7.1). Such
deviation forces can be equilibrated by arranging a specific transverse reinforcement (following the general
design procedure of Section 8.5). This solution is however uneconomical (typically in tunnels or pipes) and
unnecessary in many cases, particularly when the deviation forces are low and they can be carried by the
resistance of the concrete cover.
Following this motivation, and also to reduce the probability of accidents (as described for instance in reference
[C11.7.2-3]), a new Subsection has been introduced in the FprEN 1992-1-1:2023 providing explicit rules for
verification of deviation forces of curved reinforcement and compressive chords. The rules are based to a large
extent on Model Code 2010 provisions [C11.7.1], grounded in turn on the works by Fernández Ruiz et al.
[C11.7.2].
Figure C11.7.1: Practical cases where curved reinforcements in tension near can lead to cover spalling
Design approach, A curved reinforcement in tension originates transverse deviation forces following equilibrium conditions, see
background Fig. C11.7.2. When such forces are applied near to the surface and push against the concrete cover, a brittle
failure by cover spalling can occur. This condition is attained when the deviation force (ftr,d = Ftd/r) equals the
effective tensile resistance of the concrete cover (fct,eff,d ∙cu) [C11.7.2].
Figure C11.7.2: Deviation forces of curved reinforcement: (a) deviation forces; and (b) spalling resistance
The value of the effective tensile resistance depends upon a number of parameters [C11.7.2]:
Concrete strength (tensile strength)
Concrete cover
Spacing of reinforcement
Level of deformation of the reinforcement (with higher strains in the reinforcement associated to a
reduced strength to spalling due to its interaction with bond near to cracks)
Size of the member
Interaction with tensile stresses generated by bond in laps
For design purposes, it is normally sufficient to consider that the reinforcement reaches its yield strength
(maximum strain and force in the reinforcement, due to the strain effect, the effective resistance in case of lower
steel stresses would be higher). In that case, according to [C11.7.2], it results:
fct,eff,d = fctd∙el ≈ 0.38 fctd (C11.7.1)
This expression was slightly rearranged and the coefficient rounded for MC2010 [C11.7.1] as:
fct,eff,d = (1/3)∙fctk,0.05/C (C11.7.2)
For FprEN 1992-1-1:2023, a similar Formula is presented, expressed in terms of the compressive strength of
concrete in the following manner:
fct,eff,d = (1/8)∙fck1/2/C (C11.7.3)
As it can be noted, the expression gives almost identical results for low strength concrete (difference lower than
3% for a C30/37) and slightly more conservative estimates for higher concrete strengths.
Deviation of forces in These provisions define the cases concerned by deviation forces (11.7(1)). In general, Section 8.5 can be
curved reinforcement applied for their design (as a discontinuity region), requiring arrangement of transverse reinforcement (11.7(2)).
near to the free surface Alternatively, the Formula in 11.7(3) can be applied for verification of the resistance against deviation forces for
a simple verification or to avoid transverse reinforcement (Formula according to the previous background
11.7(1-3)
[C11.7.1,2]). The width where the effective tensile strength is considered to develop is calculated according to
the geometric definitions of Fig. 11.3 (same definitions as for verifying anchorages and laps).
Curves laps near to the The presence of curved laps is accounted for in this provision. As already observed by Fernández Ruiz et al
free surface [C11.7.2], laps have a detrimental effect on the resistance against cover spalling of curved reinforcement. This
is attributed to the interaction between bond stresses (laps) and deviation forces (curved reinforcement).
11.7(4)
A detailed procedure for such verification is presented in [C11.7.2] (reducing the value of coefficient el in
Formula (C11.7.1)). In order to provide a simpler approach for design, FprEN 1992-1-1:2023 proposes a linear
interaction between the resistance against deviation forces and the resistance against spalling in laps.
References [C11.7.1] International Federation for Structural Concrete, “fib Model Code for Concrete Structures 2010,” Ernst
& Sohn, Berlin, Germany, 434 p., 2013.
[C11.7.2] Fernández Ruiz, M., Plumey, S. and Muttoni, A., Interaction between Bond and Deviation Forces in
Spalling Failures of Arch-Shaped Members without Transverse Reinforcement, ACI Structural
Journal, V. 107, No. 3, 2010, pp. 346-354
[C11.7.3] Intichar, M.; Ebner, M.; and Sparowitz, L., Deviation Forces in Curved Beams (in German:
Umlenkkräfte in gekrümmten Stahlbetonbalken), Österreichische Ingenieur- und Architekten-
Zeitschrift, V. 149, No. 1, Austria, 2004, pp. 11-16.
For citations, copying and referencing see Note on Front page! page 487
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
12 Background to FprEN 1992-1-1:2023
Introduction
This document provides the background to section 12 (Detailing of members and particular rules). The current
Eurocode section 9 (Detailing) has 28 NDPs.
Ease of use
The initial objective of the revision of Section 9 EN1992-1-1:2004 was the reduction of NDPs. However, for the
reasons given below, in many cases some national choice was still required. Therefore, NDPs were
consolidated into clear tables and where helpful some indication of the purpose of the requirements are given. It
is believed that by having the majority of NDP’s in tables they are more readily identified, and this constitutes
ease of use.
A compilation of NDPs in EN1992-1-1:2004 was created as part of the PRB Project, [1]. Its purpose was to
facilitate consensus-building in the reduction of NDPs to enhance ease of use of the Eurocode 2. The following
Figure C12.1 gives on overview of the section 9 NDPs. It can be seen, that in most cases (green boxes) the
recommended values were accepted by the NSBs. However, for every NDP at least some NSBs either adopted
a different value (red) or adopted the recommended value but with further conditions. In many cases the NSB
values/rules are based on local practices and local approaches to safety, not purely on a technical basis. As
such there was a strong demand to keep national choice.
amount of unbonded tendon was relatively small compared to the bonded steel. It was also noted that the
calculated resistance of an unbonded tendon was dependent on assumptions about the overall deflected shape.
On this basis it seemed to be more logical to remove the 1.15 factor but to be more specific as to the basis
under which the resistance of the unbonded (or external) tendon was calculated. To this end a requirement to
calculate the resistance of the unbonded tendon based on the prestressing stress has been added. This avoids
any reliance on the deflected shape.
As NEd affects both the cracking moment and the bending strength and as the same value of prestress is used,
in both cases, for unbonded tendons, it may be thought that little advantage can be taken from these tendons in
satisfying minimum steel requirements. However, the contribution of the prestress is greater when considering
the bending strength. This is because the lever arm between the resultant of the prestressing force and the
centroid of the cracked section is greater than the lever arm in the uncracked case. Therefore, subject to
satisfying other requirements, such as those for tying, it is possible to satisfy the minimum steel requirements
solely with unbonded tendons.
In the case of lightly loaded beams clause 12.2(2) can lead to large amounts of reinforcement. For statically
determinate elements an alternative rule is provided in 12.2(3). This recognises that when the load is
significantly less than the cracking moment, cracking may still occur for other reasons. When such cracking
does occur, the reinforcement must be sufficient to absorb the energy released. This reinforcement can be
calculated by comparing the energy released at cracking with the energy absorbed through the opening of a
single crack, noting that reinforcement less than required by 12.2(2) will not ensure distributed cracking.
The so-called energy balance method is illustrated in Figure C12.2.1, where WI denotes the available internal
dissipation and WE denotes the released potential energy when sudden cracking takes place under constant
external loading, ME.
A more detailed explanation of the model can be found in references [2] & [3]. The model applies to members
where the yield hinge consists of one single crack. This means members where the cross-section yield moment,
My, is smaller than the cracking moment, Mcr, as indicated in Figure C12.2.1. For members where multiple
cracks can develop, i.e. Mcr < My, the model underestimates the deformation capacity and leads to conservative
results.
Figure C12.2.1: Illustration of energy balance concept to verify sufficient strength and deformation capacity
of members containing less reinforcement than “the deem to work” minimum reinforcement
requirement.
It should be noted that the proposed code provision contains a simplification. In the original formulation [3], the
required moment capacity is given as:
2
M Rd ,min kdc M Ed ; kdc
f 1
1 tk 1
f yk
Here, MRd,min corresponds to My as seen in Figure C12.2.1 and the factor ξ is calculated as follows:
u lo u y
f t fy
y
fy
wu f ct 2 f ct 4
wy fy
y lo
2 f ct 4
Where:
fy = yield stress of reinforcement
ft = tensile strength of reinforcement
fct = tensile strength of concrete
φ= bar diameter
εy = yield strain of reinforcement
εu = strain of reinforcement at stress ft
lo = debonding length at vicinity of crack
The debonding length has an effect on the deformation capacity of the member but no effect on the strength of
the cross section. It is always conservative to disregard the debonding length, i.e. assuming lo = 0 mm. By
doing so, a simplified expression for ξ can be obtained:
f
2 u
1 t 1 1
y f
y
The above equation has been used to calculate the kdc-values in FprEN1992-1-1:2023.
To study the effect of the simplification, calculations of the required yield moment, My, as a function of the
applied moment, ME, have been carried out. The results, normalised with respect to the cracking moment, Mcr,
are shown in Figures C12.2.2 – 4. Table 1 shows the input data used in the calculations. It can be seen that the
model requires more reinforcement when lo = 0 mm. The increase is however not significant.
Bar diameter Ø 12 mm
Cover c 20 mm
Class A Class A
(a) (b)
Figure C12.2.2. Class A steel. Required yield moment, My, as function of applied moment, ME. (a) model
with debonding length lo ≠ 0 and (b) model with debonding length lo = 0.
Class B Class B
(a) (b)
Figure C12.2.3. Class B steel. Required yield moment, My, as function of applied moment, ME. (a) model
with debonding length lo ≠ 0 and (b) model with debonding length lo = 0.
Class C Class C
(a) (b)
Figure 12.2.4. Class C steel. Required yield moment, My, as function of applied moment, ME. (a) model
To address issues around anchorage in tension zones Figure 8.5 on EN1992-1-1:2004 has been expanded into
Figure 12.2 of FprEN1992-1-1:2023. The new figure provides additional cases to highlight that laps in closed
stirrups are still acceptable in tension zones whilst the text notes that cases b) to d) (see Figure C12.3.1) with
open stirrups, are not. It is noted that Figure 12.2 a) is not mentioned with regards to its suitability in a tension
zone, this was because different countries had different practices with regards this detail.
Figure C12.3.1. Anchorage of links a)-e) and closing stirrups f)-h) (Figure 12.2 FprEN1992-1-1:2023)
For torsion links in T-sections a further example of an acceptable link has been added to figure 12.3 of
FprEN1992-1-1:2023. This allows “U-Stirrups” to form the link if the laps are confined within the flange. This
removes the risk of anchorage being compromised by corner spalling which occurs in beams predominantly
loaded in torsion. There is also a requirement to design the lap for 1.2fyd to ensure ductile yielding rather than a
brittle bond failure that may invalidate the design model.
It should be noted that the Figure 12.3d) does not show the full lap length within the depth of the flange and this
is an area where further national guidance may be helpful to reflect national traditions.
See separate paper by Schmidt and Hegger, “Background Document to sub-section 12.5.1(3)”
become less ductile, however ductility can be increased by confinement. If only the concrete is considered, a
measure of the plastic rotation capacity can be found by considering the difference between the yield strain of
the concrete and the ultimate strain. In accordance with 5.1.6(3) of FprEN 1992-1-1:2023, this can be presented
as shown in Figure C12.6.1.
It can be seen that whilst the rotation capacity reduces as fcm increases there is a substantial drop off at fcm =
58MPa. However, the above graph ignores the requirement for requirement for minimum confinement links in all
columns and this increases the ductility for all grades of concrete. The current minimum confinement rules do
not necessarily produce consistent ductility even for the same concrete strength. The column ductility can be
related to the ratio of the confinement stress to the design strength of the concrete. It was found that by using
confinement stress of 0.02fcd the ductility of higher strength concrete columns was similar to that of normal
strength concrete columns using the historical deemed to satisfy link requirements. Thus, for concretes with
fck>50 MPa, transverse reinforcement shall be provided to achieve a minimum confinement of k.fcd, where k is a
NDP with recommended value of 0.02. Away from end regions where ductility is less important the principal
function of the minimum links is to prevent buckling of reinforcement. For prestressed columns this is not
relevant and so the minimum spacing based on bar size can be ignored for columns where all longitudinal bars
are prestressed.
In addition, specific rules are given for footings and pile caps without shear reinforcement where the effective
lever arm may be less than that assumed for plane sections remaining plane. Further details are given in [5] and
[6].
existing horizontal tie requirements for walls and columns were maintained.
The tie requirements are summarized in table 12.5(NDP) recognizing that the final values chosen will be NDPs
and ease of use is improved by combining them in a single table.
It should be noted that the second generation version of EN1991-1-7 was not available when FprEN 1992-1-
1:2023 was written, the expectation is that the tie forces will generally align with those given in the first
generation, however as the table is an NDP national bodies can modify if required.
EN1992-1-1:2004 section 10 also included several tables giving minimum dimensions and “ineffective” bearing
lengths. These values could not be justified from first principles and in some situations could be either
conservative or unsafe. Therefore, whilst the need to consider ineffective edge distances is stated it is expected
that the Engineer will use both their judgement, the other material Eurocodes and NCCI to determine the
requirements for their design. Similarly, matters of deviation are dealt with in project specifications and whilst the
need to consider them is stated no values are recommended.
[1] CEN-TC250-SC2-WG1_N0265_Collection_of_EN_1992-1-1_NDPs_2015-08.pdf
References
[2] Hoang, Linh Cao; Ruiz, Miguel Fernández: Minimum reinforcement rules for robustness –
Performance based calculation of As,min for bending problems. Presentation at the 23rd CEN/TC
250/SC2/WG1 meeting, 7-8 November, 2018, Delft.
[3] Hoang, Linh Cao; Hallgren, Mikael: Minimum reinforcement rules for robustness – Performance
based verification by use of NLFEA and simplified Rigid-plastic modelling. Presentation at the 24th
CEN/TC 250/SC2/WG1 meeting, 25-26 March, 2019, Copenhagen.
[4] Ruiz et al. Post-Punching Behaviour of Flat Slabs, ACI Structural Journal, September-
October 2013.
[5] Simões J.T., Faria D.M.V, Fernández Ruiz M., Muttoni A., Strength of reinforced concrete
footings without transverse reinforcement according to limit analysis, Engineering Structures,
2016
[6] Muttoni A, Fernández Ruiz M, Simões JT. The theoretical principles of the critical shear crack
theory for punching shear failures and derivation of consistent closed-form design
expressions. Structural Concrete, 2017
For citations, copying and referencing see Note on Front page! page 496
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
12.2 (4) Background to FprEN 1992-1-1:2023
Introduction This document has been prepared to clarify the background of the design provisions of FprEN 1992-1-1:2023
with respect to the provisions of 12.2(4) concerning the minimum amount of shear reinforcement for ductility
classes B and C. These provisions are new as the topic was not specifically addressed by EN 1992-1-1:2004.
Since the phenomena governing the minimum reinforcement ratio are the same as those governing the
minimum inclination of the compression field in webs, the values of cotmin defined in 8.2.3(4) are also modified
for class A reinforcement.
The clause on minimum shear reinforcement is also applicable to minimum torsion reinforcement. Since torsion
is resolved into shear, the considerations in this document are equally applicable to torsion.
Reasons for change The shear design of beams and slabs is performed in FprEN 1992-1-1:2023on the basis of a variable-angle
truss model. Such approach, consistent with EN 1992-1-:2004, is an equilibrium-based model for design
grounded on the application of limit analysis [C12.2(4).1-4]. Other than equilibrium, the strength conditions of
the materials have to be respected and the materials shall present sufficient deformation capacity. In particular,
it is considered that reinforcement in tension is sufficiently ductile so that failure in the compression field governs
before rupture of the stirrups.
In usual design situations, the deformation capacity of the reinforcement is not relevant to limit the shear
resistance. This might however be the case for members with relatively low amounts of shear reinforcement (as
for members designed with minimum amounts of shear reinforcement). This is justified in these cases by the
fact that relatively flat angles of the compression field can be required to ensure equilibrium, leading to large
deformation demands in the stirrups. Following the relatively high deformation demand of the stirrups, the
ductility class of the reinforcement (defining its deformation capacity together with its hardening properties) can
be governing. With this respect, it is worth mentioning that according to 7.3.3.1(5), “Plastic analysis shall only be
used for reinforcement steel in Class B or C.”
Theoretically, when failure occurs by rupture of the stirrups, the shear resistance depends on the deformation
capacity of the reinforcement. This is due to the fact that rupture of the reinforcement will limit the minimum
angle of the compression field and thus the number of activated stirrups at ultimate. Consequently, when
rupture of the stirrups governs, a reinforcement class A for the stirrups leads to a lower shear resistance than a
Class B or C reinforcement. Other consequence is that, in order to ensure a certain level of resistance with a
minimum amount of reinforcement (assuming stirrup rupture), the amount of stirrups can be lowered for a Class
B or C reinforcement with respect to a Class A reinforcement. These considerations allow for savings in the
amounts of reinforcement and ensure a more uniform level of safety and enhanced consistency with the values
of the compression field recommended in 8.2.3(4).
Design approach, Remark: The material presented hereafter is a concise version of [C12.2(4).7] which is referred to for further
background details.
The main aim of the minimum amount of shear reinforcement is to allow for a safe application of shear design
provisions (VATM) of Section 8.2.3 (“Members with shear reinforcement”). This implies two aspects:
Control of cracking: Avoid crack localization and ensure that the strength reduction factors of the
compression field in 8.2.3 are applicable
Allow for redistribution of internal stresses (variable orientation of compression struts, sufficient
deformation capacity)
The VATM is based on limit analysis and allows for design based only on equilibrium conditions. Nevertheless,
the state of strains is also implicitely considered by means of the efficiency factors for the compression field
(depending on the principal tensile strains) and the limits of compression field inclination cotmin (8.2.3(4)). An
explicit calculation of such strain state is considered in Formula (8.45), where the maximum tensile strain (1) is
estimated assuming that the compressive strain in concrete is 0,001 and that the strain tensor is parallel to the
stress field (assuming cracks to have the same inclination as the compression field):
1 = x +( x + 0.001)∙cot2 (C12.2(4).1)
Consistently with that approach, depending on the horizontal strains (x), the vertical strains can also be
determined (see Mohr’s circle in Figure C12.2(4).1):
y =( x + 0.001)∙cot2- 0.001 (C12.2(4).2)
Such vertical average strain increases as the angle of the compression field reduces as can be seen in Figure
C12.2(4).2 and also for higher levels of longitudinal deformation (x).
1.4
x/y [-]
1.2
0.5
1.0 0.375
0.25
w,ave [%]
0.8
0.125
0.6
0
0.4
0.2 wy
0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
cot [°]
Figure C12.2(4).2: Average vertical strain in the compression field according to Formula (C12.2(4).2)
For some of the cases shown in Fig. C12.2(4).2, the vertical strains reach 1.3%, which indicates the risk of
failures in tension for brittle reinforcement when tension-stiffening effects are considered [C12.2(4).5] (roughly
estimated as half of the rebar strain at rupture uk = 2.5%).
When compared to an extensive database of members with varying deformation capacity of the reinforcement,
such effect can be clearly observed. This fact is presented in Figure C12.2(4).3, where the model of FprEN
1992-1-1:2023 has been compared to selected slender shear beams [C12.2(4).6] by considering a constant
value = 0.5 and cot = 2.5. In that Figure, the value of w,min is adopted as 0.08fc1/2/fy according to Formula
(12.4) of FprEN 1992-1-1:2023 (without modification for other ductility classes). The results show that all tests
with ductility Class C are safely estimated by the code, only one of Class B is (slightly) estimated in an unsafe
manner, but a large number of tests with reinforcement Class A are unsafe, independently of the shear
reinforcement ratio (in some cases, tests with brittle reinforcement and w = 2÷3ꞏw,min show insufficient
resistance). On the other side, tests with ductile reinforcement (classes B and C) and a shear reinforcement
ratio smaller than the minimum value show sufficient resistances.
2.5
Class A
2.0
Class B
Class C
1.5
0.5
prEN 1992-1-1:2018(D3)
v = 0.5 ; cot = 2.5
0.0
0.0 1.0 2.0 3.0 4.0
ρw /ρw,min [-]
Figure C12.2(4).3: Comparison to test results assuming cotmin = 2.5 for all ductility classes
2.5
Class A
2.0
Class B
Class C
1.5
test /R [-]
1.0
0.5
prEN 1992-1-1:2020-11(D7)
(variable limits cotmin)
0.0
0.0 1.0 2.0 3.0 4.0
ρw /ρw,min [-]
Figure C12.2(4).4: Comparison to test results with a penalty for Class A reinforcement in the definition of
cotmin
References [C12.2(4).1] Nielsen M. P., Braestrup M. W. and Bach F. “Rational Analysis of Shear in Reinforced Concrete
Beams.” IABSE Colloquium Proceedings P-15, Vol. 2, 1978, 16 p.
[C12.2(4).2] Thürlimann B. “Plastic Analysis of Reinforced Concrete Beams.” IABSE Colloquium - Plasticity in
Reinforced Concrete, Reports of the working commissions, Vol. 28, 1979, p. 71-90.
[C12.2(4).3] Nielsen M.P., Hoang L.C. Limit analysis and concrete plasticity. 3rd ed. Boca Raton (USA), CRC
Press, 2011, 788 p.
[C12.2(4).4] Sigrist, V., Bentz, E., Fernández Ruiz, M., Foster, S.J., Muttoni, A., Background to the Model
Code 2010 Shear Provisions - Part I: Beams and Slabs, Structural Concrete, Vol. 14, No. 3,
2013, pp. 204--214
[C12.2(4).5] Marti, P., Alvarez. M., Kaufmann, W. and Sigrist, V., “Tension chord model for structural
concrete”, Structural Engineering International, IABSE, Vol 8, No 4, 1998, 287–298.
[C12.2(4).6] Test results from: Huber P., Huber T., Kollegger J., 2016; Yoon Y.-S., Cook W. D., Mitchell D.,
1996; Piyamahant S., Shima H., 2002; Angelakos D., 1999; Bresler B., Scordelis A. C., 1963;
Vecchio F.J., Shim W., 2004; Yoshida Y., 2000; Cladera A., Mari A. R., 2005; F. Monney 2019;
Campana S. et al., 2013
[C12.2(4).7] Monney F, Fernandez Ruiz M, Muttoni A. Influence of amount of shear reinforcement and its
post-yield response on the shear resistance of reinforced concrete members. Structural Concrete.
2022. 1–33. https://doi.org/10.1002/suco.202200331
Background document to
clauses 12.5.1 and 12.4.2
Detailing of punching shear reinforcement -
maximum diameter of punching shear
reinforcement
24.01.2023
For citations, copying and referencing see Note on Front page! page 501
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
12.4.2 and 12.5.1 Background to FprEN 1992-1-1:2023
Introduction
This document has been prepared in order to give a background to the proposed EC2 provisions for the
maximum allowable diameter of punching shear reinforcement Øsw,max in flat slabs and column bases as well as
shear reinforcement assemblies in slabs. The proposal was derived from database evaluations published in [1]
and previously discussed in CEN/TC250/SC2/WG1/TG4.
Database evaluation
A punching database evaluation was conducted to derive application limits and to assess current detailing
provisions as systematic test series usable for quantifying the influence of Øsw,max are not available. The
database was discussed in CEN/TC250/SC2/WG1/TG4 and is more detailed explained in [1]. All tests failing
with any kind of punching failure inside or outside the shear-reinforced zone or on the level of Vmax were
selected for evaluation to get an overview of experimentally tested values of Øsw. Considering deviating
anchorage and thus activation conditions of different types of shear reinforcement, three different categories are
defined in accordance with the given examples in Figure 12.6 from section 12.5.1 of prEC2. Other shear
reinforcement elements which cannot be clearly classified and specific types such as bent-up bars or I-beam
section cuts were excluded for this database evaluation. However, bent-up bars revealing well anchorage can
generally be assigned to the same category as double headed studs. A summary of all selected punching tests
is given in the appendix in Table A and can also be found in [1].
Single leg links and open stirrups exhibiting a small anchorage stiffness
Closed stirrups and bars with similar anchorage exhibiting an average anchorage stiffness
Evaluating and deriving Øsw,max depending on the slab’s effective depth d turned out to be the most suitable
method as the Øsw/d-ratio strongly influences the bond conditions and thus the activation of punching shear
reinforcement [3–5]. Figure X.1 visualizes the database selection and also depicts corresponding permissible
upper limits according to MC2010, Eurocode 2 in combination with the German National Annex (EC2/NAD) and
prEC2.
32 MC2010
32 32
(a) (b) (c)
16 16 16
11 d / 200 10 d /200
8 8 8
MC2010 MC2010
prEC2 prEC2
d >160 mm d >160 mm test d > 160 mm test
0 0 0
0 100 200 300 400 500 600 0 100 200 300 400 500 600 0 100 200 300 400 500 600
Effective depth d [mm] Effective depth d [mm] Effective depth d [mm]
The point cloud evolving from the selected punching tests reveals a degressive trend of the upper limit of tested
Øsw for increasing d and indicates that better anchorage of shear reinforcement may go along with larger
permissible bar diameters. Slabs reinforced with stiffly anchored double headed studs principally featured larger
diameters than slabs with stirrups. The filtered database for shear links is small and represents only a very
limited range of d.
It is now interesting to compare the database with current detailing provisions. The stepwise graduated upper
limit for Øsw,max according MC2010 does not distinguish between the type of shear reinforcement and only
reflects double headed studs adequately. Whereas general EC2 does not mention any upper limit for Øsw at all,
EC2/NAD provides a linear relationship for stirrup-reinforced slabs resulting in overestimating Øsw,max for larger
values of d compared to the experimentally tested values. The prEC2 square root approach, recently defined by
CEN/TC250/SC2/WG1/TG4 using the presented database, represents the upper limit quite well by introducing
simple factors to differentiate between the three shear reinforcement categories (formula shown in Figure X.1).
An additional limitation ensuring proper anchorage of shear reinforcement in slabs is the minimum slab height
specified as hslab ≥ 200 mm according to EC2, EC2/NAD and prEC2 or alternatively defined as dv ≥ 160 mm
according to MC2010 which is also plotted in Figure X.1.
Conclusions
All in all, the presented database evaluation allows for a more profound definition of Øsw,max replacing the roughly
estimated upper limits of MC2010 or EC2/NAD and paves the way for the novel prEC2 approach enabling a
straightforward formula. This continuous definition of Øsw,max distinguishing between three different anchorage
categories can be extended by further categories at any time. Future punching test series with systematically
varying Øsw in combination with different tangential spacings and types of shear reinforcement may allow for
deriving more progressive upper limits.
Although the derivation is based on a punching test database, the proposed limitations for shear links shall also
be applied to shear reinforcement assemblies in slabs (cf. clause 12.4.2. (6)).
Detailed list of database entries can be presented if required.
[1] Schmidt P, Kalus M, Hegger J. Discussion of the level of maximum punching strength of flat slabs
References
and column bases using database evaluations. Structural Concrete 2021 (submitted for publication).
[2] fib Model Code for Concrete Structures 2010, Ernst & Sohn, Berlin, 2013.
[3] Fernández Ruiz M, Muttoni A. Applications of the Critical Shear Crack Theory to Punching of RC
Slabs with Transverse Reinforcement. ACI SJ 2009;106(4):485–94.
[4] Brantschen F. Influence of bond and anchorage conditions of the shear reinforcement on the
punching strength of RC slabs. PhD-Thesis. Lausanne, Switzerland; 2016.
[5] Schmidt P. Punching in Shear-reinforced Flat Slabs and Column Bases. Dissertation. Aachen; 2021.
For citations, copying and referencing see Note on Front page! page 504
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
12.5.2 Background to FprEN 1992-1-1:2023
This document has been prepared to clarify the background of the design provisions of FprEN 1992-1-1:2023
Introduction
with respect to clause 12.5.2 (“Integrity reinforcement against progressive collapse of flat slabs”). These
provisions extend and complete clause 9.4.1(3) of EN 1992-1-1:2004, providing a consistent design approach
for integrity reinforcement in flat slabs.
Flat slabs are particularly vulnerable to punching shear failures. This is due to the fact that a brittle failure at a
Reasons for change
slab-column connection leads to a significant loss of its load-carrying capacity (unless specific measures are
adopted). Thus, following a first punching failure, neighbouring connections are overloaded, with the risk of also
triggering punching failure at them. As a consequence, a local punching at one connection can propagate
through the whole flat slab, leading to its eventual progressive collapse [C12.5.2-1].
The clause 9.4.1(3) of EN 1992-1-1:2004 provides guidance on the detailing of integrity reinforcement, stating
that at least two well-anchored bars shall be arranged in each orthogonal direction on the compression side of a
slab passing over a column. Some Countries had supplementary rules for these bars that required them to have
a tensile capacity equal to the accidental load reaction at the column. If the aim is that such bars would carry the
whole accidental reaction, greater deformation and ductility would be required than is available. These
provisions give good guidance, but are relatively incomplete in terms of design and detailing of such
reinforcement. Recent state-of-the-art on this topic [C12.5.2-1,2] has shown that rational approaches can be
proposed for design of integrity reinforcement. In addition, these works have shown that previous assumption
(full activation of the yield strength of integrity bars in vertical direction to carry the shear force) is clearly too
optimistic. On that basis, Model Code 2010 [C12.5.2-3] provisions a number of rules for the design of the
integrity reinforcement, accounting for the rebar arrangement and ductility class of the reinforcement. In the
revision of EN 1992-1-1:2004, this state-of-the-art was incorporated in an effort to provide design guidance for
integrity reinforcement of new structures and an assessment method for existing ones.
In addition, since punching shear reinforcement is increasingly used in many Countries, there is a need for
consistent provisions on how to combine integrity and punching shear reinforcement.
Slabs without shear Several strategies can be followed to enhance the post-punching resistance of a slab-column connection after a
reinforcement punching failure. They normally consist of arranging shear reinforcement in the form of vertical or bent-up bars
(enhancing the deformation capacity and toughness of the connection) and/or providing the so-called integrity
12.5.2(1) reinforcement. The latter is typically a horizontal reinforcement provided in the compression zone which can
carry the shear force after punching by catenary action. This reinforcement can be activated in tension following
the significant deformations occurring after a punching shear failure.
The performance of the different solutions is highly variable, particularly with respect to the integrity
reinforcement. This is for instance shown in Figure C12.5.2.1 with reference to a number of tests performed by
[C12.5.2-1] considering different integrity reinforcement details. As can be seen, without any shear
reinforcement or reinforcement arranged on the compression side (Figure C12.5.2.1a), a sudden drop in the
load occurs after punching with a relatively low residual resistance (this residual resistance after punching being
associated to the catenary action of hogging reinforcement which is limited by the dowel resistance associated
to the concrete tensile strength [C12.5.2-1]). When longitudinal reinforcement is placed on the compression side
(Figure C12.5.2.1b) or bent-up bars (Figure C12.5.2.1c) are provided, the performance is significantly better,
and levels of load comparable to those leading to first punching can be attained after a reasonable
displacement. These tests also show (Figure C12.5.2.1d) the significance of the anchorage conditions of the
hogging reinforcement, governing both the punching and post-punching performances.
Figure C12.5.2.1: Deviation forces of curved reinforcement: (a) deviation forces; and (b) spalling resistance
(from [C12.5.2-1])
Following these experimental findings, a rational model was proposed by [C12.5.2-1,2] for calculating the post-
punching resistance of flat slabs. The model considered notably:
The contribution of the reinforcement in the compression zone which can be activated by catenary
action after some vertical displacement following punching failure. The performance of such
reinforcement is shown to be highly dependent on its ductility class and the capacity of the concrete
core around the column to anchor it;
The contribution of bent-up bars intersected by the punching cone, whose angle with the slab plane
makes their contribution more efficient for post-punching resistance than straight bars (the catenary
action being activated for significantly smaller deformations);
The contribution of the hogging reinforcement. This contribution depends on the tensile strength of the
concrete cover and its anchorage conditions.
On the basis of the rational model of [C12.5.2-1], simple design provisions were implemented in fib Model Code
2010 [C12.5.2-3]. For such provisions, the following considerations were accounted for:
The contribution of the hogging reinforcement is neglected (as it depends on the tensile strength of
concrete, it only contributes to a relatively low post-punching resistance, Figure C12.5.2.1a and its
efficiency highly depends on the anchorage conditions [C12.5.2-1], Figure C12.5.2.1d);
The contribution of the reinforcement in the compression zone or of bent-up bars is considered by
means of their strength at rupture and their angle b associated to catenary action (refer to Figure
C12.5.2.2a). For calculation of the angle, the deformation capacity of the reinforcement is governing;
A condition is introduced limiting the catenary action by rupture (pull-out) of the concrete core around
the column (Figure C12.5.2.2b).
(a) (b)
Figure C12.5.2.2: Post-punching resistance: (a) dowelling of compression reinforcement and resulting angle
(ult); and (b) rupture (pull-out) of concrete core limiting the angle of dowelled bars (from
[C12.5.2-1])
The resulting expression of fib Model Code 2010 reads:
0.5 fck
VRd ,int As fyd (ft / fy )k sin b d res bint (C12.5.2.1)
c
where As refers to the sections of reinforcement in the compression zone suitably anchored over the column or
passing through it (see definition in Figure C12.5.2.3), fyd to their design value of the yield strength, (ft/fy)k to the
characteristic ratio between tensile and yield strength, b to the angle of the integrity reinforcement at rupture
(0° for class A, 20° for class B and 25° for class C and 40° for bent-up bars class BC, fck to the characteristic
value of the compressive strength of concrete, C to its partial safety factor, dres is the core height between layers
(Figure C12.5.2.2b) and bint the integrity perimeter. As it can be noted, in Formula (C.12.5.2.1), the left-hand
side part refers to the capacity of the catenary action of the reinforcement (initially horizontal or bent-up bars)
while the right-hand side refers to the upper limit related to the capacity of the concrete core to carry the shear
force in the integrity reinforcement.
The value of the post-punching resistance provided by the integrity reinforcement (Formula (C.12.5.2.1)) is to be
compared to the action considering an accidental design situation (with reduced partial safety factors for the
actions, refer to commentary to §7.3.5.6 of MC2010). Also, MC2010 indicates that integrity reinforcement
according to Formula (C.12.5.2.1) is not required provided that the deformation capacity of the slab is sufficient
(MC2010 commentary to §7.3.5.1), which can be ensured by arrangement of conventional shear reinforcement
(MC2010 §7.3.5.6, designed to comply with the minimum amount given in Formula (7.3-68) of MC2010).
Based on the Model Code 2010 provisions, a design expression was proposed for FprEN 1992-1-1:2023, by
considering some simplifications:
The considerations of ductility Class of reinforcement, influencing both ult and (ft/fy)k in Formula
(C.12.5.2.1), can be lumped into one single parameter
The check for the capacity of the concrete core is only governing in special cases (low concrete
strength and/or shallow slabs). It is thus disregarded for the purpose of ease-of-use.
From these considerations, the following design formula results:
where As,int refers to the sections of reinforcement suitably anchored over the column or passing through it (see
Figure C12.5.2.3), kint defines the performance of the integrity reinforcement and depends on the ductility class
and VEd refers to the action under accidental situation. The value of coefficient kint, lumping the information on
the angle of the reinforcement related to catenary action and tensile strength, results 0.37 for Class B
reinforcement and 0.49 for Class C reinforcement. As it can be noted, this formulation is comparable to that of
MC2010 (Formula (C.12.5.2.1)). Due to its limited deformation capacity, class A reinforcement is not allowed as
integrity reinforcement (as in fib MC2010).
Slabs with shear As previously stated, in MC2010, no integrity reinforcement is required provided that shear reinforcement is
reinforcement provided (complying with the conditions of minimum amount given in Formula (7.3-68) of MC2010). Different to
that consideration, FprEN 1992-1-1:2023 states that integrity reinforcement is required for shear-reinforced
12.5.2(2)
slabs when the contribution of the shear reinforcement during the post-punching regime (VRd,w,int) is not sufficient
to ensure the resistance of the connection under accidental actions (VEd):
This Formula considers that, during the post-punching regime, the displacements are sufficiently large to ensure
yielding of the shear reinforcement. This can be considered as a reasonable assumption according to test
results (refer for instance to Figure C12.5.2.1).
When the resistance of the shear reinforcement in the post-punching regime (VRd,w,in) is lower than the shear
demand (VEd), integrity reinforcement has to be arranged according to Formula (12.10) providing a resistance
higher than VEd – VRd,w,int . This considers thus that a part of the shear force is carried by the shear reinforcement
and the remaining part by the integrity reinforcement.
Tests conducted with slabs reinforced with bent-up bars have shown that this type of reinforcement is not only
efficient in increasing the punching shear resistance, but can also be easily activated as integrity reinforcement
[C12.5.2-1] (see also Fig. C12.5.2.1c). For this reason, in fib MC 2010 [C12.5.2-3], bent-up bars are mentioned
as an efficient type of integrity reinforcement. In the approach of FprEN 1992-1-1:2023, such reinforcement can
be considered as a shear reinforcement leading in fact to the same result.
Detailing rules for It is worth mentioning that to ensure the required efficiency, integrity reinforcement should fulfil the detailing
integrity reinforcement rules defined in 12.5.2(1). The bars considered in the design Formula (12.10) should be:
of ductility class B or C (ductility class A reinforcement does not have sufficiently deformation capacity
12.5.2(1)
to activate catenary action),
anchored in the column or passing through it (see in Figure C12.5.2.3 the sections which may be
considered) and
placed on the compression side of the slab within the vertical column reinforcement (ideally, integrity
reinforcement should be placed as a 1st and 2nd layer of reinforcement to ensure sufficient resistance
against rupture of concrete core (see Figure C12.5.2.2b), otherwise, the verification according the
rigth-hand side part of Formula (C12.5.2.1) of Model Code 2010 becomes governing. Concerning the
requirement of the position with respect to the column reinforcement, it is necessary to avoid local
failures at the edge of the column, which can potentially limit the capacity of the integrity
reinforcement to transmit its force to the column).
In addition, punching shear reinforcement which is considered in the reduction of the required integrity
reinforcement according to Formula (12.11) shall enclose the hogging reinforcement according to 12.4.2(4)
(studs should be arranged so that the upper level of their heads is at the upper level of the hogging
reinforcements, see Figure 12.2e) and fulfil the requirements of 12.5 (particularly with respect to the distance
between the column face and the first shear reinforcement in Figure 12.7).
Figure C12.5.2.3: Example of integrity reinforcement over an edge column with 6 sections of bars which may
be considered
Contribution of hogging As shown in Figure C12.5.2.1a, the contribution of significant flexural reinforcement is not negligible. As already
reinforcement described above, the residual resistance after punching associated to hogging reinforcement results from the
catenary action which is limited by the dowel resistance associated to the concrete tensile strength [C12.5.2-1]).
12.5.2(3)
According to [C12.5.2-1], this contribution can be calculated as:
f ck
VRd ,hog nhog bef ,hog (12.12)
C
where nhog is number of bars crossing the control perimeter b0,5 (for the sake of simplicity, compared to [C12.5.2-
1], only the bars inside the control perimeter are considered) and fully anchored at a distance 4d from the
control perimeter (bars crossing two times the control perimeter can be considered twice, whereas bars
anchored within the column are considered only once in a similar manner as shown in Figure C12.5.2.3).
According to [C12.5.2-1], the catenary action of the hogging reinforcement has to be transferred to the concrete
of the slab through dowel action activating the concrete tensile strength (approximated as 0 ,5 f ck ) over an
area 2∙∙ bef,hog for every bar, where bef,hog can be calculated as:
where s is the spacing of the hogging reinforcement and c is the cover of the hogging reinforcement.
Alternative forms of Alternatives to the forms of integrity reinforcement described above can consist of special shear reinforcements
integrity reinforcement or shearheads, providing a significant deformation capacity and a sufficient residual resistance after punching
(an instance based on bars ensuring significant dowel action with ductile behaviour and efficient anchorages is
described in [C12.5.2-4]).
Economic considerations With respect to EN 1992-1-1:2004, the provisions of FprEN 1992-1-1:2023 require more integrity reinforcement
in slabs without punching shear reinforcement and with moderate hogging reinforcement. On the other side, the
new provisions allow combining the contributions of shear, hogging and integrity reinforcement. From an
economic perspective, in most cases, it is more efficient to invest in a reinforcement which can be used for both
the accidental as well as for persistent and transient design situations, allowing to reduce the slab thickness
or/and the column dimensions.
[C12.5.2-1] Fernández Ruiz, M., Mirzaei, Y., Muttoni, A., Post-punching behavior of flat slabs, American
References
For citations, copying and referencing see Note on Front page! page 510
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
13.5.1 Background to FprEN 1992-1-1:2023
Introduction
This document was prepared in order to provide background information to the specifications regarding the
minimum concrete cover for pre-tensioning tendons. A sufficient concrete cover (cmin,b) and clear tendon spacing
(s) shall guarantee crack-free transfer of prestressing within the anchorage area and shall ensure the bond
between concrete and prestressed tendon.
Background
In general, it should be differentiated between the required minimum concrete cover for strands and indented
wires due to the different bond mechanisms.
For strands, the specifications of Table 13.1 are in accordance with the work of NITSCH [1] to avoid splitting of
concrete and were also discussed by GEßNER [2]. In contrast to Model Code 2010 provisions [3], no distinction
is made between normal-strength and high-strength concrete since higher concrete strengths are automatically
accompanied by shorter transfer lengths and thus higher bond stresses leading to larger splitting forces [4].
Even if the tests of DEN UIJL [5] seem to allow smaller concrete covers for high-strength concrete, this effect is
omitted to provide minimum concrete covers on the safe side. The illustration of DEN UIJL [5] shown in Figure
13.1, which was also picked up by GEßNER [2], summarizes the interplay of occurring bond-splitting cracks
depending on concrete cover and clear spacing of tendons.
Figure 13.1: Occurrence of bond-splitting cracks in normal- and high-strength concrete as a function of
strand spacing and concrete cover reproduced from DEN UIJL [5]
Note: “1/3” means 1 specimen with splitting cracks out of 3 test specimens. Grey marked
zone for specimens without splitting cracks
Overall, it must be considered that most tests such as of NITSCH [1] and DEN UIJL [5] were conducted with 0.5’’
strands and that it is not clearly verified whether the relative values of cmin,b/Ø and s/Ø can also be directly
applied for larger strand diameters. In case of members with multiple prestressing tendons in one layer and little
or no transverse reinforcement, NITSCH [1] pointed out that splitting can not be totally ruled out regarding the
recommended minimum concrete cover values. The use of the provisions of clause 13.5 should, without further
evidence, be limited to a maximum strand diameter of 15.7 mm.
For indented wires, previous specifications were given in Model Code 2010 [3]. However, an experimental test
campaign by GEßNER [2] revealed that prestressed indented wires failed more often due to longitudinal cracks
compared to tests with strands or smooth wires. Consequently, a concrete cover of cmin,b/Ø = 2.5 is insufficient
for indented wires. Tests with cmin,b/Ø = 3.5 showed better results [2] but anchorage failure could still not be
excluded with certainty. As a result, an increased concrete cover of cmin,b/Ø = 4.0 is proposed for indented wires,
which is nevertheless smaller compared to the specifications of corresponding technical approvals. In the tests
of GEßNER [2], the clear spacing of prestressed tendons was s > 2.5Ø. To account for smaller clear spacings of
s = 2.0Ø, the concrete cover for indented wires should be increased by 0.5Ø in accordance to the approach of
NITSCH [1] for strands.
Furthermore, the values for the minimum concrete cover given in the technical approvals of pre-tensioning
tendons have generally to be considered. For example for special applications such as hollow-core slabs, other
values of tendon spacing and concrete cover may be used for pre-tensioned members to fullfill all requirements
regarding the ultimate limit state as well as serviceability limit state and providing that satisfactory behaviour in
service is demonstrated by testing or permitted by the specific European product standard with factory
production control.
[1] A. Nitsch, “Spannbetonfertigteile mit teilweiser Vorspannung aus hochfestem Beton“, Dissertation,
References
RWTH Aachen University, Aachen, 2001.
[2] S. Geßner, “Bond and anchorage of pre-tensioning tendons“, Dissertation, RWTH Aachen University,
Aachen, 2019.
[3] International Federation for Structural Concrete, “fib Model Code for Concrete Structures (Model
Code 2010)”, Ernst & Sohn, Berlin, 2013.
[4] G. Bertram, “Zum Verbund- und Querkrafttragverhalten von Spannbetonträgern aus Ultra-hochfestem
Beton“, Dissertation, RWTH Aachen University, Aachen, 2012.
[5] J.A. den Uijl, “Transfer length of prestressing strand in HPC“, Progress in Concrete Research, Vol. 4,
TU Delft, 1995.
For citations, copying and referencing see Note on Front page! page 513
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
13.5.3 and 13.5.4 Background to FprEN 1992-1-1:2022
Introduction
This document has been prepared in order to explain the slight modifications in the presentation of transmission
length and anchorage length of pre-tensioning tendons introduced in FprEN 1992-1-1:2023 when compared
with EN 1992-1-1:2004. These changes are mainly of editorial character with the intent not to significantly
change the actual transmission and anchorage length.
Summary of results
The above changes resulted in the formulae (13.4) and (13.9) for transmission length and anchorage length,
respectively, in FprEN 1992-1-1:2023. There were no changes in the coefficient 1 for release of tendons, nor
for the coefficient 1 for bond conditions:
Comparison of results
Figures C13-5.1 and C13-5.2 show a comparison of transmission and anchorage length according to EN 1992-
1-1:2004 and FprEN 1992-1-1:2023 as a function of the concrete compressive strength at the relevant time,
fck(t). For the comparison, good bond conditions and gradual release of tendons were assumed. In addition, for
the anchorage length, a ratio (pd - pm)/pm0 = 0,38 was assumed. The vertical axis has been made
dimensionless by presenting (lpt x 1)/(p x 1) and (lbpd x 1)/(p x 1), respectively.
Figure C13-5.1: Comparison of transmission length between EN 1992-1-1:2004 and FprEN 1992-1-1:2022
Figure C13-5.2: Comparison of anchorage length between EN 1992-1-1:2004 and FprEN 1992-1-1:2023
Conclusions
‐ The two approaches in EN 1992-1-1:2004 and FprEN 1992-1-1:2023 give very similar results for
transmission length at relatively low concrete strength which is considered typical for transmission.
The differences between the two approaches are slightly larger for wires.
‐ The two approaches in EN 1992-1-1:2004 and FprEN 1992-1-1:2023 give very similar results for
anchorage length at medium to higher concrete strength which is considered typical for ULS
verifications. The differences between the two approaches are slightly larger for wires.
‐ According to the knowledge of the author of this background document, there is no testing available
for strands diameters larger than 15.7mm. Hence, the use of provisions 13.5.3 and 13.5.4 should
without further evidence be limited to a maximum strand diameter of 15.7mm.
References
For citations, copying and referencing see Note on Front page! page 517
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
13.5.5 Background to FprEN 1992-1-1:2023
Introduction
This document has been prepared to explain the changes regarding pure shear tension capacity of prestressed
members without shear reinforcement.
Ibw
fctd
2
VRd,c = l cpfctd (1)
S
In equation (1) the shear resistance is verified by checking the principal tensile stress to the design tensile
strength.
This description of the shear resistance is moved to main clause 13 as it applies only to prestressed members
which usually refers to precast elements. It should be noted that the limitation to simply supported members as
contained in EN 1992-1-1:2004 was removed during the preparation of FprEN 1992-1-1:2023 since the verification
method applies to continuous members also.
The provision in this clause is limited to the effective depth d not larger than 500 mm unless the size effect is
considered in a refined analysis. This is to conservatively consider the size effect on the brittle type of failure
caused by exceeding the tensile strength of plain concrete. The Commentary to EC2 [1], i.e. the background to
EN 1992-1-1:2004, specifically refers to prestressed hollow core slabs when exemplifying special cases where
pure shear tension failure can occur. Hollow core slabs typically do not have effective depths exceeding 500 mm.
There are very few tests on prestressed beams failing in pure shear tension and with d exceeding 500 mm that
are reported in the literature. The ACI-DAfStB database of shear tests on prestressed beams without shear
reinforcement contains only a few tests with d ranging from 575 mm to 825 mm, and only one test with d = 1025
mm, see Figure 1.
Figure 1: Number of tested prestressed beams versus the effective depth d, adopted from [2].
Tests by Bertram [3] include tests on prestressed girders with d up to 917 mm, but these beams also contained
steel fibres. However, the test [3] indicate a decreasing load reserve after shear cracking with increase of beam
height, se Figure 2.
Figure 2: Shear test results of prestressed girders of SFRC without shear reinforcement, from [3].
Resent research by Roosen [4] on shear failure of prestressed girders in regions without flexural cracks also
confirms that the shear strength depends on the size of the region subjected to high tensile stresses.
The provisions for shear resistance of 13.5.5 based on principal stress may be applied to members with larger
effective depth than 500 mm only if the size effect is considered in a refined analysis. However, since beams
typically have a minimum shear reinforcement to Clause 12, it is suggested to alternatively verify shear capacity
with the refined provisions of 8.2.3 accounting for the provided minimum shear reinforcement and considering the
minimal inclination of the compression field under the design prestressing force.
Equation (1) describes the resistance based on the principal tensile stress in the cross section at the centroidal
axis. In case of irregular shaped cross sections, it is very well possible that the maximum principal tensile stress
is found in another fibre than that at the centroidal axis. Therefore, the new clause 13.5.5(2) introduces a limit of
the tensile stress in concrete based on the general formulation of principal tensile stress using Mohr´s circle:
x,Ed x,Ed
𝜎1,Ed 𝑦 𝜏Ed 𝑦 (2)
where the normal stress σx,Ed(y) and the shear stress τEd(y) in the cross section are assessed in a fibre at distance
y from the centroidal axis. The shear stress is defined by
Ed
𝜏Ed 𝑦 (3)
⋅
where VEd is the design shear force at the cross section, S(y) is the first moment of the cross section area, I is the
second moment of the cross section area, b(y) is the width at y.
With this general formulation, the maximum principal tensile stress can be assessed for a cross section with any
given cross-sectional shape.
The design condition follows
ctk,0,05
𝜎1,Ed (4)
where the design tensile strength is given by fctk0.05/C, and is not affected by a factor for high sustained load for
this design.
Another elaboration of this verifcation can be found in EN 1168 [2008], in which the method for prestressed hollow-
core slabs is given and in which also the shear stress due to the transfer of the prestress is considered.
03.01.2023
For citations, copying and referencing see Note on Front page! page 520
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
14.4.5.2 Background to FprEN 1992-1-1:2023
Introduction
This document has been prepared in order to provide additional background information to the revision of the
design rule to the determine the ultimate load bearing resistance of a braced unreinforced concrete wall loaded
by an axial force with a maximum first order eccentricity.
where etot is the sum of the first order eccentricities caused by the loads, the floor wall interaction and additional
eccentricity covering the effects of geometrical imperfections and l0 is the effective length of the wall.
While discussing prEN 1992-1-1:2004/A1:2014 it was proposed to change equation 12.11 to:
Thereafter the several proposals were discussed during the 7th WG1 meeting in Ispra. As a result of that
discussion document WG1/N171 [1] is written. In that document the several proposals were compared and
additional also a comparison with results from structural analyses based on the finite difference method was
made. In these structural analyses a bilinear stress-strain-relation as described in figure 3.4 of
EN 1992-1-1:2004 is used as a starting point for the determination of the moment-curvature relation. From this it
is concluded that:
- formula (3) leads to more conservative results than formula (4);
- formula (4) leads to more conservative results than formula (2);
- formula (2) leads to a capacity which is comparable to the capacity that can be found when a structural
analysis, in this case using the finite difference method, is performed;
- the difference between the theoretical results from the structural analysis and formula (1) is covered by
introducing an additional eccentricity due to creep: e = 0,001 l0.
The proposal from N171 was adapted for EN 1992-1-1:2004/A1:2014. So, in EN 1992-1-1:2004/A1:2014
equation 12.11 [1] was not changed, while the variable e is introduced to contribute to etot. Formula (2) was
maintained in the early drafts of prEN 1992-1-1, where the number of the formula was changed to (14.11).
In WG1/N179 again formula (4) is maintained as French proposal for equation 14.11. In the proposal it is stated
that when using formula (4) the addition of e is not required. Formula 4 is the result of many calculations
performed. Also, it is learned that the effective creep coefficient, calculated from annex B of EN 1992-1-1:2004
is considered in the background of that proposal.
For citations, copying and referencing see Note on Front page! page 521
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
14.4.5.2 Background to FprEN 1992-1-1:2023
0.5
EN 1996‐1‐1:2011 (0.35)
EN 1996‐1‐1:2022 (0.35)
0.4
0.3
0.2
0.1
0.0
0 5 10 15 20 25 30
l0/hw
Figure 14.4.5.2‐1: Comparison of several formulas for F with a variation of l0/hw, for etot/hw ratios of 0,05, 0,20
and 0,35
As a result from the comparison it is learned that when the creep coefficient implictly is taken into account, as
according to the background documentation, is done in formula (4), obvious smaller values for will be found.
Whether the creep coefficient should be applied to a second order analysis for braced unreinforced walls is
discussed between several member states represented in WG1.
It may be noted that from the background information for annex F of prEN 1996-1-1:1999 [2] it can be learned
that in this code for masonry structures creep effects are not considered in the design of vertical loaded walls,
see figure 14.4.5.2-2. In EN 1996-1-1:2005+A1:2012 creep was considered by applying and additional
eccentricity for slender walls (lo/hw 15). In FprEN 1996-1-1:2021 this additional eccentricity is not described.
For citations, copying and referencing see Note on Front page! page 522
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
14.4.5.2 Background to FprEN 1992-1-1:2023
Figure 14.4.5.2‐2: Figure 2 from SC6/N65 background information to annex F of prEN 1996-1-1 [2]
Comparison of the proposed formula with experimental data and the formula for e/h = 0
Therefore a proposal for formula (14.11) in FprEN 1992-1-1:2023 is derived in which the creep coefficient is a
variabel.
Assuming a first order eccentricity which is constant over the length of the wall leads to the largest second order
effect for the particular wall. Therefore the method can be used as a conservative approach where in case of
lateral loading the largest eccentricity resulting from the lateral loading is used in the formula.
With the analysis model iterative calculations are performed for an increasing axial force in which both material
and geometrical non-linear behaviour are considered. From the maximum value of the axial force for which an
equilibrium can be found, the relative factor is derived.
For citations, copying and referencing see Note on Front page! page 523
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
14.4.5.2 Background to FprEN 1992-1-1:2023
1.1
1.0
0.9
0.7
0.6
C20/2
C20/25 5 = e0f f = 0
‐ jeff
0.5
0.4 C20/2
C20/25 5 = e1f f = 1
‐ jeff
0.3 C40/5
C40/50 0 = e0f f = 0
‐ jeff
0.2 C40/5
C40/50 0 = e1f f = 1
‐ jeff
0.1
0.0
0.000 ‐0.001 ‐0.002 ‐0.003 ‐0.004 ‐0.005 ‐0.006 ‐0.007 ‐0.008
strain
Figure 14.4.5.2‐4: Relative stress-strain relation for plain C20/25 en C50/60 with eff equal to 0 and 1
As can be seen from figure 14.4.5.2-4, due to the use of the stress-strain relation as described in figure 5.1 of
FprEN 1992-1-1:2023 the relation between the stress and the strain no longer is linear related to the
compressive strength. Therefore, calculations have to be performed for several concrete grades. For this
C20/25 and C40/50 have been chosen as being a minimum and maximum grade applied in practice of plain
concrete walls.
Results of the several calculations are summarized in table 14.4.5.2-1 and figure 14.4.5.2-5 hereafter.
For citations, copying and referencing see Note on Front page! page 524
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
14.4.5.2 Background to FprEN 1992-1-1:2023
0.9
0.8 G00.05‐20‐0
.05 20 0,0
G00.05‐20‐0.6
.05 20 0,6
G00.05‐20‐1.4
.05 20 1,4
0.7
G00.05‐40‐0
.05 40 0,0
G00.05‐40‐0.6
.05 40 0,6
0.6 G00.05‐40‐1.4
.05 40 1,4
G00.2‐20‐0
.20 20 0,0
G00.2‐20‐0.6
.20 20 0,6
0.5
G00.2‐20‐1.4
.20 20 1,4
G00.2‐40‐0
.20 40 0,0
0.4 G00.2‐40‐0.6
.20 40 0,6
G00.2‐40‐1.4
.20 40 1,4
G00.35‐20‐0
.35 20 0,0
0.3
G00.35‐20‐0.6
.35 20 0,6
G00.35‐20‐1.4
.35 20 1,4
0.2 G00.35‐40‐0
.35 40 0,0
G00.35‐40‐0.6
.35 40 0,6
G00.35‐40‐1.4
.35 40 1,4
0.1
0
0 5 10 15 20 25 30
l0/hw
Figure 14.4.5.2‐5: Maximum value of related to a variation of eccentricity ratio, concrete strength, creep
coefficient and aspect ratio
Based on [3] and the results presented in table 14.4.5.2-1 a new formula for has been derived. It is similar to
that in [3], but with slight modifications proposed by the author of [3], Bo Westerberg:"
l e
1 2,1 0,02 0 tot
h h
= 2 0,6
(5)
l0 etot 0,8 eff fcd,pl
1 0,9 6 20
h h
1000
Where etot is the sum of e0 and ei, e0 is the first order eccentricity caused by the load, ei is the additional
eccentricity covering the effects of geometrical imperfections and fcd,pl is the design value of the compressive
strength of plain concrete.
In the figures 14.4.5.2-6 and -7 hereafter the results of the formula (5) are compared with the results
from the analysis model. Each concrete class considered, is presented in a separate figure.
For citations, copying and referencing see Note on Front page! page 525
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
14.4.5.2 Background to FprEN 1992-1-1:2023
1.0
e t o t /h w -f c k - eff
e t o t /h w -f c k - eff
0.9
0.05‐20‐0
0.05‐20‐0.6
0.8 0.05‐20‐1.4
0.2‐20‐0
0.7 0.2‐20‐0.6
0.2‐20‐1.4
0.6 0.35‐20‐0
0.35‐20‐0.6
0.35‐20‐1.4
0.5
(5) 0.05‐20‐0
(5) 0.05‐20‐0.6
0.4 (5) 0.05‐20‐1.4
(5) 0.2‐20‐0
0.3 (5) 0.2‐20‐0.6
(5) 0.2‐20‐1.4
0.2 (5) 0.35‐20‐0
(5) 0.35‐20‐0.6
(5) 0.35‐20‐1.4
0.1
0.0
0 5 10 15 20 25 30
l0/hw
Figure 14.4.5.2‐6: Comparison for C20/25 between calculation results (dots) and result from
formula (5) (line)
1.0
e t o t /h w -f c k - eff
0.9
0.05‐40‐0
0.05‐40‐0.6
0.8 0.05‐40‐1.4
0.2‐40‐0
0.7 0.2‐40‐0.6
0.2‐40‐1.4
0.6 0.35‐40‐0
0.35‐40‐0.6
0.35‐40‐1.4
0.5
(5) 0.05‐40‐0
(5) 0.05‐40‐0.6
0.4 (5) 0.05‐40‐1.4
(5) 0.2‐40‐0
0.3 (5) 0.2‐40‐0.6
(5) 0.2‐40‐1.4
0.2 (5) 0.35‐40‐0
(5) 0.35‐40‐0.6
(5) 0.35‐40‐1.4
0.1
0.0
0 5 10 15 20 25 30
l0/hw
Figure 14.4.5.2‐7: Comparison for C40/50 between calculation results and formula (5)
From the figures it can be concluded that the formula (5) results in a reasonable value for the factor when it is
compared to the results of a more detailed calculation according to the general method to determine second
order effects. Formula (5) is presented in prEN 1992-1-1 as formula 14-11.
Influence of CE
In the previous, the partial factor for the modulus of elasticity of concrete is taken as 1,2. In table 4.3 of the final
version of FprEN 1992-1-1:2023, the partial factor is a NDP for which a value of 1,5 is advised. Hereafter the
influence of the difference between 1,2 and 1,5 for CE is presented. First this is done by comparing some results
of the general calculations for the different values of CE. For C20/25 and for C40/50 this is done in figure
14.4.5.2-8.
For citations, copying and referencing see Note on Front page! page 526
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
14.4.5.2 Background to FprEN 1992-1-1:2023
1.0 1.0
0.9 0.9
0.5
1.2‐ 0.2‐20‐1.4 0.5 1.2‐ 0.2‐40‐1.4
1.5‐0.2‐20‐0 1.5‐0.2‐40‐0
0.4 0.4
1.5‐0.2‐20‐1.4 1.5‐0.2‐40‐1.4
1.5‐0.35‐20‐1.4 1.5‐0.35‐40‐1.4
0.1 0.1
0.0 0.0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
l0/hw l0/hw
Figure 14.4.5.2‐8: Influence of CE equal to 1,2 or 1,5, for C20/25 (left) and C40/50 (right)
From figure 14.4.5.2-8 it shows, that the influence of the variation of CE is limited, and only significant in case of
higher slenderness’s. Based on this small difference a small adjustment of formula (5) is suggested in case CE
is equal to 1,5, see formula (6) hereafter:
l e
1 2,1 0,02 0 tot
h h
= 2 0,6 (6)
l0 etot 1 eff fcd,pl
1 0,9 6
h h 1000 20
l e
1 2,1 0,02 0 tot
h h
= 2 0,6 (7)
l e 0,67 CE eff fcd,pl
1 0 0,9 6 tot 20
h
h 1000
In the figures 14.4.5.2-9 and -10 hereafter the results of the formula (6) are compared with the results
from the analysis model. Each concrete class considered, is presented in a separate figure.
For citations, copying and referencing see Note on Front page! page 527
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
14.4.5.2 Background to FprEN 1992-1-1:2023
1.0
0.9
e t o t /h w -f c k - eff
0.8 0.05‐20‐0
0.05‐20‐1.4
0.7
0.2‐20‐0
0.2‐20‐1.4
0.6
0.35‐20‐0
0.35‐20‐1.4
0.5
(6) 0.05‐20‐0
(6) 0.05‐20‐1.4
0.4
(6) 0.2‐20‐0
(6) 0.2‐20‐1.4
0.3
(6) 0.35‐20‐0
0.1
0.0
0 5 10 15 20 25 30
l0/hw
Figure 14.4.5.2‐9: Comparison for C20/25 between calculation results (dots) and result from
formula (6) (line)
1.0
0.9
e t o t /h w -f c k - eff
0.8 0.05‐40‐0
0.05‐40‐1.4
0.7 0.2‐40‐0
0.2‐40‐1.4
0.6
0.35‐40‐0
0.35‐40‐1.4
0.5
(6) 0.05‐40‐0
(6) 0.05‐40‐1.4
0.4
(6) 0.2‐40‐0
(6) 0.2‐40‐1.4
0.3
(6) 0.35‐40‐0
0.1
0.0
0 5 10 15 20 25 30
l0/hw
Figure 14.4.5.2‐10: Comparison for C40/50 between calculation results and formula (6)
From the figures it can be concluded that the formula (6) results in a reasonable value for the factor in case
CE is equal to 1,5.
For citations, copying and referencing see Note on Front page! page 528
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
14.4.5.2 Background to FprEN 1992-1-1:2023
For citations, copying and referencing see Note on Front page! page 529
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
20.01.2023
For citations, copying and referencing see Note on Front page! page 530
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
B.3 to B.8 Background to FprEN 1992-1-1:2023
B.3 General Related to Note (1) it should be mentioned that in contrast to the approach for creep in [2], the new creep approach
in [1] separates distinctly between the components basic creep and drying creep. This distinction is also given by
the creep model in [13].
The time development of drying shrinkage (Note (2)) depends on the migration of water molecules through
hardened concrete, which may be perfectly described by Fick’s 2nd law, i.e. drying shrinkage is a diffusion-type
process. This has to be taken into consideration for the modelling of this process.
It should be mentioned that both deformation components, i.e. creep and shrinkage, depend on the compressive
strength of concrete which serves as a substitute parameter for the effect of the composition of concrete (see
background document (BD) to Section 5.1.5. While basic shrinkage increases with the compressive strength, drying
shrinkage decreases. Related to creep, both components, i.e. basic creep and drying creep decrease with
increasing compressive strength.
Referring to Note (3), it has firstly to be stated that investigations carried out in the past show that the average
shrinkage is not significantly affected by short term varying ambient humidities [30]. However, considering seasonal
changes, pronounced differences are observed. In the winter season shrinkage is much lower than in the summer
season [31, 32], what can be considered in design. However, the average shrinkage over years corresponds to the
shrinkage which is obtained for mean constant ambient conditions. This means that reduced shrinkage in winter is
compensated by an increased shrinkage in summer; for more information see also B.6.
Note (4) does not at all mean that the presented models in FprEN 1992-1-1:2023 are not reliable for today’s
concrete. On the contrary, the reliability has been checked carefully even though there are much less experimental
data available from the years 2000 to 2020. However, the reliability is definitively reduced when the concrete
composition deviates from the limiting values defined in EN 206.
Table B.1 indicates the strength development classes of concrete. The related background is indicated in the BD
to Section 5.1.5.
The subsequent Table 1 corresponds very closely to Table B.1 and indicates some editorial improvements and
clarifications. Table 1 was developed by Task Group 7 (TG 7) "Time dependent effects" of CEN TC 250 SC 2 WG
1 during the process of discussing the draft of FprEN 1992-1-1:2023, and all changes compared to Table B.1 were
accepted by WG 1 and SC 2. However, due to an oversight Table B.1 was forgotten to be replaced by Table 1 in
the final document.
For citations, copying and referencing see Note on Front page! page 531
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
B.3 to B.8 Background to FprEN 1992-1-1:2023
Clause (3) defines the range of validity of the creep and shrinkage models, if the structural concrete complies with
EN 206. For further information see BD to Section 5.1.1.
B.4 Development of The aging function for the mean compressive strength (Formulae (B.1) and (B.2)) have principally been maintained
from the previous standard [2]. However, the function for the aging coefficient βcc(t) was extended and readjusted,
concrete strength and compare Formula (B.2) in [1] and Expression (3.2) in [2]. The origin of this equation is reference [12].
stiffness with time While the coefficient s according to Expression (3.2) in [2] depends only on the type of cement used, the new
coefficient sC in Formula (B.2) takes into account the influence of the concrete strength and its development over
time, see Tables B.2 and B.1 in [1]. The values for sC have been derived from new experimental data (see BD to
Annex D). This new approach is much more precise and is also up-to-date, as the influencing variables take into
account today's concrete composition.
The relationship for the aging coefficient βcc(t), i.e. Formula (B.2) was extended by the introduction of the reference
age tref, which can deviate from the usual reference age of 28 days; see BD to Section 5.1.3. For tref = 28 days the
original Expression (3.2) in [2] is obtained.
Formula (B.3) in [1] has been updated compared to Expression (3.4) in [2]; for details see BD to Annex D.
Formula (B.4) in [1] is different to the Expression (3.5) in [2]. The reasons are detailed founded in the BD to Section
5.1.4.
B.5 Basic formulae for Concerning the background of the prediction of the creep coefficient, refer to the BD to Section 5.1.5, Clause (1) -
(3). Further, as the new creep model is very closely related to the corresponding model given in [3], detailed
determining the creep background information may be found in [4]. Moreover, additional information is given in [6, 7]. Nevertheless, few
coefficient aspects will be highlighted subsequently.
When comparing the determination of the creep coefficient in [1] with the corresponding determination in [2], it
becomes evident that by means of Formula (B.5) in [1], a distinct separation of the creep components into basic
creep and drying creep has been realized. From a physical point of few this separation is well founded by the fact
that different microstructural mechanisms prevail for the development of the individual strain components. This
effect is well- known since roughly half a century as the so-called Pickett effect, and is well described in textbooks,
see e.g. [8, 11].
From a practical point of view, i.e. from the intention to derive the best possible approach for estimating the creep
coefficient, the separation into the strain components is associated with two fundamental advantages. Firstly, it is
known that the time development of basic creep and drying creep is different, i.e. different types of mathematical
functions have to be used to describe closely the observed time development. Secondly, if the compressive
strength of concrete is used as a substitute parameter to describe the effect of the concrete composition on creep,
what is the case in all creep models given in standards, the fact has to be faced that basic creep and drying creep
are differently affected by the magnitude of the compressive strength. Test results show that the ratio of basic creep
and drying creep increases significantly with increasing strength. Hence, if there is no separation into the both
creep components, as this is the case for the previous standard [2], neither a correct consideration of the effect of
the compressive strength nor of the effect of the ambient conditions on the time development of creep is possible.
Related to the consideration of the parameter compressive strength of concrete, both for creep and shrinkage
modelling, refer to the BD to Section 5.1.5, Clause (1) - (3). The major aspects related to the time development
functions for creep, i.e. the functions given by Formulas (B.8) and (B.13) are subsequently summarized.
For citations, copying and referencing see Note on Front page! page 532
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
B.3 to B.8 Background to FprEN 1992-1-1:2023
Figure 1: Long-term creep of concrete (specific creep is the ratio of the creep strain and the creep inducing
stress); basic creep (RH = 99 % and 100 %) and drying creep (RH = 50 % and 70 %) data from
Troxell and regression analysis applying different time functions
In Figure 1, the logarithmic function of the type ln [1 + const∙(t-t0)] – this is in principle identical to Formula (B.8) –
is plotted next to the test data, which also visually represents a good regression to the basic creep (measured
values for RH = 99 % and RH = 100 %). In the usual semi-logarithmic representation of creep curves, the time
development changes to a linear development after a slight leftward bend and they have no final value.
If one describes the time development of the basic creep with a logarithmic function and the time development of
the overlapping drying creep with a hyperbolic function according to Formula (B.13), the regression result is
obtained as shown in Figure 1 for the test values with RH = 50 % and RH = 70 %. Although a final value for the
time development is no longer given, a good adaptation to the measured values is achieved even after a very long
period of loading.
Since many of the formulae used today for the design of concrete components require a final creep coefficient, a
calculated final creep coefficient must be defined when using the logarithmic time function for the basic creep. For
this purpose, the service life of the structure can be used as the load duration to be included in the calculation. The
For citations, copying and referencing see Note on Front page! page 533
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
B.3 to B.8 Background to FprEN 1992-1-1:2023
service life, which is usually specified in guidelines, varies between 50 and 200 years depending on the guideline
and/or the type of structure. In FprEN 1992-1-1:2023 [1] it is given as 50 years. For a structural member made of
a usual structural concrete C30/37, which has an effective member size of hn = 200 mm and is loaded at the age
of 28 days in a normal environment (RH = 65 %), the increase in deformation between 50 and 200 years of load
duration amounts to approx. 5-10 % of the creep deformation occurring after 50 years when extrapolating the
superimposed logarithmic and hyperbolic time development functions. For further details see [4].
Figure 2: Time development of creep of concrete loaded at different ages; test data from [18] and
predictions from MC 2010 (curves are identical to predictions of FprEN 1992-1-1:2023) and EN
1992-1-1:2004
A likewise non-affine consideration of the age at loading is given in the representation of the basic creep according
to Formula (B.8) in [1], being first introduced in the MC 2010 [3]. There, the influence of the age at loading t0 is
integrated into the logarithmic function. This representation has the effect that the creep deformations decrease
with increasing loading age, but at the same time the creep curves also run parallel in the usual graphical
representation (strain over the logarithm of time) after longer load duration. This is in good agreement with
experimental investigations and can be justified thermodynamically [19].
For citations, copying and referencing see Note on Front page! page 534
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
B.3 to B.8 Background to FprEN 1992-1-1:2023
Figure 3: Prediction of creep recovery (left) and stress relaxation (right) when applying the principle of
superposition in combination with the creep models given by EN 1992-1-1:2004 and FprEN 1992-
1-1:2023
The product-type approach has the weakness that in combination with the application of the superposition principle
according to Formulae (B.41) or (B.42) under the boundary conditions of full unloading and pure relaxation, it leads
to mechanically nonsensical deformation or stress predictions [5, 21], respectively, see also Figure 3.
Irrespective of the fact that the use of the superposition principle must inevitably lead to erroneous predictions
because concrete is not a linear viscoelastic material as considered by this principle, the strain reversal after
unloading (Figure 3, left) and the change of sign for stress from compression to tension during relaxation (Figure
3, right), both occurring for the creep model in EN 1992-1-1:2004, must be eliminated as far as possible. This can
successfully be realized if the relevant time functions for the age at loading and the time development are matched
in a suitable way, as has been done for MC 2010 [3, 22] and FprEN 1992-1-1:2023, respectively.
Without going into further constitutive details, it may be stated here that a satisfying solution is obtained by the age
dependency of the exponent in Formula (B.13) and the particular combination of aging and time development in
the increment of the logarithmic function of Formula (B.8) in FprEN 1992-1-1:2023. Note, that this is quite different
to the approach in EN 1992-1-1:2004. In this way, the weaknesses of the product-type approach can be overcome
to a large extent. An easier mathematical solution does not exist.
Non-linear creep
In the design of concrete structures, concrete is considered to be a linear viscoelastic material, but in fact it has a
complex non-linearity that depends, among other things, on the magnitude of the applied stress for constant
For citations, copying and referencing see Note on Front page! page 535
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
B.3 to B.8 Background to FprEN 1992-1-1:2023
loading, the ambient conditions, and on the loading history for variable stresses. For details refer to [4, 5, 9, 21,
38].
Formula (B.19) in [1] has been developed originally already for MC 1990 [12] in order to allow for estimating creep
when high initial stresses are applied as this is the case for early high prestressing, even though the overall creep
model is in principal a stress-linear model.
The level of stress at the transition from linear to non-linear creep was fixed to kσ = 0,4 = constant. As a matter of
fact, this level depends actually on the strength of the concrete (porosity, age at loading), which can be neglected
for sake of simplicity. The increasing non-linearity can be described appropriately with the help of the exponential
function (Formula (B.19) in [1]). It is a simple approximation for the non-linear creep behaviour of concrete, but
neglects the influence of the age at loading as well as the influence of the duration of loading. In fact, the non-linear
part of the creep deformation decreases pronouncedly with increasing load duration. Experimental evidence for
Formula (B.19) is shown and discussed e.g. in [4].
In Formula (B.19) the stress level (stress-strength ratio) is defined as kσ = σc / fcm(t0), i.e. the creep generating stress
σc is related to the mean value of the concrete strength fcm(t0) at the age t0. This type of definition is also applied in
[3, 12]. Correspondingly, the limit value in Formula (B.19) for the onset of non-linearity is given as kσ,lim = 0,40. In
contrast, in EN 1992-1-1:2004 [2], the definition kσ* = σc / fck(t0) is used and kσ,lim = 0,45 is set. With this definition
and the given limit value, a somewhat lower stress σc for the transition from linear creep to non-linear creep is
calculated for structural concretes < C60/75. Although this is of no severe practical relevance, conceptually the
approach for kσ* in conjunction with Expression (3.7) in [2] is not correct. The linearity limit and the related
exponential equation were derived directly from experimental investigations, whereby the concretes were
characterized by the experimentally determined value for fcm. In contrast, fck is a characteristic value derived under
safety considerations, starting from the known strength distribution. Therefore, it is consistent to apply Formula
(B.19) using the experimentally determined concrete compressive strength fcm. If this value is not known, it can be
estimated from the concrete strength class or fck using fcm = fck + 8 N/mm².
The non-linearity is much more complicated if the creep-inducing stress is variable. Deviations from the predicted
strain by applying the principle of superposition may occur even if the stress-level is low (kσ < 0,40). Details may
be found in [5], and partially also in [6, 7]. The observed effects can only be described by a complex non-linear
model and require a special superposition principle for the mathematical treatment of creep problems with variable
loads [21]. Further information on the non-linearity presented here can also be found in [23].
B.6 Basic formulae for It should be noted that by definition (sign convention) shrinkage is calculated as a positive strain (elongation) when
applying the Formulae (B.23) to (B.33). In fact, however, shrinkage is a contraction, while swelling is an elongation.
determining the
Important background information on the new shrinkage model in FprEN 1992-1-1:2023, Annex B.6 [1] are already
shrinkage strain given in the BD to Section 5.1.5, Clause (4). It is explained and founded there why the new model gives
approximately 30 to 40 % higher values for shrinkage than given by EN 1992-1-1:2004 [2]. Subsequently, some
additional background information is provided.
As mentioned above, the new shrinkage model in Annex B of [1] is closely related to the corresponding model in
MC 2010 [3]. However, slight improvements have been introduced by the introduction of the Formulae (B.29) to
(B.31) and Formula (B.32) and by replacing the effect of the class of cement (N, S, R) and the strength class of
cement (32,5; 42,5; 52,5), see [3] and [2] as well, by the newly introduced strength development classes of concrete
(see Table B.1 in [1]). Background information is given in the BD to Section 5.1.5, Clause (1) - (3).
The basic approach for determining the shrinkage strain, i.e. the separation of the total strain into the components
basic shrinkage and drying shrinkage has been kept (compare Formula (B.23) in [1] with Expression (3.8) in [2]).
The introduction of the coefficients αNDP,b and αNDP,d in Formulae (B.24) and (B.25) is new. The default values for
both coefficients are 1,0, however, a National Annex may give different values.
From a scientific point of view such a deviation of the coefficients αNDP,b and αNDP,d from 1,0 is only justified, if the
composition of the concrete deviates from the average composition used in Europe. For example, if in a particular
country concretes are primarily produced using high strength limestone aggregate and/or with a low binder content
and/or with binders for which low shrinkage values have been experimentally proven, it makes sense to apply
αNDP,b < 1,0 and αNDP,d < 1,0. It is recommended to run test in order to derive reliable values for αNDP,b and αNDP,d.
For citations, copying and referencing see Note on Front page! page 536
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
B.3 to B.8 Background to FprEN 1992-1-1:2023
Otherwise, the αNDP values should be kept as 1,0, even if for example a high ambient average or a low ambient
average humidity prevails in the particular country. The reason is that the effect of the ambient average humidity is
considered correctly by the formulae (B.29) to (B.32).
Formula (B.26) in [1] which gives the magnitude of the notional basic shrinkage differs from the Expression (3.12)
in [2] in relation to the mathematical form, the application of fcm,28 instead of fck, and by considering the early strength
development of concrete. This results in differences for the prediction of basic shrinkage up to approx. 0,04 ‰
within the validity range of the model, which is not that much. Nevertheless, the new approach is physically correct,
whereas the previous approach is not.
Further, Formula (B.26) is identical to the corresponding Formula in MC 2010 [3], and was already introduced in
MC 1990-99 [24]. The choice of a hyperbolic function as given in the MC 2010 and now in FprEN 1992-1-1:2023
has the advantage that no change of sign occurs in the calculation of the shrinkage values for low strength values
and that a final value for the notional basic shrinkage is reached for high strength values. Both effects make physical
sense but are not given by the simple linear relationship of Expression (3.12) in EN 1992-1-1:2004. Hence, TG 7
(see Introduction) decided to switch for FprEN 1992-1-1:2023 to the corresponding equation of the MC 2020.
Related to the use of fcm in [1] instead of fck in [2], refer to the BD to Section 5.1.3, Clause (1) - (6). The use of the
compressive strength as a parameter, which affects shrinkage (and creep) is elaborately explained in the BD to
Section 5.1.5.
The Formulae (B.29) to (B.31) and Formula (B.32) describe the effect of the ambient humidity RH on the magnitude
of the shrinkage. This parameter is related to the reference value RHeq, which indicates the average relative
humidity inside a member in a state of moisture equilibrium. For an average concrete strength of fcm = 35 N/mm²,
RHeq = 99 %, see Formula (B.32). With increasing concrete strength, the value for RHeq decreases continuously
and as an example amounts to RHeq = 91 % for a concrete strength of fcm = 80 N/mm². The physical consequence
is that in the assumed case of an ambient humidity of RH > RHeq = 91 %, this concrete no longer shrinks but swells,
i.e. its volume increases significantly, because it accumulates water molecules from the environment inside its pore
structure.
The described effect was not correctly considered by Expression (B.11) in [2]. The Formulae (B.29) to (B.32) in [1]
correctly defines three ranges without a discontinuity (jump) at the transition between the ranges. The range
20 % ≤ RH ≤ RHeq covers the drying shrinkage and the range RHeq < RH < 100 % the swelling of the concrete. For
RH = 100 %, swelling also occurs, but this results from water absorption when the member surface is exposed to
water in liquid form. In this case, the capillary transport mechanism of the pore system becomes active, which leads
to increased water absorption and, as a consequence, to an increased volume increase.
The Formulae (B.29) to (B.32) describe only those swelling processes that can occur in a concrete if it has not
previously shrunk as a result of a water release. If, for example, a concrete member has reached a shrinkage of
0.5 ‰ after a long drying phase and is subsequently exposed to a different ambient climate resulting in a continuous
uptake of water molecules (e.g. RH = 99…100 %), the occurrence of the resulting swelling deformation cannot be
determined by means of Formula (B.31). In the relevant literature, no relationships are known with which the
swelling occurring under the conditions mentioned could be calculated. As a guideline for practical applications, it
can be assumed that after long periods of high moisture supply, about 50-60 % of the previously occurring
shrinkage deformation is reversible. Thus, swelling after previous shrinkage of -0.5 ‰ is calculated to +0.25 ‰ to
+0.30 ‰, possibly plus the deformation resulting from the summand "+0.25" according to equation (B.31) in
connection with Formula (B.25), if water can be absorbed in liquid form.
When comparing the functions for the time development for drying shrinkage, it is evident that the Expression (3.10)
in [2] is quite different to the Formula (B.33) in [1]. This is visualized in Figure 4, which gives the time development
of drying shrinkage as predicted from EN 1992-1-1:2004 [2], EN 1992-2:2005 [13] and MC 2010 (identical to FprEN
1992-1-1:2023 [1]).
Concerning the time function of FprEN 1992-1-1:2023, Formula (B.33) in [1], it has to be stated that this kind of
hyperbolic function complies with the diffusion theory and may be directly derived from this process which governs
drying shrinkage.
Firstly, this means that the member size hn affects just the time development exactly according to the square root
((t-ts)/(c∙hn2 + (t-ts))0,5, where the parameter c corresponds to the (average) diffusion coefficient. Hence, the member
size hn does not affect the magnitude of drying shrinkage but the rate of its time development, see e.g. [4].
Moreover, this time function meets very well the average time development observed in hundreds of experiments.
The time function according to EN 1992-1-1:2004 is a purely empirical function that does not meet the typical
development of drying shrinkage of structural concretes. The shrinkage model in EN 1992-2:2005 takes into
account correctly the hyperbolic size effect (t-ts)/(c∙hn2 + (t-ts)) but does not consider the square root for the time
development (see above), which is characteristic for diffusion processes.
For citations, copying and referencing see Note on Front page! page 537
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
B.3 to B.8 Background to FprEN 1992-1-1:2023
The time development of the FprEN 1992-1-1:2023 curve lies roughly in the middle of the two curves being given
by EN 1992-2:2005 [13], apart from the initial phase, when presented in the selected semi-logarithmic scale (see
Figure 4). The silica fume (SF) added to the concrete mix (criterion only "contained" or "not contained") causes a
considerable delay in the drying shrinkage curve, regardless of the quantity, which is at least surprising.
In contrast, the drying shrinkage according to EN 1992-1-1:2004 [2] develops much faster. Half of the final value is
reached here after approx. 150 days, whereas with FprEN 1992-1-1:2023 it is only reached after approx. 700 days.
The possible reason why the time development curve of the EN 1992-1-1:2004 deviates so much from the curve
given by MC 2010 (FprEN 1992-1-1:2023) is subsequently explained.
As already mentioned above and in the BD to Section 5.1.5, the factors 0.85 and kh were introduced in the shrinkage
model in EN 1992-1-1:2004 [2] in order to lower the final values of the calculated drying shrinkage to a level given
by former guidelines (e.g. MC 1978 [25]). Obviously, it was then realised that the initial shrinkage is as well lowered
so that for small member sizes it turned out to be much smaller than predicted by MC 1978. To compensate this
weakness, it was necessary to change the time function (more rapid development of shrinkage, see Figure 4) in
order to increase the calculated values after a short drying period. Hence, the originally considered time
development function – see function in prEN 1992-1-1:2001 [26], which is identical to Formula (B.33) – had to be
replaced. In retrospect, it can be seen that by introducing two “errors” (factors 0.85 and kh) a further error
(accelerated time development) had to be introduced in order to achieve a sufficient adaptation to the shrinkage
model of MC 1978, which was considered to be correct. This procedure ignored the reasons for the change from
the MC 1978 shrinkage model to the MC 1990 shrinkage model (diffusion-type model).
For citations, copying and referencing see Note on Front page! page 538
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
B.3 to B.8 Background to FprEN 1992-1-1:2023
For citations, copying and referencing see Note on Front page! page 539
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
B.3 to B.8 Background to FprEN 1992-1-1:2023
and shrinkage was also investigated (three different square cross-sections, edge length 1000 mm, 500 mm and
250 mm). Parallel to the concreting of large test specimens on the construction site, laboratory specimens were
also produced on site from the same concrete mix. Unlike the large test specimens, they did not remain on the
construction site, but were tested in the laboratory under the typical test conditions (constant 20 °C, 65 % relative
humidity).
The preliminary results obtained show that the MC 2010 model is able to predict the shrinkage deformations
measured on the test specimens in the laboratory as well as on the construction site with mostly good and partly
very good approximation. With regard to the influence of the member size, the tests confirm the validity of Fick's
2nd law for concrete shrinkage.
All in all, the very valuable new tests of Kollegger confirm the fact that the influences of relative humidity and
member size, which are physically correctly taken into account in the diffusion-type shrinkage models, ensure that
laboratory results can on average be accurately transferred to practical conditions. On the other hand, the tests
also show that the seasonal climatic fluctuations exert a pronounced influence with increasing slenderness of the
members that was not previously known or proven in this way and is therefore not represented by shrinkage models
in standards. This applies equally to the influence of the season of the member production on the shrinkage
behaviour of concrete, which was recognised in [31, 33], and which results from a coupled ambient temperature
and moisture effect.
B.7 Tests on elastic The FprEN 1992-1-1:2023 [1] specifies in Section B.7 criteria for the necessity of experimental investigations on
concrete creep and shrinkage as a supplement to the sole application of the prediction models for creep and
deformations, creep and shrinkage. With the Formulae (B.34) to (B.39) relationships are given for an improved model that takes into account
shrinkage the test results by means of additional parameters which are obtained from a regression analysis.
Clause (3) gives the relevant test guidelines. Unfortunately, EN 12390-16 denotes the basic shrinkage as
autogenous shrinkage, as this shrinkage component was termed in former times. Today’s understanding is that
the basic shrinkage results from the superposition of the chemical shrinkage and the autogenous shrinkage. The
chemical shrinkage is the volume decrease caused by the chemical reaction of H2O with the cement clinker phases.
The autogenous shrinkage results from the internal drying as a consequence of the formation of the gel
microstructure of the binder matrix. In both cases there is no moisture exchange with the environment.
The consideration of test results for the considered structural concrete allows the coefficient of variation V for the
prediction of creep and shrinkage to be reduced from approx. V ≈ 30 % to V ≈ 10 %. This increases the factors
given in B.7, Clause (2) from 0,6 and 0,5 to 0,87 and 0,84, respectively, and decreases them from 1,4 and 1,5 to
1,13 and 1,16. Thus, as an example, if the creep coefficient is calculated as φ(50y, 7d) = 3, the upper value for
design is φ0,95(50y, 7d) = 3 ∙ 1,5 = 4,5 if no experiments have been carried out, and drops to
φ0,95(50y, 7d) = 3 ∙ 1,16 = 3,5 (23 % less), if experimental results are considered.
The regression parameters ξbc1, ξbc2, ξds1 and ξds2 are to be determined according to the method of the least squares.
For mathematical reasons, the regression parameters ξbc1 and ξbc2 correlate and cannot be determined
independently of each other. However, this does not imply any limitations.
For such creep and shrinkage tests on usual test specimens, a minimum loading period or drying period of 3 months
must be observed in order to enable a reliable regression of the fitting functions. If drying creep or drying shrinkage
is considered, the strain development curves will show an inflection point in the semi-logarithmic scale (y-axis
strain, x-axis log of time, see of Figures 1 and 4). Waiting for this inflection point in the tests is absolutely crucial
for a reliable regression or extrapolation of the shrinkage and creep process.
Since the drying shrinkage and partially also drying creep is a diffusion-controlled process, the specimen size and
the concrete grade play a decisive role in view of the onset of the inflection point. For usual structural concretes
and specimen diameters of 150 mm, this inflection point occurs roughly after 50 ± 10 days, so that a reliable
extrapolation is possible after a test period of between 90 and 120 days. For slimmer specimens, the inflection
point is reached earlier. This can be estimated using Formulae (B.13) and (B.33).
For creep tests in a moist environment in order to determine the basic creep there will be no inflection point.
However, the minimum loading period of approx. 3 months is sufficient to be able to carry out a reliable regression.
For shrinkage tests in a moist environment in order to determine the basic shrinkage there will occur an inflection
point as well in the semi-logarithmic scale (for mathematical reasons) which is achieved roughly after 30 days. After
roughly 20 days, approx. 60 % of the final strain is reached. Formula (B.27) in [1] may be used for a more precise
estimation.
For citations, copying and referencing see Note on Front page! page 540
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
B.3 to B.8 Background to FprEN 1992-1-1:2023
B.8 Detailed analysis for While in Formula (B.40) the creep function J(t,t0) is described by means of the conventional parameters Ec(t0), Ec,28
and φ(t,t0), i.e. with the elastic moduli and the creep coefficient, Bažant [34, 35] gives a different definition of J(t,t0).
creep at variable loading However, both approaches for J(t,t0) are equally correct if the definitions of the respective deformation components
are consistent with each other.
Formula (B.40) applies exclusively to the effect of a constant stress σc(t0). If, on the other hand, the stress is time-
variable, Formula (B.40) shall not simply use σc(t) instead of σc(t0), as this is sometimes found in the literature, but
for constitutive reasons the superposition principle, also called Boltzmann's superposition principle, shall be used.
According to this, the strain caused by the stress curve σc(t) is obtained by decomposing the stress curve into small
stress increments Δσc(τi) applied at time τi and summing up the strains caused by this, which leads to Formula
(B.41).
It is important to mention that Formulae (B.41) and (B.42) are valid only for a linear viscoelastic material. This holds
not true for concrete, which feature a complex non-linearity. Nevertheless, both Formulae may be applied for
concrete, as it is usually done, and they give in most case reasonably good prediction results. For further details,
see [5, 9, 21].
As a general constitutive relation for concrete, Formula (B.42) allows the treatment of arbitrary creep and relaxation
problems if, given a known material law (creep function J(t,t0)), either the stress history σ(t,t0) or the strain history
ε(t,t0) is known. In the latter case, Formula (B.42) mathematically represents a Volterra's integral equation, which
can only be solved numerically for complex creep functions as used in design today.
For the sake of completeness, it should be mentioned that the viscoelastic behaviour of concrete can also be
described by means of a differential equation that is equivalent to Formula (B.42). However, Formula (B.42) can
only be directly converted into a differential equation if the creep function J(t,t0) fulfils certain mathematical criteria.
Further details can be found in [9, 21].
Since the application of Formula (B.42) is mathematically complex and also requires a high storage capacity for
the numerical solution of creep and relaxation problems - in this respect the differential equation mentioned above
would be more favourable - the effective modulus of elasticity ("age-adjusted effective modulus") according to the
subsequent equation is generally used for calculations in practice:
Ec, 28
Ec, 28, eff
1 ξ t , t0 φ t , t0
This equation uses the "relaxation coefficient" ξ(t,t0) according to Trost [36], ("aging coefficient" according to Bažant
[37]) and thus allows the treatment of creep at variable stresses and strains as a quasi-elastic problem in a simple
algebraic way. Although there are disadvantages associated with this application, such a calculation is possible in
many practical cases with a good approximation because the value for ξ(t,t0) with 0,5 ≤ ξ(t,t0) ≤ 1,0 varies within
narrow limits and can mostly be assumed to be constant as ξ = 0,8. An exact calculation of ξ(t,t0) is complex; details
are given in [10]. The use of equation for Ec,28,eff also has the great advantage that the mathematical type of the
approach for the creep coefficient φ(t,t0) no longer plays a role. Only the accuracy of the prediction of the creep
coefficient is decisive.
[1] FprEN 1992-1-1:2023 (2023) Eurocode 2: Design of concrete structures – Part 1-1: General rules, rules
References for buildings, bridges and civil engineering structures. CEN, Brussels.
[3] Fédération internationale du béton (fib) (2013) fib Model Code for Concrete Structures 2010. Wiley c/o
Verlag Ernst & Sohn, Berlin.
[4] fib Bulletin 70 (2013) Code-type models for structural behaviour of concrete: Background of the
constitutive relations and material models in the fib Model Code for Concrete Structures 2010. State-
of-art report, Lausanne.
[5] CEB Bulletin 199 (1990) Evaluation of the time dependent behavior of concrete. CEB Bulletin
d'Information, No. 199, Comité Euro-International du Béton, Lausanne.
[6] Müller, H. S.; Acosta, F.; Kvitsel, V. (2021) Models for predicting shrinkage and creep of concrete - Part
For citations, copying and referencing see Note on Front page! page 541
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
B.3 to B.8 Background to FprEN 1992-1-1:2023
1: Analysis of the shrinkage model in DIN EN 1992-1-1:2011 and new approach in Eurocode 2 prEN
1992-1-1:2020 in: Beton- und Stahlbetonbau 116, H. 1, pp. 2-18; Part 2a: Creep - Principles and
analysis of the creep model in DIN EN 1992-1-1:2011 in: Beton- und Stahlbetonbau 116, H. 9, pp. 660-
676; Part 2b: Creep - New approach in Eurocode 2 prEN 1992-1-1:2020 in: Beton- und Stahlbetonbau
116, H. 10, pp. 808-820 (in German).
[7] Müller, H. S.; Acosta F.; Kvitsel, V.; Torrenti, J. M. (2023) Prediction of creep and shrinkage – improved
approaches for fib Model Code 2020 in: Structural Concrete, in preparation.
[8] Neville, A. M.; Dilger, W. H.; Brooks, J. J. (1983) Creep of plain and structural concrete. Construction
Press, London, Great Britain.
[9] Bažant, Z. P. (1988) Mathematical modelling of creep and shrinkage of concrete. John Wiley & Sons
Ltd., Chichester, Great Britain.
[10] CEB Bulletin 215 (1993) Structural effects of time dependent behaviour of concrete. Comité Euro-
International du Béton, Lausanne, Switzerland.
[11] Bažant, Z. P; Jirasek, M. (2018) Creep and Hygrothermal Effects in Concrete Structures. Springer
Science+Business Media, Dordrecht, The Netherlands.
[12] CEB-FIP Model Code 1990 (1993) Design Code. Committée Euro-International du Béton (CEB),
Thomas Telford Services Ltd., London.
[14] ACI Committee 209 (2008) Guide for modeling and calculating shrinkage and creep in hardened
concrete. ACI Report 209.2R-08, Farmington Hills.
[15] Troxell, G. E.; Raphael, J. M.; Davis, R. E. (1958) Long time creep and shrinkage tests of plain and
reinforced concrete. ASTM Proceedings, Vol. 58, pp. 1101-1120.
[16] Stöckl, S. (1981) Experiments on the influence of the load level on the creep of concrete. German
Committee for Reinforced Concrete (Deutscher Ausschuss für Stahlbeton), H. 324, Beuth Verlag
GmbH, Berlin (in German).
[17] Brooks, J. J. (2005) 30-year creep and shrinkage of concrete in: Magazine of Concrete Research, Vol.
57 Issue 9, November 2005, pp. 545-556.
[18] Hilsdorf, H. K.; Rüsch, H.; Kordina, K. (1970) Unpublished experiments at the MPA Munich (in German)
[19] Le Roy, R.; Le Maou, F.; Torrenti, J. M. (2017) Long term basic creep behavior of high performance
concrete: data and modelling in: Materials and Structures 50, H. 1, pp. 1-11.
[20] CEB Bulletin 72/73 (1970) CEB-FIP Model Code 1970: International Recommendations for the design
and construction of concrete structures. Comité Européenne du Béton, Paris, London.
[21] Müller, H. S. (1986) On the prediction of the creep of structural concrete [Dissertation]. University of
Karlsruhe (in German).
[22] Müller, H. S.; Anders, I., Breiner R., Vogel, M. (2013) Concrete: treatment of types and properties in fib
Model Code 2010 in: Structural Concrete, Vol. 14, No 4, pp. 320-334.
[23] Anders, I. (2017) Material law for the description of the creep and relaxation behaviour of young normal-
and high-strength concretes. Karlsruhe Institute of Technology (KIT). Karlsruher Reihe Massivbau,
Baustofftechnologie, Materialprüfung; Heft 73 and Deutscher Ausschuss für Stahlbeton, H. 624, Beuth
Verlag GmbH, Berlin, 2017 (in German).
[24] Müller, H. S.; Küttner, C. H.; Kvitsel V. (1999) Creep and shrinkage models of normal and high-
performance concrete – concept for a unified code-type approach in: Revue française de génie civil,
Vol. 3, No. 3-4, pp. 113-132.
[25] CEB Bulletin 124/125 (1978) CEB-FIP Model Code 1978: International System of Unified Standard –
Codes of Practice for Structures. Comité Européenne du Béton, Paris, Lausanne.
For citations, copying and referencing see Note on Front page! page 542
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
B.3 to B.8 Background to FprEN 1992-1-1:2023
[26] prEN 1992-1-1:2001-04 (2001) Eurocode 2: Design of concrete structures – Part 1: General rules and
rules for buildings. European Committee for Standardization.
[27] Wagner, O. (1958) The creep of unreinforced concrete. German Committee for Reinforced Concrete
(Deutscher Ausschuss für Stahlbeton), H. 131, Verlag Ernst & Sohn, Berlin (in German).
[28] Hilsdorf, H. K., Müller, H. S. (1983) Discussion of a note "Comments on the use of Ross' hyperbola and
recent comparisons of various practical creep prediction models" by Z. P. Bažant and J. C. Chern in:
Cement and Concrete Research, Vol. 13, pp. 441–443.
[29] Rüsch, H., Jungwirth, D., Hilsdorf, H. K. (1983) Creep and Shrinkage. Their Effect on the Behavior of
Concrete Structures. Springer Verlag, New York.
[30] Müller, H. S. and Pristl, M. (1993) Creep and shrinkage of concrete at variable ambient conditions in:
Bažant, Z. P. und Carol I., [Ed.] Creep and Shrinkage of Concrete, Proceedings of the 5th International
RILEM Symposium on Creep and Shrinkage of Concrete (ConCreep 5). & FN Spon, London, pp. 15–
26.
[31] Kollegger J. (2020) Creep and shrinkage - Development of practical models for the prediction of creep
and shrinkage of concrete. Report to the Austrian Association for Structural Engineering (öbv) (in
German).
[32] Kollegger J. et al. (2023) Creep and shrinkage - Development of practical models for the prediction of
creep and shrinkage of concrete. Publication in Beton- und Stahlbetonbau, in preparation (in German).
[33] Suza, D. (2020) Influence of the scale effect and environmental conditions on the creep and shrinkage
of concrete [Dissertation]. Vienna University of Technology.
[34] Bažant, Z. P.; Panula, L. (1978) Practical prediction of time-dependent deformations of concrete in:
Matériaux and Constructions, Vol. 11, No. 65, 1978; Vol. 11, No. 66, 1978; Vol. 12, No. 69, 1979.
[35] RILEM draft recommendation: TC-242-MDC multi-decade creep and shrinkage of concrete: material
model and structural analysis. Model B4 for creep, drying shrinkage and autogenous shrinkage of
normal and high-strength concretes with multi-decade applicability. RILEM Technical Committee TC-
242-MDC (Zdenek P. Bažant, chair) in: Materials and Structures (2015), Vol. 48, pp. 753-770.
[36] Trost, H. (1967) Effects of the superposition principle on creep and relaxation problems in concrete and
prestressed concrete in: Beton- und Stahlbetonbau, H. 10 und 11 (in German).
[37] Bažant, Z. P. (1972) Prediction of concrete creep effects using age-adjusted effective modulus method
in: ACI Journal, Vol. 69.
[38] Torrenti, J. M. (2018) Basic creep of concrete-coupling between high stresses and elevated
temperatures in: European Journal of Environmental and Civil Engineering, Vol. 22, No. 12, pp. 1419-
1428.
[39] Torrenti, J. M.; Benboudjema, F. (2013) Desiccation shrinkage of large structures: is there a size effect?
in: Proceedings of ConCreep9, MIT, Cambridge, MA, USA.
For citations, copying and referencing see Note on Front page! page 543
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
For citations, copying and referencing see Note on Front page! page 544
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
B.9 Background to FprEN 1992-1-1:2023
Introduction
This document has been prepared in order to validate the provisions in FprEN 1992-1-1:2023, Annex B.9, for
relaxation of prestressing steel.
For the effect of initial stress between 70% and 80% of the tensile strength, relaxation values at 1000 hours may
be linearly interpolated between the specified or tested values at these stress levels. For stresses below 70% of
tensile strength, relaxation values at 1000 hours may be linearly interpolated between the value specified at
stress of 70% and zero relaxation loss assumed at a stress of 50% of the tensile strength. Table B.9-1 shows
the assumed losses at 50% as well as the specified maximum values at 70% and 80% of the tensile strength of
the prestressing steel (corresponding to Table B.4 in FprEN 1992-1-1:2023).
Relaxation losses of prestressing steel are sensitive to temperature, see [B.9-1]. Hence, testing of relaxation
B.9(6)
losses at elevated temperature is recommended if the prestressing steel is expected to be exposed over
extended periods of time to temperatures of 30°C or higher. For temperatures below 20°C, the relaxation values
at 20°C may conservatively be assumed. For prestressed members subject to heat treatment, the provisions in
Clause 13 apply regarding relaxation loss during the period of heat treatment.
(a) Actual / specified strength per test (b) Actual / specified strength distribution
Figure B.9-1: Actual / specified tensile strength of 7-wire prestressing strands (sample selection, see text above)
Figure B.9-2 illustrates a limited evaluation of the coefficient kp as a function of the initial stress level. As can be
seen, the coefficient varies with initial stress level. However, kp = 0,16 is a fair value for initial stress levels
between 70% and 80% of the tensile strength.
If actual results of relaxation tests are available, Level 2 approach can be used. Figure B.9-3 shows two
examples of relaxation tests, both performed at an initial stress of 70% of the tensile strength up to 1000 hours,
one on 12.9mm diameter strand, the other on 15.7mm strand of 1860 MPa specified tensile strength. The figure
shows the actual test values together with the extrapolated relaxation loss at 50 years design service life
(highlighted with red circle). The formulae for the best fit straight lines are also shown in the figures.
(a) 12.9mm strand stressed to 70% tensile strength (b) 15.7mm strand stressed to 70% tensile strength
Figure B.9-3: Examples of best fit approximation for relaxation loss based on test results (Level 2 approach)
The extrapolated relaxation losses at 50 years design service life in Figure B.9-3 are 6.1% and 4.1% for the
12.9mm strand and the 15.7mm strand, respectively. These two samples had actual values of 1000 = 1.37% and
1000 = 1.09% for the 12.9mm and 15.7mm strand, respectively. This can be compared with the estimations
based on Level 1 approach of 6.6% based on the maximum specified value of 1000 = 2.5% A limited evaluation
of another 10 relaxation tests from three manufacturers with strands 12.5mm to 15.7mm, tensile strength of
1770MPa and 1860MPa, and stressed to 70% and 80% of actual tensile strength, gave extrapolated relaxation
values at 50 years design service life between 2.7% and 6.6% with an average value of 4.5%. Hence, while with
the estimated loss of 6.6% based on Level 1 approach some conservatism may be present in some cases, it is
not conservative in other and certainly not in all cases.
Conclusions
‐ The changes introduced for relaxation losses in FprEN 1992-1-1:2023, Annex B.9, are justified based
on the above data and are easy to use.
‐ The Level 1 approach based on maximum specified relaxation loss at 1000 hours from product
standards and Formula (B.43) with an assumed coefficient kp = 0.16 give fair values of final relaxation
loss at the end of the design service life. The slight conservatism in some cases is felt to have a
marginal effect, if any, on the design and economy of prestressed structures, in particular for bonded
tendons where relaxation accounts for about one third of the total long-term losses.
‐ For large projects Level 2 approach based on actual relaxation test data or past relaxation testing of
the same type of prestressing steel from the manufacturer may be appropriate in some cases where
the effective prestress is relevant in design.
[B.9-1] fib Model Code for Concrete Structures 2010, Fédération Internationale du Béton, Lausanne, 2013.
References
[B.9-2] A. W. Beeby, R. S. Narayanan and H. Gulvanessian, “Designers' Guide to Eurocode 2: Design of
Concrete Structures”, Institution of Civil Engineers, London, 2005.
[B.9-1] CEN/TC 250/SC 2/WG 1 N 834.
Background to
ANNEX C (NORMATIVE):
REQUIREMENTS FOR MATERIALS
Frank Fingerloos DBV (German Society for Concrete and Construction Technology), DE
Alexander Lindorf DBV (German Society for Concrete and Construction Technology), DE
Lars Meyer DBV (German Society for Concrete and Construction Technology), DE
Tony Jones The Concrete Centre, UK
Werner Fuchs (for C.8) University of Stuttgart, DE
21.02.2023
For citations, copying and referencing see Note on Front page! page 548
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
Introduction
This background document has been prepared to explain the necessity and content of Annex C:
Requirements to Materials.
NOTE This background document is subject to changes according to the ongoing discussion within
CEN/TC250 and its AHG Interface between Eurocodes and hENs. Therefore, the following explanations are
preliminary.
C.1 Use of this annex The design according to Eurocode 2 is based under the assumption that reliable and accurate
C.2 Scope and field of ap- information at a sufficient confidence level is available for the construction materials and construc-
plication tion products used. This information can be obtained from the following sources:
Case 1 from Declarations of Performance (DoP) based on harmonised technical specifications
(hEN/EAD) based on the European Construction Products Regulation (CPR) [1];
Case 2 from other technical provisions (e.g. non-harmonized European standards);
Case 3 from other and, if applicable, project-related specifications.
The required performance of the construction materials and construction products to be fulfilled as
a minimum when used in accordance with Eurocode 2 shall be defined in the form of requirements.
These requirements, which are based on the fitness for use of materials and products, are initially
expected in material or product standards. However, after the changeover from the Construction
Products Directive (CPD) to the CPR in 2013, such requirements may no longer be included in
harmonised specifications (hEN/EAD) based on the CPR. Therefore, they must be specified in
Eurocode 2 for these cases.
The normative Annex C serves this purpose: It is necessary to formulate the requirements for prop-
erties to be fulfilled for those materials and products that are to be used in the design according to
Eurocode 2 under the assumptions given there.
Case 1 provides a technical basis to assess the performance of construction products. The manu-
facturer has to deliver a Declaration of Performance (DoP) in which at least one (essential) char-
acteristic is declared. However, this characteristic can be also irrelevant for design with Eurocode 2.
To ensure a consistent system of parameters needed for the usability of a construction product in
connection with a Eurocode design, the design codes of CEN/TC 250 shall name product charac-
teristics needed for design – if needed in conjunction with threshold values [2]. Therefore, all per-
formance characteristics with necessary requirements (in values or pass criteria of tests) shall be
given in Eurocode 2, Annex C. This is because there is no other document to which the Eurocodes
could refer, and which would contain these necessary requirements and the rules on how to demon-
strate compliance with them. A "voluntary DoP" can be used for those properties that cannot (yet)
be explained via a mandatory DoP because they are not listed in Annex ZA of a hEN or in an EAD,
but which are required for design according to Eurocode 2 [2]. The necessary 'set of properties' is
included in the normative Annex C.
It is expected, that the hENs for reinforcing and prestressing steels will not be finally published in
the Official Journal of EU (OJEU), when this Eurocode 2 is published in 2023. So national product
standards or approvals may be used if compatible with material requirements of Annex C. There-
fore NDP-formulations are established in the code-text as follows:
(…) … steels used for structures designed in accordance with this Eurocode shall comply with
the relevant standards for … steel.
NOTE 1 The National Annex can specify relevant standards for … steel.
NOTE 2 The harmonised product standard EN … for reinforcing steel is currently under development.
Still unsolved is a short-term solution how to implement missing parameters in the scope of product
hENs and therefore in the DoPs of the producers. A general way to proceed is given in document
JIS Action 5, 2018.08.24: CPR-Procedure to develop a standardisation request. Unfortunately
based on this it must be expected that it will last about 5 years until new hENs will be drafted and
published in the OJEU [2].
Because the new Eurocode 2 should be finalised in 2023 a short-term solution with Annex C is
established. Three general case scenarios for construction product properties were detected (see
Figure BD.C.1).
Case 1 contains properties which are covered in harmonised technical specifications (hENs/EADs)
or in non-harmonised documents and do not need any modification. Furthermore, Case 1 is divided
into the sub-cases 1a), 1b) and 1c).
– 1a) Since the required properties in terms of accuracy, reliability and confidence level are fully
covered by the harmonized technical specification based on the CPR, no further action is re-
quired in Case 1a).
22.02.2023
For citations, copying and referencing see Note on Front page! page 549
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
– 1b) In Case 1b) a parameter is covered by the system of the CPR, but it is not listed in an Annex
ZA (and therefore cannot be declared in a DoP). The required level of accuracy, reliability and
the level of confidence have to be established in Eurocode 2.
– 1c) For parameters not addressed in harmonised specifications, a declaration of the existing
performance in a (mandatory) DoP is not given, see case 1c). Here, the information must be
declared otherwise – e.g. in a voluntary DoP. The basis for such a voluntary DoP can also be
non-harmonized rules.
Case 2 describes cases where a required property is in principle addressed in an EN or in another
rule but requires an adaptation rule in the Eurocode regarding accuracy, reliability or confidence
level. A simple declaration of performance in a voluntary DoP would be insufficient here.
Case 3 occurs when no EN or other rule exists for a required property. In this case, requirements
for the property must be defined in the Eurocode – including regulations on accuracy, reliability and
confidence level.
Figure BD.C.1. The different cases for determining the properties necessary for the design [2]
Additional information on the requirements applicable to a product in a specific Member State can
be obtained from the national Product Contact Point for Construction established by each Member
State.
A mechanism for completing the standards and strengthening the DoP should be established in
the EU-Member States (e.g. with the support of industry) by drawing up up-to-date lists of the re-
quired essential characteristics for construction products for certain uses. Member States shall de-
termine under their own responsibility the performance requirements for the essential characteris-
tics in relation to the intended use of a construction product. In the context of the currently ongoing
consultations on the revision of the CPR, corresponding comments from the industry (FIEC, CPE,
SBS) were submitted and presented as "Option F" [3]. The Commission is also aware of this prob-
lem and has launched a so-called acquis process to drive forward the completion of the standards
[4].
Whilst the preceding topics concerns Members of the European Union, the Eurocodes assumes
non-EU states adopting the Eurocodes will have similar procedures.
C.1 (1) Precast concrete Under the umbrella of the non-harmonised EN 13369:2018: Common rules for precast concrete
products products [5] the most of precast concrete products are covered by harmonised standards (hEN).
The manufacturers shall draw up a declaration of performance DoP based on an appropriate hEN.
The difficulty with precast concrete elements is that the rules for the concrete used are not harmo-
nised in the member states and to that extent differ. The same applies currently and until the final
publication of prEN 10080/prEN 10370 for reinforcing steel and of EN 10138 for prestressing steel.
In addition, different national implementations of FprEN 1992-1-1 are to be expected for the design.
Against this background, the following requirement is formulated in C.1 (1) in FprEN 1992-1-1:
(1) …This Normative Annex also applies to precast concrete products according to EN 13369 de-
signed according to this standard.
Precast concrete products that do not take these national differences into account may not be used
in another member state without further specifications. In this respect, compliance with the rules
valid in the place of use is relevant for usability – not compliance with the rules in the place of
manufacture.
22.02.2023
For citations, copying and referencing see Note on Front page! page 550
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
In EN 206 [6]/EN 12620 [7] the limitation of the aggregate size is designated in terms of lower
(d) and upper (D) sieve sizes expressed as d / D. In the design the values of Dmax shall be specified
with Dupper (for minimum cover, spacing of bars, tendons and couplers) and Dlower (shear and
punching resistance):
– Dlower: smallest value of upper sieve size D for the coarsest fraction of aggregates in the concrete
permitted by the specification of concrete,
– Dupper: largest value of upper sieve size D for the coarsest fraction of aggregates in the concrete
permitted by the specification of concrete
– Dmax: Declared value of upper sieve size D of the coarsest fraction of aggregates which is actu-
ally used in the concrete
The definition of D via Dlower in EN 12620 can lead to a range of aggregate gradings. EN 206 does
not specify a minimum coarse aggregate content.
The shear and punching models are calibrated against tests carried out with typical gradings. Dlower
should be equal or greater than 8 mm. The use of non-typical aggregate gradings where the per-
centage of larger aggregate sizes in relation to Dlower is small can result in different behaviour. This
can be avoided by specifying grading parameters in addition to Dlower.
It is expected that the new Exposure Resistance Classes (ERC) according to FprEN 1992-1-1-
clause 6.4 will be implemented in a new prEN 206-100. In the event EN 206 or prEN 206-100 does
not refer to ERC, a National Application Document to EN 206 (or a National Annex of EN1992-1-1
can provide the necessary advice on how to implement ERC rules in a country.
C.4 Reinforcing steel Relevant: Case 1b) in Figure BD.C.1 Voluntary DoP for strength and ductility classes required.
C.4.1 Carbon reinforcing
steel For FprEN-1992-1-1-design strength classes (B400, B450, B500, B550, B600, B700) and ductil-
ity classes (A, B, C) of reinforcing steel are required. prEN 10080 [10] does not contain these
classes. Therefore, an interface between the defined FprEN-1992-1-1-classes according to Tables
C.1/C.2 and the DoP according to Annex ZA of prEN 10080 is necessary. Table BD.C.1 contains
the interface conditions. See also CEN-TC250-SC2_N1438 [11].
Intermediate strength classes can be defined and used in the National Annex (see 5.2.2, Table 5.4).
It is expected, that reinforcing steel for ULS and SLS-design is produced under Assessment and
Verification of Constancy of Performance (AVCP) system 1+ addressed in prEN 10080, Table ZA.3.
Each reinforcing steel shall bear on one rib or indentation row, a numerical mark identifying the
manufacturer's works. This mark shall be repeated at an interval of not more than 1,5 m. An addi-
tional voluntary marking with max. 6 characters of Re, Rm, Agt or classes (like 500B) with real
numbers directly on the reinforcing steel is expected to reach the assumed design reliability on
execution on site. A combination with the product identification number allocated by the manufac-
turer himself, added by a code/link to declaration of performance on web is possible.
Table BD.C.1. Interface conditions prEN 10080 [10] to FprEN 1992-1-1
prEN 10080: Table ZA.1 – Relevant clauses for FprEN 1992-1-1
reinforcing steel and intended use
Essential Cl. Classes Notes Required classes and/or threshold
characteristics and/or levels
threshold
levels
Elongation at 5.2 – Declared value For specifying ductility classes A, B, C with
maximum load (10% fractile value) elongation at maximum load as a parameter:
Declared value Agt (used as uk) as 10%
quantileb value for production lotc:
Ductility class A: Agt 2,5 %
Ductility class B: Agt 5,0 %
Ductility class C: Agt 7,5 %
22.02.2023
For citations, copying and referencing see Note on Front page! page 551
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
Indented reinforcement
The indentations shall form an angle of incli-
nation with the longitudinal axis of 35° to 75°.
The indentation flank inclination a shall be
40° to 80°. Requirements of prEN 10080, 7.4,
are accepted.
Shear force 5.3 – Declared values Usable in design and execution of welded
(for lattice girders (minimum values for fabrics and lattice girders
and welded fab- Fw,min, Fs,min and Fd,min) C.4.1(3): In welded fabric the declared mini-
mum shear force value shall be
rics)
Fs ≥ 0,25 Re,nom An (C.1)
where Re, nom is the nominal characteristic
yield strength and An is the nominal cross-
sectional area of either:
a) the larger wire at the joint in a single wire
welded fabric or
b) one of the twin wires in a twin wire welded
fabric (twin wires in one direction)
Stress ratio 5.2 – Declared value For specifying ductility classes A, B, C with
(ult. tens. strength/ (fractile value) ratio tensile strength/yield strength and ratio
tens. yield actual value of yield strength/nominal value of
yield strength (Rm/Re and Re,act/Re,nom)
strength)
as parameters:
22.02.2023
For citations, copying and referencing see Note on Front page! page 552
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
Fatiguea 5.4 – Declared values For reinforcing steel under fatigue design:
(N at 2σa and σmax) Declared values as 10% quantileb value for
production lotc:
Fatigue stress range 2a in testing for N 2 ×
106 cycles based on a maximum stress level
of max = 0,6fyk:
160 MPa for bars/de-coiled bars ≤ 12 mm
140 MPa for bars/de-coiled bars 12 mm <
≤ 16 mm
130 MPa for bars/de-coiled bars 16 mm <
≤ 20 mm and for straight bars > 20 mm
100 MPa for wire fabrics ≤ 12 mm
80 MPa for wire fabrics > 12 mm
Different fatigue requirements can be set in a
National Annex (see footnote b in Table C.1).
Cyclic load 5.7 – Declared value No requirements
strength for NCLS at εc Not used in FprEN 1992-1-1
Strength at 5.6 – Declared value No requirements for normal design
elevated for RT at T NOTE: In FprEN 1992-1-1, 1.1 is stated, that
temperature the design rules are valid under temperature
conditions between –40 °C and +100 °C gen-
erally.
22.02.2023
For citations, copying and referencing see Note on Front page! page 553
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
C.4.2 Stainless reinforcing Relevant: Case 1b) in Figure BD.C.1 Voluntary DoP for strength classes required.
steel
For FprEN 1992-1-1-design strength classes (B400, B450, B500, B550, B600, B700) and ductil-
ity classes (A, B, C) of stainless reinforcing steel are required. prEN 10370 [13] does not contain
strength classes. Therefore, an interface between the defined FprEN 1992-1-1-classes according
to Tables C.3 and C.4 and the DoP according to Annex ZA of prEN 10370 is necessary. Table
BD.C.2 contains the interface conditions.
Intermediate strength classes can be defined and used in the National Annex (see 5.2.2, Table 5.4).
It is expected that stainless reinforcing steel for ULS and SLS-design is produced under Assess-
ment and Verification of Constancy of Performance (AVCP) system 1+ addressed in prEN 10370,
Table ZA.3.
Table BD.C.2. Interface conditions prEN 10370 [13] to FprEN 1992-1-1
prEN 10370: Tables ZA.1x – Relevant clauses for FprEN 1992-1-1
reinforcing steel (stainless steel) to be used for
reinforcement of concrete
Essential Cl. Classes Notes Required classes and/or threshold
characteristics and/or levels
threshold
levels
Elongation 4.1 Threshold Declared value For specifying ductility classes A, B, C with
elongation at maximum load as a parameter:
Declared value Agt (used as uk) as 10%
quantileb value for production lotc:
Ductility class A: Agt 2,5 %
Ductility class B: Agt 5,0 %
Ductility class C: Agt 7,5 %
Weldability 4.2 Threshold Declared steel Usable in design and execution of welded
designation reinforcement with stainless steels designa-
tions is given in EN 10088-1 or EN 10088-5.
Sections and 4.3 Threshold Declared nominal Declared value for max deviation from nomi-
tolerances on diameter nal mass of bar or wire:
sizes Δm 4,5% 8 mm
Δm 6,0% 8 mm
22.02.2023
For citations, copying and referencing see Note on Front page! page 554
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
Bonding strength 4.5 – Declared geometry Usable in design with minimum values of
relative rib area fR,min for ribbed bars:
Ribbed reinforcement
Declared 5% quantileb value for minimum
relative rib area fR,min for production lotc
≤ 6 mm: 0,039
6,5 to 8,5 mm: 0,045
9 to 10,5 mm: 0,052
11 to 50 mm: 0,056
Indented reinforcement
The indentations shall form an angle of incli-
nation with the longitudinal axis of 35° to 75°.
Shear force 4.6 – Declared value Usable in design and execution of welded
according to the fabrics and lattice girders
declared test method C.4.1 (3): In welded fabric the declared mini-
mum shear force value shall be
Fs ≥ 0,25 Re,nom An (C.1)
where Re, nom is the nominal characteristic
yield strength and An is the nominal cross-
sectional area of either:
a) the larger wire at the joint in a single wire
welded fabric or
b) one of the twin wires in a twin wire welded
fabric (twin wires in one direction).
Stress ratio 4.7 Threshold Declared ratio For specifying ductility classes A, B, C with
(maximum ratio tensile strength/0,2 % proof strength
strength/ tensile Rm/Rp0,2 (where Rm should be limited to stress
at 7 % elongation) as parameters:
yield strength)
Declared value Rm/R0,2 as 10% quantileb
value for production lotc:
Ductility class A: Rm/Rp0,2 1,05
Ductility class B: Rm/Rp0,2 1,08
Ductility class C: Rm/Rp0,2 1,15 and < 1,35
Tensile yield 4.8 Threshold Declared value For specifying strength classes B with value
strength of 0,2 % proof strength Rp0,2k (used as fyk) as
parameter:
Declared value Rp0,2 as 10% quantileb value
for production lotc:
Strength class B400: Rp0,2 400 MPa
Strength class B450: Rp0,2 450 MPa
Strength class B500: Rp0,2 500 MPa
Strength class B550: Rp0,2 550 MPa
Strength class B600: Rp0,2 600 MPa
Strength class B700: Rp0,2 700 MPa
Tensile strength 4.9 Threshold Declared value Indirectly specified with declared ratio value
Rm/Rp0,2 for ductility classes
Cyclic load 4.10 – Conformity No requirements
Not used in FprEN 1992-1-1
Fatigue 4.11 – Declared value For stainless reinforcing steel under fatigue
design:
Declared values as 10% quantileb value for
production lotc:
Fatigue stress range 2a in testing for N 5 ×
106 cycles based on a stress ratio min/max =
0,2 for bars and de-coiled products:
200 MPa for ≤ 16 mm
185 MPa for 16 mm < ≤ 20 mm
170 MPa for 20 mm < ≤ 25 mm
160 MPa for 25 mm < ≤ 32 mm
150 MPa for 32 mm < ≤ 50 mm
Different fatigue requirements can be set in a
National Annex (see footnote b in Table C.3).
Strength at 4.12 – Declared value No requirements for normal design
elevated NOTE: In FprEN 1992-1-1, 1.1 is stated, that
temperature the design rules are valid under temperature
conditions between –40 °C and +100 °C
generally.
Assumed parameters for fire design accord-
ing to FprEN 1992-1-2, 5.1 (6):
slope of the linear elastic range Es,θ
proportional limit fsp,θ
maximum stress level fsy,θ
22.02.2023
For citations, copying and referencing see Note on Front page! page 555
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
Table 2 of prEN 10370 [13] contains the appropriate requirements of ductility classes A, B, C
(the upper limit 600 MPa in Note has to be updated). A “voluntary” DoP for the ductility class is
expected.
Fatigue requirements has been included in FprEN 1992-1-1 based on BS 6744:2016 [14]. This
standard adopts the same S-N curves for B500 grade stainless steel reinforcing as for carbon steel
based on a modified regime of testing. The code provision extends this for other grades of stainless
reinforcing steel and provides countries with a national choice as to whether this approach is
adopted or modified.
The minimum relative rib area in prEN 10370 [13] is widely conservative compared with the re-
quirements for carbon steel. Therefore, the design rules for bond in FprEN 1992-1-1 are also valid
for stainless reinforcing steel.
C.5 Prestressing steel Relevant: Case 1b) in Figure BD.C.1 Voluntary DoP for strength classes required.
prEN 10138-3 [16] specifies the requirements and definitions for the essential characteristics of
prestressing bars, which are used for the prestressing of concrete and which are delivered as fin-
ished products in the form of:
– hot rolled bars,
– processed bars,
– plain and threaded bars.
22.02.2023
For citations, copying and referencing see Note on Front page! page 556
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
Part 3 is the first available new draft of prEN 10138. The new parts for wires and strands are still
under preparation.
For FprEN 1992-1-1-design strength classes Y of prestressing steel are required. prEN 10138-3
[16] contains strength classes (grades of prestressing bars) in its Table 4. The steel designation
follows EN 10027 [18]. The FprEN-1992-1-1 classes according to Tables C.5 to C.7 referring to
Table 4 of prEN 10138. Table BD.C.3 contains the interface conditions.
A “voluntary” DoP is expected for the actual tensile strength Rp,act for each individual test-specimen
with these upper limits:
– Y1030: Rp,act 1180 MPa
– Y1050: Rp,act 1210 MPa
– Y1100: Rp,act 1260 MPa
– Y1230: Rp,act 1370 MPa
Intermediate strength classes can be defined and used in the National Annex (see 5.3.2, Table 5.6).
It is expected that prestressing steel for ULS and SLS-design is produced under Assessment and
Verification of Constancy of Performance (AVCP) system 1+ addressed in prEN 10138, Table ZA.3.
Table BD.C.3. Interface conditions prEN 10138-3 [16] to FprEN 1992-1-1
prEN 10138: Tables ZA.1 – Relevant clauses for FprEN 1992-1-1
prestressing steel to be used for
reinforcement of concrete
Essential Cl. Classes Notes Required classes and/or threshold
characteristics and/or levels
threshold
levels
Stress ratio 5.1 – Declared value Rpk/Rp0,1k 1,10
10% fractile value at
90% probability
Tensile strength 5.1 – Declared value For specifying strength classes:
5% fractile value at Declared value Rpk as 5% quantilea value for
90% probability production lotb:
Y1030: Rpk 1030 MPa
Y1050: Rpk 1050 MPa
Y1100: Rpk 1100 MPa
Y1230: Rpk 1230 MPa
Tensile yield 5.1 – Declared value For specifying strength classes:
strength 5% fractile value at Declared value Rp0,1k as 5% quantilea value
90% probability for production lotb:
Y1030: Rpk 835 MPa
Y1050: Rpk 950 MPa
Y1100: Rpk 900 MPa
Y1230: Rpk 1080 MPa
Elongation at 5.1 – Declared value Declared value Agt as 10% quantilea value
maximum force 10% fractile value at for production lotb:
90% probability Agt 3,5 %
Relaxation 5.2 – Declared value Percentage loss of initial force F0 determined
Frt in % of F0 in isothermal relaxation test in accordance
with EN ISO 15630-3 [18].
Relaxation at 1000 h as Frt:
see Table C.7 for required 1000 [%]
Sections and toler- 5.6 – Declared ranges Values of 5.6 accepted.
ances on sizes for m, l and s
Surface geometry 5.5 – Declared ranges No parameters for plain bars. The DoP is
for thread geometry only P.
parameters Full set of the thread parameters of 5.5 for
threaded bars is accepted.
Fatigue 5.3 – Declared value For prestressing steel under fatigue design:
N at a force range Fr Declared values as 10% quantilea value for
and a maximum force production lotb:
Fup Fatigue stress range 2a in testing for N 2 x
106 cycles with an upper limit of Rpk:
200 MPa for plain p ≤ 40 mm
150 MPa for plain p > 40 mm
180 MPa for ribbed p ≤ 40 mm
120 MPa for plain p > 40 mm
Different fatigue requirements can be set in a
National Annex (see footnote a in Table C.7).
Modulus of 5.1 – Declared value 205 000 MPa
elasticity 10% fractile value at
90% probability
22.02.2023
For citations, copying and referencing see Note on Front page! page 557
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
The relevant EAD for Post-tensioning systems for prestressing of structures (see Fig. BD.C.2) is
EAD 160004-00-0301 [19] according to FprEN 1992-1-1, 5.4 Prestressing Systems.
C.6 Couplers Relevant: Case 1b) in Figure BD.C.1 “Voluntary DoP” is required.
The National Annex can specify relevant standards (or national approvals) for couplers designed
in accordance with FprEN 1992-1-1.
The available EAD for ETA-based DoPs is EAD 160129-00-0301: Couplers for mechanical splices
of reinforcing steel bars [20] (pending for citation in OJEU). The sizes of couplers range from 8 mm
to 50 mm. The load bearing parts of the couplers are completely made of steel or cast steel.
Types of couplers covered by the EAD are standard couplers, position couplers, bridging couplers
or transition couplers (see Fig. BD.C.3).
a) Standard coupler
b) Positional coupler
c) Transition coupler
Fig. BD.C.3. Types of couplers [20]
22.02.2023
For citations, copying and referencing see Note on Front page! page 558
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
Concerning product packaging, transport, storage, maintenance, replacement and repair it is the
responsibility of the manufacturer to undertake the appropriate measures and to advise his clients
on the transport, storage, maintenance, replacement and repair of the product as he considers
necessary [20]. It is assumed that all couplers will be installed according to the manufacturer’s
instructions. Relevant manufacturer’s stipulations having influence on the performance of the prod-
uct covered by the EAD [20] shall be considered for the determination of the performance and
detailed ETA or in the technical product information.
It is expected that couplers for ULS and SLS-design are produced under Assessment and Verifi-
cation of Constancy of Performance (AVCP) system 1+ addressed in EAD [20].
Couplers are intended to be used for mechanical splices of reinforcing steel bars for concrete struc-
tures designed according to FprEN 1992-1-1 and prEN 1998-1 for:
– Transfer of axial tension and/or compression forces of the connected,
– Limitation of slip,
– Resistance to high-cycle fatigue loading according to FprEN 1992-1-1,
– Resistance to low-cycle seismic loading according to prEN 1998-1.
The EAD [20] covers the following specifications of the intended use:
– Connection between reinforcing bars avoiding lapped splicing,
– Mechanical splices of reinforcing steel bars with a nominal yield strength of 400 MPa Re,nom
600 MPa and of ductility classes B or C according to FprEN 1992-1-1,
– Mechanical splices of reinforcing steel bars positioned such that the concrete cover complies
with FprEN 1992-1-1.
It shall be ensured that As,nom,bar ꞏ Re,nom,bar ≤ As,nom,coupler ꞏ Re,nom,coupler.
When the EAD is defined as the relevant standard the essential characteristics 1, 2, 3 (see Table
BD.C.4) shall be declared in a DoP addressed in an ETA. The characteristics 4, 5, 6, 7, 8 shall be
declared if relevant in design.
In the DoP, the percentage total elongation at maximum load Agt,act shall also be declared according
to requirements for reinforcing steel in Table BD.C.1.
If couplers are used in fatigue design the requirements of Tables C.1 and C.2 apply. The DoP may
declare other (often lower) values Rsk, k1, k2 for the S-N curve of a coupler. This has to be con-
sidered in an adapted fatigue design and accepted in the project specification. A National Annex
can provide further requirements (e.g. regarding fatigue strength) and advice to couplers.
22.02.2023
For citations, copying and referencing see Note on Front page! page 559
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
C.7 Headed bars The National Annex can specify relevant standards (or national approvals) for headed bars de-
signed in accordance with FprEN 1992-1-1.
Headed bars:
The head of headed bars should have the same steel strength as the reinforcing steel bar. The
head-to-bar connection shall resist to the yield force Fyd of the bar diameter (to declare in the
DoP).
The bearing surface of heads should be at >80 degrees to the bar unless other justified by testing.
The thickness of the head tH should not normally be smaller than ¼ of the maximum head width/di-
ameter h (see Fig. BD.C.4). The geometry of anchoring heads may be circular, quadratic or rec-
tangular.
h
th
t
reinforcing
steel bar with
It is expected that headed bars for ULS and SLS-design are produced under Assessment and
Verification of Constancy of Performance (AVCP) system 1+ addressed in EAD [21].
Table BD.C.5. Essential characteristics of headed bars [21]
Double-headed studs:
The EAD [22] covers double headed studs made of ribbed or smooth shafts and a head at both
ends as punching shear reinforcement elements (see Fig. BD.C.5). The reinforcement elements
comprise at least two double headed studs with the same diameter and shape (ripped or smooth)
and the following specifications:
– double headed studs with shaft diameter dA of 10 mm to 25 mm
– double headed studs with a head diameter dk 3dA
– double headed studs with shafts made of weldable ribbed reinforcement bars or weldable struc-
tural steel with the following characteristics:
yield strength: fyk 500 MPa
ratio of tensile strength over yield strength: (ft/fy)k 1,05
22.02.2023
For citations, copying and referencing see Note on Front page! page 560
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
It is expected that double headed studs for ULS and SLS-design are produced under Assessment
and Verification of Constancy of Performance (AVCP) system 1+ addressed in the EAD [22].
Table BD.C.6. Essential characteristics of double-headed studs [22]
22.02.2023
For citations, copying and referencing see Note on Front page! page 561
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
C.8 Post-installed rein- FprEN 1992-1-1, clause C.8 provides the requirements for product properties to ensure the safe
forcing steel systems use of post-installed reinforcing steel systems in concrete structures.
C.8 (1) Post-installed steel reinforcing bar connections carried out with anchoring mortars are capable of
developing bond behaviour equivalent to cast-in-place steel reinforcing bars if they are properly
performed with an anchoring mortar sufficiently qualified to fulfil the requirements for safety, ser-
viceability, and durability of concrete structures in service and the product specific installation tools
[23]. Correct installation with these tools will ensure a durable bond in the interfaces between the
reinforcing steel bar and the anchoring mortar as well as the anchoring mortar and the borehole
wall.
Therefore, applicability of the installation tools, suitability and allowable use conditions for the an-
ticipated field of application of a system for post-installation of deformed reinforcing steel bars
should be verified by means of a European Technical Product Specification.
The prequalification process given e.g. in EAD 330087 [24] consists of appropriate prequalification
tests taking into account installation and environmental conditions on site and in service as well as
the range of reinforcing steel bar diameters and concrete strength classes at independent tests
laboratories considering the type of loading and an assessment of test results by an independent
organization. The result of the evaluation corresponding to the prequalification procedure shall be
documented in a European Technical Product Specification. This document should contain all in-
formation for the installation and design of post-installed reinforcing steel bars required by FprEN
1992-1-1.
Then, the design of post-installed reinforcing steel bars can be performed according to the provi-
sions given in FprEN 1992-1-1, clause 11.4.8.
C.8(2) The bond behaviour (bond stiffness and bond strength) of post-installed reinforcing steel bars shall
be at least equal with the bond behaviour of cast-in reinforcing steel bars under the same condi-
tions. Then the required minimum mean bond strength fbm,pi,rqd of Table C.7 for bond efficiency
class CPI-1,0 is achieved (bond efficiency factor kb,pi = 1,0) when tested and assessed in accord-
ance with EAD 330087 [24]. The minimum mean bond strength of post-installed reinforcing steel
bars is evaluated from the results of suitable pre-qualification tests taking into account the param-
eters given in C.7(3). The use of anchoring mortars referring to bond efficiency class CPI-1,0 is
recommended.
However, there are anchoring mortars for post-installed rebar applications with characteristics in-
ferior to cast-in-place reinforcing steel bars possible. Their assessment results in a bond behaviour
with bond stiffness and/or bond strength lower than that of cast-in-place reinforcing steel bars. In
this case the bond efficiency classes CPI-0,7 to CPI-0,9 apply, provided the assessment in accord-
ance with EAD 330087 indicates that the corresponding minimum mean bond strength of Table
C.7 for the relevant bond efficiency class is met. The lower bond efficiency class yields an increase
of the design anchorage length by a factor of up to 1/kb,pi = 1/0,7 = 1,43 which might result in
uneconomic design.
The factor kb,pi is product dependent and shall be published in the relevant European Technical
Product Specification. For post-installed reinforcing steel bars where kb,pi < 0,7 a design according
FprEN 1992-1-1 is not possible.
C.8(3) Clause C.8(3) lists all parameters influencing the performance of post-installed reinforcing steel
bars during installation and in service. They are taken into account in the derivation of the required
minimum mean bond strength fbm,rqd of Table C.7 for the determination of the bond efficiency class
in accordance with EAD 330087.
22.02.2023
For citations, copying and referencing see Note on Front page! page 562
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex C Background to FprEN 1992-1-1:2023
References [1] Construction Products Regulation (CPR). Regulation (EU) No. 305/2011 of the European
Parliament and of the Council of 9th March 2011 laying down harmonised conditions for the
marketing of construction products and repealing Council Directive 89/106/EEC.
[2] CEN-TC250-SC2_N2328 Guidance note on the listing and categorization of material and
product parameters required for Eurocode design (State 2019-10-14).
[3] Oscar Nieto et al: CPR Future – Option F https://prezi.com/p/rwlxncpqg3rs/cpr-future-
option-f/
[4] European Commission: CPR Acquis https://ec.europa.eu/growth/sectors/construc-
tion/product-regulation/acquis_en
[5] EN 13369:2018: Common rules for precast concrete products.
[6] EN 206:2021: Concrete: Specification, performance, production and conformity
(EN 206:2013+A2:2021).
[7] EN 12620:2008: Aggregates for concrete.
[8] EN 14889-1:2006: Fibres for concrete – Part 1: Steel fibres – Definitions, specifications and
conformity.
[9] EN 14651:2007: Test method for metallic fibre concrete – Measuring the flexural tensile
strength (limit or proportionality (LOP), residual).
[10] prEN 10080:2020 (new 2023-01): Steel for the reinforcement of concrete – Weldable rein-
forcing steel – General.
[11] CEN-TC250-SC2_N1438_Moersch-Letter_Reinforcing_Steel_EN_10080-2018-09-21
[12] EN ISO 15630-1:2019: Steel for the reinforcement and prestressing of concrete – Test meth-
ods – Part 1: Reinforcing bars, rods and wire.
[13] prEN 10370:2019: Steel for the reinforcement of concrete – Stainless steel.
[14] BS 6744:2016: Stainless steel bars. Reinforcement of concrete – Requirements and test
methods.
[15] EN 10088-1: Stainless steels – Part 1: List of stainless steels.
[16] prEN 10138-3:2020: Steel for the prestressing of concrete – Part 3: Bars
[17] EN ISO 15630-3:2019: Steel for the reinforcement and prestressing of concrete – Test meth-
ods – Part 3: Prestressing steels.
[18] EN 10027: Designation systems for steels.
[19] EAD 160004-00-0301: Post-tensioning systems for prestressing of structures.
[20] EAD 160129-00-0301: Couplers for mechanical splices of reinforcing steel bars.
[21] EAD 160012-00-0301: Headed reinforcement steel bars
[22] EAD 160003-00-0301: Double headed studs for the increase of punching shear resistance
of flat slabs or footings and ground slabs.
[23] Fuchs, W., Hofmann, J.: Post-installed reinforcing bars – Requirements for their reliable use.
Developments in the Built Environment.
DOI link: https://doi.org/10.1016/j.dibe.2020.100040, Elsevier, 2021.
[24] EAD 330087-00-0601: Systems for post-installed rebar connections with mortar.
22.02.2023
For citations, copying and referencing see Note on Front page! page 563
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex D Background to FprEN 1992-1-1: 2023
20.03.2023
Introduction
This document has been prepared to explain and validate the text of the new
Annex D and give reference to the most relevant papers and guidelines.
The work on Annex D has been done within CEN TC250/SC2/WG1/TG7: Time-
dependent effects and CEN TC250/SC2/PT1 (The project team for prEN 1992-1-
1).
The objective of the annex is to give normative guidance for cases where it is
agreed to use EC2 to restrict early age cracking. Together with the new “Annex H:
Guidance on design of concrete structures for water-tightness”, it is meant to
replace the content of the present “EN 1992-3. Eurocode 2 Design of concrete
structures. Part 3: Liquid retaining and containment structures.”
The background for the default values chosen in “Annex B: Time dependent
behaviour of materials: Creep, shrinkage and elastic strain of concrete and
relaxation of prestressing steel, Table B.2” is also included in this document since
these parameters are relevant for both Annex B and Annex D.
Background
D.1 General
Cracks can affect the aesthetics, durability, and tightness, whilst their control has
both economy and sustainability consequences. It is therefore important that
Eurocode 2 gives regulations that clearly describe how and when cracking shall be
accounted for in the structural design. Annex D considers the effects of
temperature changes due to the concrete’s hardening process and basic shrinkage
in combination with imposed deformations occurring at later ages.
The main objective of the methods described is to evaluate the cracking risk and
provide guidance on crack width calculations if cracking is expected to occur. The
focus is on through-cracks, which may span over the whole thickness of the
concrete member. This is typically between 2 and 30 days after casting depending
mainly on the dimension of the structural member and the binder type applied.
The typical temperature and stress histories at early ages are illustrated in Figure
1, where Tci is the fresh concrete temperature, Tc,max the maximum concrete
temperature due to hydration heat, T0 the temperature of the restraining structure,
while ΔTmin is an additional temperature drop which might be due to daily or
seasonal temperature variations. The most unfavourable moment in time for early-
age cracking is denoted tcrit, corresponding to the moment when temperature
equilibrium with the restraining structure is achieved (within 2ºC).
Related to Figure 1a) the most decisive parameters are: The fresh concrete
temperature (Tci), ambient temperature history, insulation conditions, climatic
conditions as wind velocity and solar radiation, temperature of the restraining
structure (To), and finally the additional maximum temperature difference due to
daily and seasonal variations (ΔTmin)
a) b)
Figure 1. (a)Typical temperature history for a structural concrete member, and (b)
corresponding stress history.
In the research literature the term “4000” is denoted the Arrhenius-constant, and
for accurate approaches it is often replaced by the activation energy (ET)
corresponding to the Arrhenius constant times the Gas constant (8,314 J/moloK))
expressed as [6]:
where Ti represents the temperature history, and A and B are material parameters.
The parameter B is 0 for Ti>20oC and therefore only relevant (& necessary) for low
fresh concrete temperatures (winter conditions). According to textbooks the
constants A and B depend on cement type, pozzolan content, and the w/b-ratio.
The typical variation range for A is 20000-40000 kJ/mol, while for B it is 1000-1500
kJ/mol.
For additional input parameters needed for the temperature calculations, see for
instance [5]. The values for the heat conductivity may vary within the range 1,2-3,0
∆J/(s moK). A recommended default value is 2.5 J/(s moK).The thermal diffusivity
corresponds to the conductivity divided by the heat capacity and the specific heat
(mm2/s). The surface convectivity may vary in the range from about 3.3J/(s m2oK)
for 18mm plywood formwork and no wind to about 15 J/(s m2oK) for a free concrete
surface with 5 m/s wind.The heat capacity is defined as: Specific heat ꞏ concrete
density. And the specific heat varies typically in the range 0,85-1,15 kJ/(kg oK).
while in Annex D, a more comprehensive expression is used for ß for cases where
stresses and strength at early ages are important:
ß= 𝑒𝑥𝑝 𝑠 1 (D.3)
The value for tref may be taken between 28 and 91 days, default values for the material
parameter sc are presented in Table B.2 (Annex B), while values for tdor are found in
Table D.1. Compared to present EC 2, the approach is reflecting three novel points:
1- Modification of the formulas adding the last square root-sign term including tref,,
makes it possible to express the strength development with reference strength at
tref later than 28 days without changing the sc-parameter.
2- Introduction of the parameter tdor (Figure 1, describing the “end of the dormant
phase”) defining when stiffness and strength start to be significant. This is a
decisive parameter to consistently include thermal dilation and basic shrinkage.
Both effects may develop fast at early ages, and may typically vary between 8
and 13 maturity hours (0,3-0,5 days Table D.1). The parameter may be strongly
affected by admixtures, and it is important that the choice of the parameter is
consistent with basic shrinkage and temperature development.
3- In Annex B, the three “Strength development classes for concrete” CS, CN and
CR are defined including amounts of the most common cement supplementary
materials. In addition, the parameters sc and tdor are made dependent on the
characteristic compressive cylinder strength (fck).
The default values for sc and tdor are given in Table 1 below, while the background is
explained in the following.
Table 1: Parameters for strength development (includes content in Table B.2 and D.1)
CS CN CR
Strength sc tdor(d) sc tdor(d) sc tdor(d)
fck ≤ 35 MPa 0.6 0,5 0. 5 0,45 0.3 0,35
35 MPa <fck < 60 MPa 0.5 0,45 0.4 0,4 0.2 0,3
fck ≥ 60 MPa 0.4 0,4 0.3 0,35 0,1 0,3
Two approaches were used to assess the default parameters, firstly a data base
including 105 mixes, from the references [7-11] plus unpublished data from NTNU was
established, and parameters defined by optimization procedures. Secondly formulas for
“sc” published by Vollpracht, Soutsos and Kanavaris [12], were used. In the end it was
decided to use only one significant number for the parameters, and the findings from the
two approaches were combined. For both approaches Formula B.2 was used (ie tdor=0).
The database consists of 105 test series with a majority of “modern” concretes. The
series are distributed among the strength and strength development classes as shown in
Table 2 below.
Table 2: Distribution of the 105 test series in the data base used to determine the
strength development parameter “sc”.
CS CN CR
fck ≤ 35 MPa 2 13 6
35 MPa <fck < 60 MPa 6 36 17
fck ≥ 60 MPa 6 9 10
Since the parameters refer to characteristic cylinder strength at 28 days, the following
formulas were used to transfer from the strengths reported:
In the paper by Vollpracht et al [12] 861 experimental series were included, many of
them containing slag or fly ash as supplementary SCM’s. The authors established
formulas for the sc-parameter dependent on the w/b-ratio and the amount of SCM (scm)
included:
𝑤 𝑠𝑐𝑚
𝑠 𝑐 𝑐
𝑏 𝑏
For the lower and higher strength classes the following values for the parameters c1 and
c2 were found:
Cements (normal and higher strengths): 42.5 R, 52.5 N, 52.5 R: c1=0,481, c2=0,441
In the statistical evaluation the standard deviations are of the same size of order as for
the present data base previously described.
To utilize these formulas to verify, adjust, and increase the general validity of the sc-
values obtained from the data base the well-known Bolomey’s formula [9] was used to
establish a relation between the w/b-ratio and the mean cylinder strength:
The transfer from mean to characteristic strength was carried out as described
previously, and the parameters chosen for K and α were shown to give reasonable
agreement with a selection of the results in the data base. It was assumed that w/b=0,7
is representative for fck<35 MPa, w/b=0,35 for fck > 65 MPa, while w/b=0,45 represents
the strength interval in between. It was also assumed that the CS-group of concretes
contains 55% SCM, while the CN-group contains 35% SCM. Based on these
assumptions the sc-values presented in Table 3 below were found.
As seen in Table 3, the finally chosen default values (already reported in Table 1)
correspond approximately to the average values from the data base and the values
determined from Vollpracht et al’s results.
The statistical evaluation based on the data base, showed that for the optimized values
the overall standard deviation was 1,87 MPa. The statistical evaluation reported in [12]
show comparable experience as already mentioned. Combining the two approaches to
the final values gave a standard deviation for the data base on 2,46 MPa.
fck ≤ 35 0,54 0,66 0,6 0.36 0,55 0,5 0,27 0.34 0,3
35<fck< 60 0.44 0,55 0,5 0.29 0,42 0,4 0,26 0.22 0,2
fck ≥ 60 0.41 0,47 0,4 0.29 0,37 0,3 0,25 0,17 0,1
Assessment of tdor
Accurate determination of this parameter requires compressive stress tests at very low
ages, special setting time tests or tests in special TSTM-equipment. The proposed
default values are based on test results and reflects that the parameter in general is
decreasing with increasing compressive strength, and that it is increasing with increasing
SCM-content, see for instance Kanstad et.al. [7], Klausen [8], and Bjøntegaard and
Kjellsen [13]. Some typical results are shown in Figure 2.
Figure 2.(a) Effect of cement type on the parameter tdor and (b) typical experimental
results [13].
Figure 3a) shows the ratio fc(91)/fc(28) for concretes of the data base where the 91-days
strength was available. Although the scatter is large, a tendency with decreasing ratio
with increasing strength can be seen. The scatter is partly explained by a strong
influence of the cement type and pozzolanic content. It should also be noted that the
green cross markers are representing data from a recent project where the aim was to
develop environmentally friendly high strength concretes for the Norwegian “Ferry-free
E39-project” funded by the Norwegian Public Road Administration.
Figure 3b) shows the calculated ratio fc(91)/fc(28) versus the corresponding experimental
ratio using Formula B.2 and the parameters from Table B.2. It is seen that the scatter is
relatively large, but that still the tendency in the equation is reasonable.
Figure 3.(a) Ratio fc(91)/fc(28) versus 28 days strength (fc(28)). (b) Calculated versus
experimental ratio fc(91)/fc(28).
The thermal dilation coefficient (αcth) varies with the concrete age, humidity
conditions and aggregate type within a wide range 7-14 E-6 as seen in Figure 4a)
[14]. The recommended default value is as in the present EC2 10E-6, which is a
reasonable choice as shown in Figure 4a).
The basic (autogenous) shrinkage, εcbs(t), may be calculated from Formula B24. It
should, however, be noted that the uncertainty is large, which yields both the final
value and the time function. See the relevant background documents for Annex B.
An important and relevant aspect for Annex D is that both previous and present
European standards for shrinkage measurements (EN 12390-16) prescribe that
the deformation measurements should start one day (24 h) after casting. There
are, however, concretes which develop a considerable basic shrinkage before 1
day. When relevant this should be considered in the calculations. Figure 4b)
shows a typical situation, and illustrates at the same time the need for coordinating
the thermal dilation and the basic shrinkage with the tdor-parameter.
Furthermore, the temperature influence on the basic shrinkage is not yet fully
understood, and is therefore not sufficiently considered in the model in Annex B.
Figure 4c shows the measured basic shrinkage as a function of the temperature
increase 14 days after casting for three concretes with fly ash addition.
a)
b)
c)
Using parameters described in Table 1 and Figure 1, the tensile stress in a predefined
position, including cumulative impacts, can be determined.
σ1(t)= (ΔεT(t-t2) + εcbs(t-t2))E1 ˑR1 + αcthˑ ΔTmin ˑE2 ˑR2 + εcds(t-t2)ˑE3 ˑR3
(2)
Where t2 is the time when the stress changes sign. For early age problems the time t
should be set equal to tcrit as defined in Figure 1, while if cumulative impacts are included
t is the relevant design service life. ΔεT(t-t2)=kTemp(Tc,max-To)αcth represents the part of the
temperature causing tensile stresses and εcbs(t-t2) is the corresponding part of the basic
shrinkage occurring in the cooling phase and εcds is the drying shrinkage strain. E1
=Ec(t2eq)/(1+k2*φ(t,t2eq)) is the effective E-modulus representing the cooling phase.
Default values based on amongst others Scandinavian practice [6,7] are k1=0,9 and
k2=0,8. The parameters E2 and E3 are corresponding effective E-modules representative
for the additional effects. Three different values for the degree of restraint are used in
formula (2) because this parameter will vary with the concrete age and the time period
considered, and because the structural system may change with time.
The present draft of Annex D [1] includes formulas for the tensile stresses at two
particular stages, expressed by the parameters in Figure 1.
At temperature equilibrium between restrained and restraining member ( 𝑡 ):
𝜎 𝑡 𝑅 , 𝑘 𝛼 𝑇, 𝑇 𝜀 𝑡 𝜀 𝑡 (D.4)
𝜎 𝑅 , 𝑘 𝛼 𝑇, 𝑇 𝜀 𝑡 𝜀 𝑡
,
𝑅 , 𝐸 𝛼 Δ𝑇 𝑅 , 𝜀 𝑡 𝜀 𝑡 (D.5)
,
t2 is the time when the stress changes sign from compression to tension. For the early
age problem at tcrit, kTemp αcth (Tc,max-To) represents the part of the temperature causing
tensile stresses while [εcbs(tcrit)- εcbs(t2)] is the part of the basic shrinkage occurring in the
cooling phase. Default values based partly on Scandinavian practice [16,17], and verified
below are kTemp=0,9 and χφ 0,55.
For the long-term formula, the temperature term is the same, while the effective E-
modules and the shrinkage terms are adjusted to the relevant time-intervals. The
additional temperature change Δ𝑇 corresponding to the daily temperature variation is
added as a short-time effect. The value X=0,8 is in agreement with the general
application of the Trost-Bazant method. Ec,28 is the tangent modulus of elasticity of
A typical test regime used for the verification presented in the following, consist of
hydration heat development, activation energy for maturity calculation, development of
E-modulus versus time, Temperature-Stress-Testing-Machine (TSTM), and free dilation
test results (dummy specimen for the TSTM). For more thorough explanations, see
[8,15,18]. The realistic temperature histories applied are calculated for a 700 mm thick
wall, with both fresh concrete temperature and equilibrium temperature equal to 20oC.
The predefined degree of restraint is either 0,5 or 0,7. Figure 5 shows typical E-modulus
development, while Figure 6 includes three different sets of so far unpublished TSTM-
test results.
a)
b)
c)
Figure 6. Typical TSTM test results including the realistic temperature curve applied, the
tensile strength development, and the measured and calculated temperature history.
(Unpublished results from NTNU)
In the practical applications of the method, available input may vary, and therefore the
following three approaches are chosen in the verification of the method:
1) -t2, tcrit,Tc,max, To , and E-modulus from the experiment. αcth =10-5 (10E-6) and
basic shrinkage from Formula B.26 and B.27.
2) As 1), but with t2=2 days as proposed default value in Annex D
3) As 1), but with experimental values for the free dilation (sum of basic
shrinkage and thermal dilation)
The results are presented in Table 5, 6 and 7. The number in the parenthesis in the last
column represents the ratio between calculated and experimental values. In some cases
it is expected that the first and the second approach may give large scatter because the
basic shrinkage and the thermal dilation coefficient deviate considerably from the default
values. In particular this is the case when a relatively large basic shrinkage occur before
1 day.
In Figure 7, 8 and 9 the calculated stress versus the corresponding experimental value is
presented. And it is seen that the agreement between the calculated and experimental
values is good for all the considered cases.
Table 5. Verification results,1st approach: t2, tcrit,Tc,max, To , and E-modulus from the
experiment. αcth =10-5 and basic shrinkage from Formula B.26 and B.27.
Test description Ec(t2e) Tmax- εcbs(tcrit)- σ(tcrit) calc σ(tcrit) exp
(MPa) To(oK) εcbs(t2) (MPa) (MPa)
(10-6)
w/b=0,38, Cem I SR, 34176 41,8 16 4,32 4,00(1,08)
0% SCM, R=0,5, t2=3d
tcrit=9,25d, To= 20oC
w/b=0,38, Cem II, incl 32633 36,2 17 3,61 3,75(0,96)
FA, R=0,5, t2=3d
tcrit=9,25d, To= 20oC
w/b=0,38, Cem I, R=0,5, 32835 39,5 18 3,95 3,50(1,13)
t2=3d tcrit=9,25d,
To=20oC
w/b=0,48, Cem I SR, 30844 35,2 13 3,28 2,80(1,17)
R=0,5, t2=3,2d
tcrit=9,25d, To= 20oC
w/b=0,48, Cem II incl 28902 28,8 10 2,51 2,55(0,99)
FA, R=0,5, t2=3d
tcrit=9,25d, To= 20oC
Table 6. Verification results, 2nd approach: t2=2 days, tcrit,Tc,max, To , and E-modulus from
the experiment. αcth =10-5 and basic shrinkage from Formula B.26 and B.27.
Test description Ec(t2e) Tmax- εcbs(tcrit)- σ(tcrit) calc σ(tcrit) exp
(MPa) To(oK) εcbs(t2) (MPa) (MPa)
(10-6)
w/b=0,38, Cem I SR, 33253 41,8 21 4,26 4,00(1,06)
0% SCM, R=0,5, t2=3d
tcrit=9,25d, To= 20oC
w/b=0,38, Cem II, incl 31577 36,2 22 3,54 3,75(0,94)
FA, R=0,5, t2=3d
tcrit=9,25d, To= 20oC
w/b=0,38, Cem I, R=0,5, 32077 39,5 23 3,91 3,50(1,12)
t2=3d tcrit=9,25d,
To=20oC
w/b=0,48, Cem I SR, 29510 35,2 18 3,19 2,80(1,14)
R=0,5, t2=3,2d
tcrit=9,25d, To= 20oC
w/b=0,48, Cem II incl 27519 28,8 13 2,42 2,55(0,95)
FA, R=0,5, t2=3d
tcrit=9,25d, To= 20oC
Table 7. Verification results,3rd approach: t2, tcrit,Tc,max, To , and E-modulus from the
experiment. Thermal dilation and basic shrinkage also from the experiment.
Test description Ec(t2e) Tmax- εcbs(tcrit)- σ(tcrit) calc σ(tcrit) exp
(MPa) To(oK) εcbs(t2) (MPa) (MPa)
(10-6)
w/b=0,38, Cem I SR, 34176 41,8 -1,6 4,05 4,00(1,01)
0% SCM, R=0,5, t2=3d
tcrit=9,25d, To= 20oC,
αcth=9,8ꞏ10-6
w/b=0,38, Cem II, incl 32633 36,2 7,1 3,37 3,75(0,90)
FA, R=0,5, t2=3d
tcrit=9,25d, To= 20oC,
αcth=9,6ꞏ10-6
w/b=0,38, Cem I, R=0,5, 32835 39,5 6,0 3,68 3,50(1,05)
t2=3d tcrit=9,25d,
To=20oC, αT=9,6ꞏ10-6
w/b=0,48, Cem I SR, 30844 35,2 -14 2,76 2,80(0,99)
R=0,5, t2=3,2d
tcrit=9,25d, To= 20oC,
αcth=9,2ꞏ10-6
w/b=0,48, Cem II incl 28902 28,8 -11 2,21 2,55(0,87)
FA, R=0,5, t2=3d
tcrit=9,25d, To= 20oC,
αcth=9,6ꞏ10-6
Figure 7. Calculated versus experimental results applying formula (D.4), 1st approach.
Statistics: μθ=1,07 , CoV=0,007.
Figure 8. Calculated versus experimental results applying formula (D.4), 2nd approach.
Statistics: μθ=1,04, CoV=0,007
Figure 9. Calculated versus experimental results applying formula (D.4), 3rd approach.
Statistics: μθ=0,96, CoV=0,005
𝑤 , 𝑘𝑤 ∙ 𝑘1/𝑟 ∙ 𝑆 , ∙ 𝜀 𝜀 (9.8)
wk,cal is the characteristic surface crack width, kw the conversion factor from mean to
characteristic values, Srm,cal the calculated mean crack spacing, while εsm and εcm are
the mean strains in reinforcement and concrete, respectively, and their
difference 𝜀 𝜀 represents the crack strain. The parameter k1/r allows for out
of plane curvature, for uniaxial tension as typically encountered with thermal and
shrinkage cracking this factor may be taken as 1,0.
For elements subjected to imposed strains which are restrained at the edges
(typically walls on foundation slabs), the crack strain (εsm - εcm) in Formula (9.8) may
be replaced by:
𝑓 ,
𝜀 𝜀 𝑅𝑎𝑥 𝜀 𝑘 ∙ (9.13)
𝐸
where kt is a coefficient dependent on the nature and duration of the load, fct,eff the
mean value of the tensile strength of the concrete at the time when the cracks are
expected to occur and Ecm the modulus of elasticity. Raxεfree in Formula (9.13) can be
taken as:
The parameters are described above in connection with Formulas (D.4) and (D.5).
Elements restrained at the ends can according to 9.2.3(3) be considered in the crack
formation stage and the crack strain in Formula (9.8) may be replaced by Formulae
(9.11):
(9.11)
Relevant background for Formula (D.6) and (D.7) is the UK CIRIA 660 regulations [2]
and the French report CEOS.fr [3] which both are using similar approaches.
An interesting comparison is done for the typical wall on foundation slab system,
Figure 8 [19], ie an investigation of a section one wall-thickness above the
foundation. It is seen that the agreement between the various approaches
investigated is reasonably good. It should be noted that the work was finished in
2018, and therefore based on the literature available at that time. An interesting
publication for verification of the crack width calculation combining 9.2.3 and Annex
D is the comprehensive investigation of wall-on-slab-structures presented in [20].
Furthermore, is the topic highly addressed in the recent RILEM committee TC 287-
CCS Early age and long-term crack width analysis in RC Structures.
a)
b)
References References
[1] FprEN 1992-1-1 2023: Eurocode 2: Design of concrete structures – Part 1-1:
General rules, rules for buildings, bridges and civil engineering structures,
Annex D.
[2] CIRIA 660 (P.B. Bamforth et al) Early-age thermal crack control in concrete.
London 2007 (& draft of 2015-version).
[3] CEOS.fr Cracking and shrinkage. Behaviour and evaluation of special
structures. Recommendations for the control of cracking phenomena
(January 2015).
[4] Azenha, M. et al., Recommendations of RILEM TC 287-CCS: Thermo-chemo-
mechanical modelling of massive concrete structures towards cracking risk
assessment. Materials and Structures (RILEM), Accepted for publication per
April 2021.
[5] Laube, M., Werkstoffmodell zur berechnung von temperaturspannungen in
massigen betonbauteilen im junge alter, Technical University of
Braunschweig, Doctoral thesis, Doctoral thesis, 1990. (In German)
[6] Freiesleben Hansen, P., Hardening technology 2, Decrement method, BKF-
centralen 1978 (In danish)
[7] Kanstad, T.; Hammer, T.A.; Bjøntegaard, Ø.; Sellevold, E. J. Mechanical
Properties of Young Concrete: Part I - Experimental Results related to Test
Methods and Temperature Effects. Materials and Structures 2003 ;Volume
36. p. 218-225
[16] Kanstad, T. Verification of three different calculation methods for early age
concrete. In: Crack Risk Assessment of Hardening Concrete Structures. The
Nordic Concrete Federation 2006 ISBN 82-91341-97-4. p. 101-110
[17] Larson, M.; Estimation of crack risk in early age concrete, simplified methods
for practical use, Licentiate thesis, Luleå technical university, Sweden
2000:10, ISSN 1402-1757.
[18] Klausen, Anja Birgitta Estensen; Kanstad, Terje; Bjøntegaard, Øyvind.
Hardening Concrete Exposed to Realistic Curing Temperature Regimes and
Restraint Conditions: Advanced Testing and Design Methodology. Advances
in Materials Science and Engineering 2019 ;Volume 2019. p. –
[19] Klausen, Anja; Early age crack assessment: codes, guidelines and calculation
methods. DaCS Project memo. SINTEF Building and infrastructure.
Trondgheim, Norway 2018.
[20] Jedrzejewska, A., Kanavaris, F., Zych, M., Schlicke, D., Azenha,M.;
Experiences on early age cracking of wall-on-slab concrete structures. In
Structures, Volume 27, October 2020, DOI: 10.1016/j.istruc.2020.06.013
28.11.2022
For citations, copying and referencing see Note on Front page! page 585
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F Background to FprEN 1992-1-1:2023
Introduction
This document has been prepared in order to provide background information related to the FprEN 1992-1-1 Annex F
- “Safety formats for non-linear analysis”. In the last decades, the non-linear analyses (NLAs) have increasingly
become the most common instrument able to simulate the mechanical response of structural systems, such as
reinforced concrete (RC) members, from serviceability to the ultimate limit state (ULS). Concerning the results of such
complex calculations, they need to be properly processed in order to satisfy the target reliability levels required by this
standard (i.e., EN 1990 [1]; FprEN 1992-1-1 [2] – Annex A). To this aim, several methodologies (i.e., safety formats)
have been introduced by scientific literature [3]-[7] and codes as [8]-[9]. All the different safety formats provide a
framework devoted to assess the design value of the structural resistance Rd including the relevant sources of
uncertainties: the aleatory and the epistemic ones [10]-[14].
FprEN 1992-1-1 - The FprEN 1992-1-1 - Annex F - “Safety formats for non-linear analysis” aims to provide a standard to perform
Annex F validation (with reference to general and project specific attempts) and definition (e.g., characterization of the loading
process, selection of the solution strategy and of the modelling hypotheses) of non-linear numerical (NL) models of RC
F.1 Use of this annex structural systems/members and to establish the methods (i.e., safety formats) useful to derive design values of
F.2 Scope and field of structural resistance Rd for ULS verification according to EN 1990 [1].
application
Advantages:
- Analysis is performed with properties and material behaviour closer to reality, therefore the distribution of
internal forces and deformations is more realistic;
- It enables an assessment of structures and structural elements that cannot be easily calculated with the
standard analytical formulas as these are typically valid only for beams, columns or slabs with simple geometrical
shapes without openings;
- It can provide a useful insight in the real structural behaviour and the actual failure mode;
- In some cases when deeper knowledge of the structure and its properties is available, and very well validated
model is used, it can provide higher load carrying capacity then standard methods. This is especially useful for
the assessment of existing structures.
Disadvantages:
- Nonlinear analysis is more time consuming in model preparation, analysis time as well as post-processing;
- Principle of superimposition does not apply to actions and their effects, so the analysis must be performed for each
combination of the actions separately;
compatibility and constitutive laws always lead to diverse solutions for the specific structural problem. Then, the
multiplicity of choices that may be performed defining NLN model leads to a further degree of uncertainty having
epistemic nature [10]-[17]. This source of uncertainty can be efficiently represented by the model uncertainty random
variable ϑ, which can be probabilistically characterized comparing experimental and related numerical results. This
probabilistic calibration should be carried out having care to reduce as much as possible the influence of other
epistemic sources of uncertainty as the statistical (i.e., limited sample of observation) and experimental (i.e., deviation
of tests set configuration and/or of measurement devices) ones, that inevitably may affect the probabilistic calibration
of the model uncertainty random variable ϑ.
Next, the background to safety formats implemented in FprEN 1992-1-1 Annex F aimed to account for the different
sources of uncertainty in ULS verifications of RC structures/members is reported.
Figure BG-F.1: Local and global approaches for safety verification of RC structures (Castaldo et al. 2019 [6]).
For instance, in case the ULS verification is performed using refined NLAs, the global capability of the reinforced
F.3 General concrete structures/members to redistribute stress fields and internal forces can not be neglected. In this context, the
use of non-linear numerical models requires the adoption of a global approach for evaluation of the structural safety,
comparing directly the actions combined at ULS according to the selected load case Fd and the associated design
value of structural resistance Rd. This approach is denoted as global resistance format (GRF) in line to [8]. The main
differences between the two approaches are explained in Figure BG-F.1.
In line with the global resistance format [8] and EN 1990 [1], the ULS verification of an RC structural members/system
can be exploited as follow:
Clause (5) Fd R d (BG-F.1)
where the design value of the actions Fd can be evaluated in line with the EN 1990 [1] according to the proper
combination, while, the design value of structural resistance Rd can be evaluated through NLAs with the application of
safety formats (which will be introduced next in this document) according to the following general expression:
R X rep ; a rep
Rd (BG-F.2)
R* R d
In the formula (BG-F.2), the term R{∙} denotes the structural resistance estimated by means of NLA using the
representative values Xrep for material and arep for geometric properties according to the adopted safety format, as
described next. The level of structural reliability adopted for ULS verification is involved by means of two different
partial safety factors:
- the global resistance factor γ*R, which is a partial safety factor that takes into account, at the level of global
structural response, of the influence of the aleatory uncertainties related to the material properties and
geometry. This partial safety factor can be determined in compliance with a specific target level of reliability
according to the methodology described by the selected safety format;
- the model uncertainty factor γRd, which is a partial safety factor that takes into account the uncertainty (of
epistemic nature) related to simplifications, assumptions and choices performed to define of the non-linear
numerical model. This partial safety factor is horizontal and independent from the adopted safety format.
The software used in non-linear verification of the ultimate limit state should be validated by comparison between
Clause (3) numerical and experimental or benchmark results. Similarly, the choices made with respect to the specific numerical
model should be tested by sensitivity analysis.
1) general validation attempts:
- basic material tests;
- structural tests with the characterization of relevant failure modes;
2) project specific attempts:
- mesh sensitivity tests;
- solution method tests.
In general, these issues may be solved partially by software provider and partially by the analyst which is directly
involved in the specific project.
The structural resistance R{∙} that corresponds to the achievement of the ultimate limit state (i.e., structural failure)
shall be identified by the analyst in concomitance of actions level where lack of convergence of numerical procedure
occurs. The analysts should interpret the results paying attention to distinguish and recognize potential numerical
failure and/or physical failure and to characterize the relevant failure mode (see F.5 FprEN 1992-1-1 [2] – Annex F).
The loading process implemented by a non-linear simulation in order to evaluate the structural resistance R{∙} against
the actions shall be defined according to engineering judgement consistently with the considered combination at
Clause (4) ultimate limit state. In particular, the outcomes of the NLA are significantly affected by the loading procedure that, with
particular care to staged construction methods, should be identified accounting for the actual loading sequence and
even including time dependent effects if relevant. In particular, the structural resistance R{∙} may be evaluated
increasing the actions from their initial values (e.g., characteristic values) by incremental steps, such that the
associated design values in the adopted combination are reached in the same step. The incremental process of the
actions should be continued beyond their design values with the same law until structural failure is reached. The
structural resistance R{∙} corresponds to the values of the actions which lead to structural failure. With the adoption of
this approach, the estimated design value of of structural resistance Rd can be directly compared to design value of the
axions Fd with reference to the specific combination at ULS.
The safety formats based on the above described approach, which are introduced by FprEN 1992-1-1 Annex F, are
Clause (6) the following:
i. Partial factor method (PFM);
ii. Global factor method (GFM);
iii. Full probabilistic method (FPM).
In particular cases, where higher degree of refinement for estimation of structural reliability for ULS verification is
required, the (iii.) Full Probabilistic Method (FPM) can be adopted in compliance to EN 1990 [1] and FprEN 1992-1-1
[2] – Annex A .
In the following, the principal features of the safety formats (i., ii., iii.) for NLAs of RC structural systems/members
reported in FprEN 1992-1-1 Annex F are described.
The Partial Factor Method (PFM) has been introduced by [8]. The design value of structural resistance Rd is obtained
F.4 Partial factor method by means of one NLA according to formula (BG-F.3):
(PFM)
R X d ; ad (BG-F.3)
Rd
Clause (1) Rd
where R{Xd; ad) represents the structural resistance estimated by means of a non-linear numerical analysis using the
design values of the material (Xd) and geometrical (ad) properties; γRd is the partial safety factor which accounts for
model uncertainty related to global NLAs.
With the PFM, the influence of the aleatory uncertainties is accounted for by Xd according to EN 1990 [1]. The partial
safety factors to be used for the evaluation of Xd should consider the statistical variability of the material property and
the geometric uncertainty but should not be inclusive of model uncertainty (which is accounted for trough the model
uncertainty factor γRd related to global resistance estimated by NLA). For instance, the method of EN 1990 [1] should
be implemented according to specifications of Table A.1 line (e) of FprEN 1992-1-1 [2] - Annex A and FprEN 1992-1-1
[2] - Annex A in general. The appropriate target reliability index βtgt can be adopted in the evaluation of the mentioned
above partial safety factors always in compliance to EN 1990 [1] and FprEN 1992-1-1 [2] - Annex A. In case the
geometric uncertainty is not accounted for within the evaluation of the partial safety factors associated to material
properties (as previously mentioned), the design value of geometrical properties (ad) should be used according to EN
1990 [1]. Alternatively, the NLN model should be defined using the nominal values for geometric properties (anom)
(prEN 1990:2020 [1]).
In case of adoption of simplified approach for non-linear analysis according to following assumptions:
Clause (2)
- structural resistance is determined based on numerical model established in a way that the influence of the
concrete tensile strength is neglected;
- and the model follows the assumptions and verification procedures of Section 8 and/or Annex G of FprEN
1992-1-1 [2];
the model uncertainty can be assumed to be equal to the one included in calculation of fcd and fyd according to Section
5 of FprEN 1992-1-1 [2]. In this circumstance, formula (BG-F.3) may be replaced by:
R d R f cd , f yd ; a nom
where fcd and fyd are derived from Section 5 of FprEN 1992-1-1 [2] and anom is the nominal value of the geometric
property according to EN 1990 [1].
The Global Factor Method (GFM) is based on the evaluation of the design value of structural resistance Rd by means
F.5 Global factor method of global partial safety factors (i.e., γ*R and γRd) applied to mean value of structural resistance Rm{X; a} which is
(GFM)
estimated throughout NLAs. In particular, the principles of this method descends from global resistance safety formats
F.5.1 General as the “Method of Estimation of Coefficient of Variation” (ECoV) and “Global Resistance Factor Method” (GRF)
Clause (1) reported by [8]-[9], which has been furtherly investigated by [6]-[7], [21].
The Global Factor Method (GFM) allow to estimate the design value Rd assuming that the structural resistance R is
lognormally distributed. The statistical parameters (i.e., mean value Rm and coefficient of variation V*R) should be
evaluated trough probabilistic analysis including materials (X) and geometrical (a) uncertainties using a probabilistic
model for basic variables compliant with EN 1990 [1] and FprEN 1992-1-1 [2] - Annex A. Although the assumption of
lognormal probabilistic distribution turns out to be a simplification, it is widely adopted to describe resistance random
variables and its efficiency is demonstrated by several applications [5]-[7]. In case the lognormal probabilistic model is
not able to represent the structural resistance R the Full Probabilistic Method (FPM) can be adopted to perform
reliability-based verification of the specific ULS.
The Global Factor Method (GFM) can be used to determine the design value of structural resistance Rd according to
the following expression in line to assumption of lognormal distribution:
R m X ; a (BG-F.4)
Rd
R* R d
The coefficient of variation VR,M reflect the influence of the statistical variability of material properties on
variability of structural resistance. According to the ECoV method [3],[8] and in line to the assumption of
lognormal probabilistic model for structural resistance R, the coefficient of variation can be estimated with the
following simplified approach:
1 R X m ; a nom (BG-F.6)
V R ,M ln
1 .6 5 R X k ; a nom
where R{Xk; anom} is the structural resistance estimated by means of NLA using characteristic values of
material properties (Xk) in line to FprEN 1992-1-1 [2].
VR,M [-]
- T-Beam: RC beam with “T” cross section
designed in compliance to EN1992-1-1
SM SH
T-Beam
SL CLX
Reinforcement yielding
Figure BG-F.2: Coeffiecient of variation related to uncertainties associated to material properties VR,M (Castaldo et
al. 2019 [6]) estimated by means probabilistic analysis using NLA (non-linear finite element
analsysis).
The results from probabilistic analysis of ultimate response of RC beams performed by [6] including material
uncertainties only, highlighted that, in case of missing explicit estimation of VR,M, such value can be set equal
to the upper bound of 0.15 assuming that failure is fully governed by concrete cylinder compressive strength.
This assumption is valid until the coefficient of variation of concrete cylinder compressive strength is VC≤0.15
(as generally assumed by [8],[18]). In case of higher values for VC, reference to the FprEN 1992-1-1 [2]-
Annex A can be performed and the value of VR,M should be estimated appropriately. The coefficient of
variation VR,G account for the contribution of geometric uncertainty to statistical variability of structural
resistance. The value of VR,G can be assumed according to the FprEN 1992-1-1 [2] – Annex A or, as further
simplification on safe side, can be set equal to 0.05 [8],[22] until the governing size of structural component is
bigger than 200 mm.
F.6Full Probabilistic The Full Probabilistic Method (FPM is in compliance to EN 1990 [1] - Annex C. This method allows to quantify directly
Method (FPM)
Clauses (1), (2), (3) the level of structural reliability with reference to considered ULS. In order to fulfill the safety verification, the achieved
level of reliability should be higher than the one associated to the target value EN 1990 [1], FprEN 1992-1-1 [2] –
Annex A. Additional information related to probabilistic methods for structural reliability analysis can be found in [25].
Probabilistic models for model uncertainties (both from actions and resistance sides), actions, material and geometric
properties should be defined in line to EN 1990 [1] and FprEN 1992-1-1 [2] – Annex A. With particular reference to the
probabilistic model for uncertainty associated to the resistance model established using NLA (i.e., model uncertainty),
it should be defined according to FprEN 1992-1-1 [2] – Annex F and as described in the following. More details about
probabilistic models for concrete structures can be acknowledged in [8] and [26].
The epistemic uncertainty in NLAs is associated to “missing” knowledge, hypotheses and simplifications related to
Clauses (1), (3), (4), (6) constitutive laws, kinematic compatibility and equilibrium of forces and to assumptions concerning auxiliary non-
physical variables or individual choices [11]-[17]. All the mentioned above aspects characterize the selection of a
specific modelling hypothesis [12]-[14], also denoted in literature as “solution strategy” [17]. Moreover, the level of
epistemic uncertainty can be influenced by the nature of the governing failure mode (i.e., brittle, ductile) [16]-[17].
With reference to [11]-[17], the epistemic uncertainty in resistance models may be exhaustively represented by the
resistance model uncertainty random variable denoted as:
R exp , i X , Y (BG-F.7)
i
R num , i X
The formula (BG-F.5) relates the structural resistance estimated from an experimental test Rexp,i(X,Y) to the structural
resistance estimated through numerical simulation (i.e., NLA) Rnum,i(X). X is a vector of basic variables included into
the numerical resistance model, while Y is a vector of variables that may affect the resistance mechanism but are
neglected in the model and, their unknown influence, can be indirectly incorporated including the model uncertainty
random variable ϑ within the reliability analysis.
An in-depth probabilistic calibration of the resistance model uncertainty random variable and of the related partial
safety factor γRd needs to be addressed in relation to the selected set of modelling hypotheses and the target reliability
differentiation according to EN 1990 [1] and FprEN 1992-1-1 [2] - Annex A . In line with the approaches of [12]-[17],
the probabilistic calibration of the resistance model uncertainty random variable can be performed in line to the
following steps:
1. Selection of the benchmark experimental tests: the selection of the benchmark set of experimental results
should be performed accounting for geometries, material properties and failure mechanisms inherent to the
investigated problem. In general, all the information about materials and geometry to define the non-linear
numerical models should be derived by the benchmark experimental test. In case of missing information
default values can be assumed according to FprEN 1992-1-1 [2];
2. Definition of the non-linear numerical models: the numerical models defined in order to reproduce by NLAs
of the benchmark experimental tests should be established adopting a specific set of modelling hypotheses
and assumptions. The modelling uncertainty so far estimated will be valid and representative only for the
considered sets of modelling hypotheses;
3. Quantification of the model uncertainty: the model uncertainty should be estimated for each benchmark
experimental test according to formula (BG-F.7);
4. Statistical analysis of model uncertainty: the outcomes of observed model uncertainty should be
appropriately processed from statistical point of view with the aim to determine the best estimate of
statistical parameters (e.g., mean value μϑ and coefficient of variation Vϑ) and the most likely probabilistic
distribution for model uncertainty random variable ϑ. According to [12],[17], in the most of the cases, the
preferred probabilistic distribution able to describe the model uncertainty random variable is the lognormal
one. The methods for statistical analysis should be able to account for the influence of the limited number of
observed samples ϑi on the estimation of mean value μϑ and coefficient of variation Vϑ. Techniques as the
methods of moments, maximum likelihood method (ML) [23] and Bayesian approaches [24] can be adopted.
Once the probabilistic calibration of the model uncertainty random variable ϑ has been performed, it can be possible to
derive the related partial safety factor as follows:
1 (BG-F.8)
Rd
F 1 R tgt
where F(∙) is the likely cumulative probabilistic distribution function able to describe resistance model uncertainty
random variable; Φ is the standard normal cumulative probabilistic distribution function; αR is the first-order-reliability-
method (FORM) sensitivity factor which is equal to 0.32 in the assumption of non-dominant variable in line to [8]; βtgt is
the target reliability index that should be adopted in compliance to EN 1990 [1] and FprEN 1992-1-1 [2] - Annex A
depending from the considered ULS.
In case the adopted probabilistic distribution is the lognormal one (as it happen in the large part of the cases [12]-[17]),
the formula (BG-F.8) can be rewritten as:
e xp R tgtV (BG-F.9)
Rd
In general, the assumption of non-dominant variable for model uncertainty [8] turns out to be verified until the following
expression is satisfied [27]:
V R* V (BG-F.10)
where V*R is the coefficient of variation of structural resistance inclusive of materials and geometric uncertainties (see
F.3 of FprEN 1992-1-1 [2] - Annex F and related explanation in this document) while, Vϑ is the coefficient of variation
related to model uncertainty.
In cases where the reliability analysis is dominated by the model uncertainty (i.e., the formula BG-F.10 is not fulfilled),
the non-dominant variables becomes the materials and geometric ones (i.e., aleatory). In this case the formula BG-F.9
(and the related one in EN 1992-1-1 [2] - Annex F) should be applied adopting first-order-reliability-method (FORM)
sensitivity factor αR equal to 0.8. On the opposite, the methods described by F.2 and F.3 of EN 1992-1-1 [2] - Annex F
should be applied adopting αR equal to 0.32 to account for the uncertainties related to materials and geometry within
the calculation of related partial safety factors.
Clause (2)
In the FprEN 1992-1-1 [2] - Annex F, fixed values of γRd are also provided in order to cover cases where the
probabilistic calibration of model uncertainty is not carried out by the analysis and/or software produces. These fixed
values are inclusive of the influence of statistical uncertainty and are consistent with the assumption of non-dominant
variable for model uncertainty (i.e., αR=0.32). In case the model uncertainty results to be the dominant variable, these
fixed values are not valid and the probabilistic calibration according to FprEN 1992-1-1 [2] - Annex F and as it is
previously explained in this document have to be performed. In the literature, several investigations reports studies
related model uncertainty quantification (with particular reference to non-linear finite elements analysis) [12]-[17] that
justifies the suggested values assumed on the safe side by FprEN 1992-1-1 [2] - Annex F with reference to 50 years
reference life reliability index βtgt equal to 3.8.
The investigation of Castaldo et al. 2018 [12] considered a selection of 25 structural members from several literature
references including deep beams, shear panels and walls. All the structural members have been realized with
statically determined scheme and tested up to failure with a monotonic incremental loading process. The experimental
tests have been reproduced by non-linear numerical simulations adopting 9 different modelling hypotheses
distinguishing between the software platform and concrete tensile response. A total number of 225 NL simulations has
been performed to estimate the resistance model uncertainty random variable. The results of the investigation are
reported in Table BG-F.2.
The work of Castaldo et al. 2020 [13] propose the investigation related to the epistemic uncertainty in non-linear
numerical analyses using the finite element method in case of cyclic loading process. In particular, 17 shear walls
tested by several Authors under different cyclic loading histories and statically determined static scheme has been
considered. These experimental tests cases have been reproduced by means 18 different modelling hypotheses
differentiating between the adopted software, concrete tensile behavior and shear stiffness of concrete after cracking.
A total number of 306 NLN simulation has been collected in order to characterize the resistance model uncertainty
random variable, with results listed in Table BG-F.2.
Cervenka et al. 2018 [16] reports the investigation of 33 RC structural elements which are inclusive of slabs and
beams which presented both shear and bending failure modes. The experimental tests on these members has been
reproduced by means NL numerical simulations adopting single modelling hypothesis and using monotonic
incremental loading process. The Table BG-F.2 reports the main results in terms of characterization of resistance
model uncertainty random variable distinguishing between bending, shear and punching failure modes.
Table BG-F.2: Characterization of model uncertainty random variable from different literature references with the
related model uncertainty partial safety factor γRd.
Model uncertainty ϑ Partial
Type of RC Coefficient safety
Reference Type of NLA Failure mode Probabilistic Mean value
members of variation factor
distribution μϑ γRd*1
Vϑ
2D finite Different failure
elements modes due to
Castaldo Deep beams,
(plane stress); both concrete
et al. 2018 walls, shear Lognormal 1.01 0.12 1.15
incremental crushing and
[12] panels
monotonic reinforcement
loading yielding
Different failure
2D finite modes due to
Castaldo
elements both concrete
et al. 2020 Shear walls Lognormal 0.88 0.13 1.35
(plane stress); crushing and
[13]
cyclic loading reinforcement
yielding
Mainly ductile
(failure
characterized
1.04 0.05 1.02
extensive
reinforcements
3D finite yielding)
elements
Engen et Beams, deep Mainly brittle
(brick);
al. 2017 beams, walls, (failure Lognormal
incremental
[17] frames characterized by
monotonic concrete crushing
loading 1.14 0.12 1.02
without or with
limited amount of
reinforcements
yielding)
*1 the values of γRd are derived considering 50 years reference life reliability index βtgt= 3.8 and αR= 0.32 (non-dominat).
*2 the range bounds the influence of significant/limited level of experimental uncertainty within the tests sets considered for investigation
of model uncertainty [14].
The research of Engen et al. 2017 [17] reports the results of non-linear numerical simulations performed in order to
reproduce 38 RC members experimentally tested under monotonic loading process by several authors. The structural
members were inclusive of beams, deep beams, frames and walls. Also in this case a single modelling hypothesis has
been adopted to realize the NL numerical models of the considered members. The outcomes of the investigation
leaded to the characterization of the resistance model uncertainty random variable distinguishing between ductile and
brittle failure mechanisms. Table BG-F.2 reports the main results of this last investigation.
Finally, the work of Gino et al., 2021 [14] rely to the characterization of the epistemic uncertainty related to NLAs of
slender RC members. A total number of 40 experimental outcomes of tests on reinforced concrete columns having
slenderness ratio between 15 and 275 and different geometries in line with [22] limitations have been considered. A
discussion related to the influence of the experimental uncertainty on estimation of model uncertainty is proposed and
statistical values/model uncertainty factors are derived performing different assumptions. The main results of the
investigation are reported in Table BG-F.2.
Clauses (6), (7)
The investigation of Castaldo et. al, [6] raised the issue of the influence of the probabilistic distribution of aleatory
uncertainties (i.e., materials) on prediction of failure mechanism by means numerical modelling and related
repercussions on the application of safety formats as GFM and PFM (i.e., methods that not require an extensive
probabilistic investigation of structural resistance). In particular, the results of [6] highlights that the possible presence
of concavity in the response surface of structural resistance in proximity of the design point may lead to unsafe
conclusions when the GFM and PFM are used. This particular source of uncertainty is not covered by the model
uncertainty factor γRd alone because it can manifest only in cases where the design value of structural resistance is
evaluated without full probabilistic analysis and concern the interaction between aleatory and epistemic (i.e., model)
uncertainties.
For instance, previously to perform the specific ULS verification, the sensitivity of structural response to probabilistic
distribution of material properties should be investigated [6],[28]. In line to the results of [6] and subsequent discussion
and elaboration in CEN/TC250/ SC2/ WG1/TG6, the preliminary procedure described in FprEN 1992-1-1 [2] - Annex F
have been defined in order to check the shape of response surface of structural resistance in proximity of the design
point.
In case the following relationships between the structural resistances evaluated with different values of material
properties are verified, the numerical model can be considered to be not sensible to probabilistic distribution of
material properties:
where fcm and fym are the mean values of concrete cylinder compressive strength and reinforcement yielding strength,
respectively, while fcd and fyd are the related design values determined according to FprEN 1992-1-1 [2].
In case the conditions of (BG-F.10) turn out to be not fulfilled, it means that in the proximity of the design point may
exist possible weak failure mechanisms that give rise to local modification of concavity of the response surface. In this
case, an additional uncertainty should be included within the estimation of the model uncertainty factor γRd that, in line
with [6], can be increased of 15% as safe assumption.
References
[1] EN1990 Eurocode – Basis of structural design. CEN . Brussels
[2] FprEN1992-1-1 Eurocode 2 – Design of concrete structures. CEN 2013, Brussels.
[3] Cervenka V., Global safety formats in fib Model Code 2010 for design of concrete structures, Proceedings of
the 11th Probabilistic Workshop, Brno, 2013.
[4] Cervenka V., Reliability–based non-linear analysis according to fib Model Code 2010, Structures Concrete,
Journalof the fib, vol. 14, March 2013, ISSN1464-4177, p.p.19-28, 2011.
[5] Shlune H, Gylltoft K, Plos M. Safety format for non-linear analysis of concrete structures. Magazine of
Concrete Research 2012; 64(7): 563-574.
[6] P. Castaldo, D. Gino, G. Mancini (2019): Safety formats for non-linear analysis of reinforced concrete
structures: discussion, comparison and proposals, Engineering Structures, 193, 136-153,
https://doi.org/10.1016/j.engstruct.2018.09.041.
[7] Allaix DL, Carbone VI, Mancini G. Global safety format for non-linear analysis of reinforced concrete
structures. Structural Concrete 2013; 14(1): 29-42.
[8] fib Model Code for Concrete Structures 2010. Fib 2013. Lausanne.
[9] EN 1992-2: Eurocode 2 – Design of concrete structures. Part 2: Concrete bridges – Design and detailing
rules. CEN 2005. Brussels.
[10] Kiureghian AD, Ditlevsen O. Aleatory or epistemic? Does it matter?. Structural Safety, 31:105-112, 2009.
[11] Holický M, Retief JV, Sikora M. Assessment of model uncertainties for structural resistance. Probabilistic
Engineering Mechanics 2016; 45: 188-197.
[12] P. Castaldo, D. Gino, G. Bertagnoli, G. Mancini (2018): Partial safety factor for resistance model uncertainties
in 2D non-linear analysis of reinforced concrete structures, Engineering Structures, 176, 746-762.
https://doi.org/10.1016/j.engstruct.2018.09.041.
[13] P. Castaldo, D. Gino, G. Bertagnoli, G. Mancini (2020): Resistance model uncertainty in non-linear finite
element analyses of cyclically loaded reinforced concrete systems, Engineering Structures, 211(2020),
110496, https://doi.org/10.1016/j.engstruct.2020.110496
[14] Gino D, Castaldo P, Giordano L, Mancini G. Model uncertainty in nonlinear numerical analyses of slender
reinforced concrete members. Structural Concrete. 2021;1–26. https://doi.org/10.1002/suco.202000600
[15] P. Castaldo, D. Gino, G. C. Marano, G. Mancini (2022): Aleatory uncertainties with global resistance safety
factors for non-linear analyses of slender reinforced concrete columns, Engineering Structures, 2022; 255;
113920. https://doi.org/10.1016/j.engstruct.2022.113920
[16] Cervenka V, Cervenka J and Kadlek L (2018) Model uncertainties in numerical simulations of reinforced
concrete structures, Structural Concrete, 2018;19:2004–2016.
[17] Engen M, Hendriks MAN, Köhler J, Øverli JA, Åldtstedt E. A quantification of modelling uncertainty for non-
linear finite element analysis of large concrete structures. Structural Safety 2017; 64: 1-8.
[18] JCSS. JCSS Probabilistic Model Code. 2001.
[19] fib Bulletin N°80. Partial factor methods for existing concrete structures, Lausanne, Switzerland; 2016.
[20] fib Bulletin N°45. Practitioner’s guide to finite element modelling of reinforced concrete structures – State of
the art report. Lausanne; 2008.
[21] Cervenka V., Reliability –based non-linear analysis according to fib Model Code 2010, Structures Concrete,
Journalof the fib, vol. 14, March 2013, ISSN1464-4177, p.p.19-28, 2011.
[22] EN 1992-1-1: Eurocode 2 – Design of concrete structures. Part 1-1: general rules and rules for buildings. CEN
2014. Brussels.
[23] Faber, Michael Havbro Statistics and Probability Theory, Springer, 2012.
[24] Gelman A, Carlin JB, Stern HS, Dunson DB, Vehtari A, Rubin DB. Bayesian data analysis. 3rd ed. CRC Press;
2014.
[25] ISO 2394. General Principles on Reliability for Structures; ISO 2394:2015(E); International Organization for
Standartization (ISO): Geneva, Switzerland, 2015; p. 111.
[26] JCSS. JCSS Probabilistic Model Code. 2001.
[27] fib Bulletin N°80. Partial factor methods for existing concrete structures, Lausanne, Switzerland; 2016.
[28] Yu, Q., Valeri, P., Ruiz, M.F. and Muttoni, A. 2021. A consistent safety format and design approach for brittle
systems and application to textile reinforced concrete structures. Engineering Structures, 249: 113306,
https://doi.org/10.1016/j.engstruct.2021.113306.
28.02.2021
J. Cervenka
For citations, copying and referencing see Note on Front page! page 595
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
Introduction
This document has been prepared in order to provide additional background information to the Annex F. It
extends the information provided in the background information document [12]. It summarizes the safety
formats for non-linear analysis as they are proposed in EC2-Annex F. It demonstrates their application and
compares the calculated design values with analytical formulas and experiments. This annex describes three
possible safety formats:
F.4 Partial factor method
F.5 Global factor method
F.6 Full probabilistic method
This documents demonstrates the application of these safety formats to a selected set of benchmark problems
summarizing the most typical elements and failure modes occurring in the design of reinforced concrete
structures.
The first section describes the appropriate choice of input parameters for the application of the safety formats
F.4. to F.6.
The second section describes the benchmark problems used in this study. This section includes also an
important discussion on the choice of model uncertainty partial safety factor.
The third section summarizes and compares the obtained results and some concluding remarks.
R X d ; a
Rd (1)
Rd
where:
is the design value of the material property calculating adopting partial safety factors
Xd
according to Table A.1 line (f) accounting for materials and geometric uncertainties, but
excluding model uncertainties, which are considered separately by Rd . In general case
the design value is calculated as:
Rd is the partial safety factor, which accounts for the model uncertainty according to F.5;
tgt Is the target value for the reliability index according to Table A.3;
J. Cervenka
For citations, copying and referencing see Note on Front page! page 596
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
n
VRM V I
I
2 ;
RM Is the bias factor for all uncertainties, which may be calculated as a product of all bias
n
factors: RM I .
I
For typical situations in accordance to Annex A, the material safety factors for concrete and steel are listed in
Table 1
Table 1: Standard values for material factors for the safety format F.4
persistent and transient design situations
S C
1.2 1.54 (1.46)1
Rm X ; a
Rd (2)
R* Rd
where:
Rm is the mean value of the structural resistance based on the numerical simulation and
shall be estimated by means of a probabilistic analysis including uncertainties related
to material and geometric properties;
R* is the global resistance factor for the uncertainties of material properties and geometry
according to F3.2;
Rd is the partial safety factor, which accounts for the model uncertainty according to F.5;
where:
1 R X m ; anom 1 R X m* ; anom
(4)
VR* ln ln
1.65 R X k ; ak 1.65 R X k* ; anom
In (4) R X m ; anom represents an estimate of the structural strength calculated by a nonlinear analysis using
average material parameters and nominal geometric properties. R X k ; ak is the analogic estimate of the
characteristic strength that is obtained by a nonlinear analysis using the characteristic values of the material
and geometric parameters. This approached was originally proposed in [8] and [9]. It became part of the
nonlinear safety formats proposed in the fib model code 2010 [17] along with the partial factor method F.4. In
practical cases it would be however rather inconvenient to create a model taking into account the characteristic
values of its dimensions or reinforcement location. It is therefore proposed to keep geometric parameters to
their nominal values and rather define an auxiliary material property X m* , X k* representing the uncertainty in
both material and geometric properties. This can be done using the information in Annex A and the assumption
that structural resistance Rmodel calculated by the numerical model is approximately linearly dependent on the
1
1.46 according to the new version of Annex A (N0267), while 1.54 according to the previous version of Annex A (N 222). This version of the
document was initially prepared according to Annex A version N 222. All examples are calculated according to this Annex A version with the
exception of the last punching example, which was calculated according to Annex A version N0267.
J. Cervenka
For citations, copying and referencing see Note on Front page! page 597
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
where dc is concrete section area, dy is the effective depth, f c , aIs is the actual in-situ concrete compressive
strength, and f y is the reinforcement yield strength. In general the uncertainties of the input
f ym
d ym f ym d y , nom d f yk d y , nom f ym
*
(6)
f yk
and coefficient of variation:
By analogy the uncertainties in the concrete contribution can be expressed by its mean value:
f cm
d cm f cm d c ,nom Ac Is f ck d c ,nom f cm* (8)
f ck
The material strengths in (6) and (8) can be described by auxiliary variables:
f ym
*
f ym d f yk d exp 1.645V fy f yk (10)
f yk
f ck Ac Is exp 1.645V fc f ck
f cm
f cm* Ac Is (11)
f ck
which represent the material mean strength considering the uncertainties of the material strength, effect of
insitu strength as well as geometric uncertainties.
Using the analogic assumptions the auxiliary variables for characteristic analysis in Eq. (4) can be calculated
as:
f yk* f ym
*
exp 1.645 VRS f yk d exp V fy VRS (12)
These assumptions are in agreement with the Eurocode EN 1992-1-1:2005 approach of partial factors for
materials as shown for instance in the case of F.4.
It should be noted however, that in nonlinear analysis it is usually necessary to define other parameters to
describe the behaviour of reinforced concrete structures. It is desirable to define unique rules how these
additional parameters such as for instance: tensile strength ft , fracture energy GF and others are related to
these basic strength parameters. Then all material parameters can be uniquely derived from the auxiliary steel
yield strength f y* and concrete compressive strength f c* . Alternatively, the mean and characteristic values of
the additional material parameters can be derived using JCSS probabilistic Model Code [16].
J. Cervenka
For citations, copying and referencing see Note on Front page! page 598
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
Table 2: Standard values for auxiliary steel and concrete material parameters for the safety format F.5
Parameter persistent and transient design situations
Reinforcement
Concrete
Table 3: Standard values for auxiliary steel and concrete material parameters for the full probabilistic analysis
F.6
Parameter persistent and transient design situations
Reinforcement
Coefficient of variation VRC VAc2 V2Is V fc2 VRC 0.12 0.122 0.042 0.16
- The evaluation of the structural strength Rd using the developed numerical model with the given
loading scenario. It should be understood that in most cases Rd is not a single number but it rather
represents an overall load factor for the investigated load/action combination.
J. Cervenka
For citations, copying and referencing see Note on Front page! page 599
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
Additional important tests that should be performed in each project involving application of nonlinear analysis to
structural assessment are project specific sensitivity studies that typically consist of the following types of
analyses:
- mesh sensitivity tests
- solution method tests such as effects of load steps size, solution method parameters, loading history,
various modelling idealizations, boundary conditions, etc.
For each example a table of used material properties is provided as well as the specific values that are used for
each method. This documents compares the design strength values obtained by PFM (F.4), GFM (F.5) and full
probabilistic method (F.6). The probabilistic method can be considered as the most accurate estimation of the
design strength. These three methods are also compared with the analytical formulas from EC2. The final
comparison is available in Table 9 and Table 10.
J. Cervenka
For citations, copying and referencing see Note on Front page! page 600
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
J. Cervenka
For citations, copying and referencing see Note on Front page! page 601
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
Table 5: Overview of material parameters for beams T-8, T-11 from : Bosco & Debenardi [2]
Reinforcement B500H Concrete
Yield strength Strain limit[-] Compressive Tensile Fracture
[MPa] strength [MPa] strength [MPa] energy [Nm]
Test data
f y , (COV)
---
587 (0.04)
ft (COV)
0.07 27.8 – 30.9 --- ---
672 (0.02)
f yk 500 (1)
f ck 25 (1)
EC2 Analytic
(2)
f yd 459 N/A f cd 16.7 --- ---
PFM
GFM
Mean
*
f ym 523 (2)
0.07 f cm* 26.6 *
f ctm 1.91 *
GFm 48.76
(2)
Char. f yk* 478 0.07 f ck* 20.4 *
f ctm 1.47 *
GFk 37.44
Full probabilistic
1.91
(2)
Mean
*
f ym 523 0.07 f cm* 26.6 *
f ctm *
GFk 37.44
Notes:
(1) Test material properties were matched to standard steel f yk 500 MPa and concrete to f ck 25 MPa .
(2) Geometric uncertainties for steel are modified for the case od d = 565, i.e.
J. Cervenka
For citations, copying and referencing see Note on Front page! page 602
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
Figure 4 : Experimental and numerical results for the beam B92B from Hallgren [15]
J. Cervenka
For citations, copying and referencing see Note on Front page! page 603
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
Table 6: Overview of material parameters for beam B92B from Hallgren [15]
Reinforcement B500H Concrete
Yield strength Strain limit[-] Compressive Tensile strength Fracture
[MPa] strength [MPa] [MPa] energy [Nm]
Test data
f y , (STD)
---
631 (±5)
ft , (STD)
0.03 42.1 --- ---
749 (±7)
f yk 586 f ck 36 --- ---
EC2 Analytic
f yd 510 N/A f cd 24 --- ---
PFM
f yd 488 0.025 f cd 23.4 f ctd 2.05 GFd 44
GFM
Mean *
f ym 599 0.025 f cm* 38.2 *
f ctm 3.34 *
GFm 72
Full probabilistic
Mean *
f ym 599
(1)
0.025 f cm* 38.2 *
f ctm 3.34 *
GFk 72
(1)
COV VRS 0.0452 0.052 VRC 0.12 0.12 2 0.042
--- VRC 0.16 VRC 0.16
0.067 0.16
Notes:
Geometric uncertainties for steel are not modified d < 200 mm
J. Cervenka
For citations, copying and referencing see Note on Front page! page 604
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
Figure 5 :Anderson & Ramirez [1] geometry and dimensions, (1 in = 25.4 mm, 1 kip = 4.448 kN,
1 psi = 0.00689 MPa)
Figure 6 :Anderson & Ramirez [1] load displacement curves, (1 in = 25.4 mm, 1 kip = 4.448 kN,
1 psi = 0.00689 MPa)
J. Cervenka
For citations, copying and referencing see Note on Front page! page 605
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
Table 7: Overview of material parameters for beam W1-W2 from Anderson & Ramirez [1]
Reinforcement B500H Concrete
Yield strength Strain limit[-] Compressive Tensile strength Fracture
[MPa] strength [MPa] [MPa] energy [Nm]
Test data
f y , (COV)
587 (0.04)
ft , (COV)
0.07 29.3 --- ---
672 (0.02)
f yk 500 (1)
f ck 25 (1)
EC2 Analytic
f yd 455 (2)
N/A f cd 16.7 --- ---
PFM
f yd 439 (2)
0.07 f cd 16.2 f ctd 1.17 GFd 29.87
GFM
(2)
Char. f yk* 476 0.07 f ck* 20.4 *
f ctm 1.47 *
GFk 37.44
Full probabilistic
*
523 f cm* 26.5 *
1.91
(2)
Mean f ym 0.07 f ctm *
GFk 37.44
(2)
VRS 0.0452 0.0352 VRC 0.12 0.122 0.042
COV --- VRC 0.16 VRC 0.16
0.057 0.16
Notes:
(1) Test material properties were matched to standard steel f yk 500 MPa and concrete to f ck 25 MPa .
(2) Geometric uncertainties for steel are modified for the case od d = 344, i.e.
Vd 0.05 200 / d
23
0.035 , d 1 0.05(200 d ) 2 3 0.965
J. Cervenka
For citations, copying and referencing see Note on Front page! page 606
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
Punching problem :
Punching example taken from the experiments by Guandalini et al. [14]. This example is selected because it is
a slab with large thickness, i.e. representing a large structure where nonlinear analysis can be used in practice.
Figure 7 : Punching experiment by Guandalini et al. [14], PG3 experiment with dimensions B = 6 m,
rg = 2.85 m, c = 0.52 m, h = 0.5 m, reinforcement Ø16@135 mm, cover 28 mm.
Figure 8 : Results of punching experiment of Guandalini et al. [14] and numerical simulation from the
uncertainty factor calibration [10].
J. Cervenka
For citations, copying and referencing see Note on Front page! page 607
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
Table 8: Overview of material parameters for slab PG3 Guandalini et al. [14]
Reinforcement B500H Concrete
Yield strength Strain Compressive Tensile strength Fracture
[MPa] strength [MPa] [MPa] energy [Nm]
limit[‐]
Test data
fy , 520
f ym 0.025 f ck
f cm
f yk
exp 1.645V fy exp 1.645V fc
520 483 32.4
1.077 1.18
27.5
EC2 Analytic
PFM
GFM
Mean
*
f ym 494 0.025 f cm* 30.8 *
f ctm 2.47 *
GFm 61.75
Mean
*
f ym 494 0.025 f cm* 30.8 *
f ctm 2.47 *
GFk 61.75
J. Cervenka
For citations, copying and referencing see Note on Front page! page 608
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
Conclusions and comparison of the design strength values calculated by the methods from
Annex F
This section contains the comparison of design strength values calculated for the investigated examples using
the three safety formats for nonlinear analysis from Annex F. They are:
F.4 Partial factor method,
F.5 Global factor method,
F.6 Full probabilistic method.
The three methods are also compared with the standard analytical formula from the corresponding section of
the Eurocode 2. The analytical formulas are taken from a previous draft 4 of Eurocode prEN 1992-1-1.
The results are compared in two tables: Table 9 and Table 10. The first table contains the comparison of the
actual strength values in terms of the maximal loads, while the second table contains the comparison of relative
values that are normalized with respect to the analytical formula from the Eurocode.
From the result tables, it is possible to draw the following conclusions:
- for cases with clear analytical solution, which is mainly the case of bending with steel failure the nonlinear
analysis provides same results as the analytical solutions,
- for other situations the nonlinear analysis provides slightly more conservative results than the analytical
formulas. This is however a desirable situation since it can be expected that nonlinear analysis will be
mainly used in situations when analytical formulas are not valid or can be applied only with strong
simplifying assumptions,
- GFM method provides always results on the safe side with the exception of the punching case. However,
after the application of the clause F.7(8) even for this case conservative results are obtained,
- it can be observed that more conservative results from the nonlinear analysis are obtained mainly in the
cases when concrete crushing is observed, i.e. bending with crushing, shear with concrete crushing and
punching example. This results is strongly related to the used software where the default software
settings assume rather conservative brittle behaviour of concrete during compressive softening, i.e.
crushing. This shows the importance of the sensitivity and validation studies that should be performed for
each software or material model to be used in structural assessments according to Annex F as they are
described in clauses: F.3(3) and F.7.
It should be noted that currently the results are slightly inconsistent as the material parameters for first 4
examples were calculated with the previous version of Annex A (N 222) while in the case of the last punching
example the Annex A version (N 0267) was used as is explained in the footnote on page 3. The changes in
Annex A are however minor and it should not significantly affect the main results and conclusions of this
document.
J. Cervenka
For citations, copying and referencing see Note on Front page! page 609
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
B92B1-42 [14]
Shear
Bending/shear
Bending - 296 214 1.01 229 1.01 244 1.01
concrete Rd Rd Rd
380 Shear - 223
crushing 191 204 1.13 218 1.13
1.13
[ 38 ] Rd Rd Rd
T11-A1 [2]
Shear 316 [ 30 ]
W1-W2 460-
217 [ 40 ] 228 1.13 211 1.13
Rd
240 1.13
Rd
Anderson & 549 Rd
Notes :
(1) Model uncertainty factor is increased by 15% according to the clause F.7(8)
[1] Anderson, N.S. & Ramirez, J.A. 1989. Detailing of stirrup rein-forcement. ACI Structural Journal 86(5):
References
J. Cervenka
For citations, copying and referencing see Note on Front page! page 610
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex F - Examples Background to FprEN 1992-1-1:2023
507-515.
[2] Bosco C. & Debenardi, P. G. 1993. Influence of some basic parameters on the plastic rotation of
reinforced elements. Ductility requirements for structural Concrete - Reinforcement. Comite Eruo-
international du Breton. CEB bulletin 218. pp25-44
[3] CEB-FIP (Comité Euro International du Béton; Fédération International de la Précontraint). CEB-FIB
Model Code 1990: Design Code. London: Thomas Telford.
[4] P. Castaldo, D. Gino, G. Bertagnoli, G. Mancini (2018): Partial safety factor for resistance model
uncertainties in 2D non-linear analysis of reinforced concrete structures, Engineering Structures, 176,
746-762. https://doi.org/10.1016/j.engstruct.2018.09.041.
[5] P. Castaldo, D. Gino, G. Bertagnoli, G. Mancini (2020): Resistance model uncertainty in non-linear
finite element analyses of cyclically loaded reinforced concrete systems, Engineering Structures,
211(2020), 110496, https://doi.org/10.1016/j.engstruct.2020.110496
[6] Cervenka J, Pappanikolaou V. Three dimensional combined fracture-plastic material model for
concrete. Int J Plast 2008;24(12):2192–220. https://doi.org/10, 1016/j.ijplas.2008.01.004.
[7] Cervenka, V., Cervenka, J., Jendele, L., ATENA Program Documentation, Part 1: Theory, Cervenka
Consulting s.r.o., 2020, www.cervenka.cz
[8] Cervenka, V.: Global Safety Format for Nonlinear Calculation of Reinforced Concrete. Beton- und
Stahlbetonbau 103, (2008), special edition, Ernst & Sohn, pp. 37–42.
[9] Cervenka, V., Reliability-based non-linear analysis according to Model Code 2010, Journal of FIB,
Structural Concrete 01/2013, pp.19-28
[10] Cervenka V, Cervenka J, Kadlec L. Model uncertainties in numerical simulations of reinforced concrete
structures. Structural Concrete 2018;19(6):2004–16. https://doi.org/10.1002/suco.201700287.
[11] Engen M, Hendriks MAN, Köhler J, Øverli JA, Åldtstedt E. A quantification of modelling uncertainty for
non-linear finite element analysis of large concrete structures. Structural Safety 2017; 64: 1-8.
[12] Gino D., Castaldo, P., Mancini, G., Background document to Annex F, Safety formats for nonlinear
analysis, CEN/TC250/SC2/WG1/TG6, 28. 11. 2022
[13] Gino, D., Castaldo, P., Giordano, L., Mancini, G., Model uncertainty in nonlinear numerical analyses of
slender reinforced concrete members. Structural Concrete. 2021;1–26.
https://doi.org/10.1002/suco.202000600
[14] Guandalini S, Burdet OL, Muttoni A. Punching tests of slabs with low reinforcement ratios. ACI Struct J.
2009;106:87–95.
[15] Hallgren M., Flexural and Shear Capacity of Reinforced High Strength Concrete Beams without
Stirrups, TRITA-BKN, Bulletin 9, 1994, ISSN 1103-4270
[16] JCSS, (2001), JCSS Probabilistic Model Code, Joint Committee on Structural Safety, ISBN 978-3-
909386-79-6
[17] ModelCode 2010, fib Lausanne, Ernst & Sohn: Switzerland, 2011
J. Cervenka
For citations, copying and referencing see Note on Front page! page 611
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
22.12.2022
For citations, copying and referencing see Note on Front page! page 612
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex G Background to FprEN 1992-1-1: 2023
Introduction
This document has been prepared to clarify the background for the provisions introduced in subsections G.3 -
G.5 of Annex G in FprEN-1992-1-1:2023. The subsections provide methods for ULS design of membrane-,
shell-, and slab elements and SLS analysis of membrane elements.
Figure 1: Roadmap to link the various provisions in Section 8 with Annex G. Topics written in red font are new
and not included in EN 1992-1-1:2004.
L. C. Hoang 22.12.2022
For citations, copying and referencing see Note on Front page! page 613
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex G Background to FprEN 1992-1-1: 2023
reinforcement in the x-direction. The total amount of required reinforcement in this direction, which carries a
tension force of the magnitude F = Vcot, can however be re-allocated to the positions of the chords. Hence,
half of the horizontal reinforcement (i.e. ½Vcotis allocated to the tension chord while the other half is
allocated to the compression chord as a reduction of the compression force. The re-allocation is illustrated in
Figure 4. As can be seen in Figure 4, the re-allocation of the horizontal reinforcement leads to the same design,
as if the beam was designed using subsection 8.2.3, where the shift-rule required ½Vcotto be added
(subtracted) to the tension (compression) chord.
Figure 3: Free body diagrams and design scheme for cantilever beam using the method of Annex G
Figure 4: Free body diagram and design scheme for cantilever beam using the method of Subsection 8.2.3.
L. C. Hoang 22.12.2022
For citations, copying and referencing see Note on Front page! page 614
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex G Background to FprEN 1992-1-1: 2023
L. C. Hoang 22.12.2022
For citations, copying and referencing see Note on Front page! page 615
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex G Background to FprEN 1992-1-1: 2023
i.e. x = 4y. With this design, and in a proportional loading scheme, the weak reinforcement (i.e. y) will start to
yield long before the load reaches the ultimate state. Once the weak reinforcement yields, redistribution of
stresses takes place by rotation of the strut to a flatter angle (scenario 2) while the concrete stress, c,
increases and while the stress in the strong reinforcement (x) increases as well. At the final state, = 26,6o
corresponding to cot= 2 and the strong reinforcement reaches the yield stress. In this case, x can therefore
be taken as the yield strain while y is larger, but unknown. The opposite will be the case, if instead the design
is based on cot= ½, which will lead to scenario 3).
In order to avoid a situation where the approach leads to too conservative (small) values of , guideline (c) has
been included. Guideline (c) in Clause G.3(7) covers the important situations, where the assumption of zero
tensile strength in the concrete (which is the basis for the Expressions (G.3) – (G.5)) simply lead to a large
estimate of the transverse strain and thereby too small a -value. This issue can be illustrated by the following
example:
Example:
Assume: fyk = 500 MPa, fck = 30 MPa and fctk = 2 MPa (no partial coefficient)
The calculations are performed without partial coefficients. Both cases require according to table G.1 tension
reinforcement (ftdx = ftdy = 9 MPa for case 1 and ftdx = ftdy = 1.5 MPa for case 2) and cot = 1. Since the tensile
strength of concrete is neglected, then reinforcement has to yield in both cases and therefore, 1 and thus also
will in both cases assume the same value. This gives when guideline (b) together with Expressions
(G.9) –(G. 11) are used. The consequence is that in case 1, we can verify the compressive strength (c = 2xy =
18 MPa < fc = 0.61*30 = 18.3 MPa) while in case 2 we cannot (c = 2xy = 19 MPa > 18.3 MPa). The fact is,
however, that in case 2, we actually have a higher efficiency factor because the required tension reinforcement
provides a tensile strength (ftdx = ftdy = 1.5 MPa) smaller than the tensile strength of the concrete. The member is
therefore un-cracked. Therefore, in this case 1 to be inserted into Expression (G.9) should not be determined
by taking x and y as the yield strain for of the reinforcement. The most appropriate value forx and y would in
this case be fct/Ec. This would lead to an increase of to ca. 0.89 leading to c < fc.
It should here be noted, that Section 12 (Table 12.4) of FprEN-1992-1-1:2023 provides rules for minimum
reinforcement in walls designed according to Annex G. The minimum requirements may in some situations lead
to higher reinforcement demands than what is necessary according to Annex G. The reason for this is that the
method of Annex G requires the member to have an appropriate deformation capacity ensured through the
possibility to have a distributed crack pattern (smeared cracks). The suggested minimum reinforcement in Table
12.4 (0,25Ac*fctm/fyk on each surface) corresponds well with tests, see e.g. ref. [13].
Figure 5: Yield condition for reinforced concrete membrane elements (from [2])
Due to the fact that the -factor (G.9) depends on the transverse strain and thereby the stress field inclination,
an iterative procedure must be used to calculate the shear capacity of membranes. In the following calculation
L. C. Hoang 22.12.2022
For citations, copying and referencing see Note on Front page! page 616
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex G Background to FprEN 1992-1-1: 2023
scheme, it is assumed that the reinforcement capacity in the x-direction is larger than the one in the y-direction,
i.e. x f y , x y f y , y (notation of Annex G). This assumption means that in case there is redistribution of
stresses, it will be the weak reinforcement (in the y-direction) that undergoes strains larger than the yield strain
while the strain in the strong reinforcement (in the x-direction) will be uniquely defined at the ultimate state. The
procedure includes the following steps:
----------ooOOoo----------
Step 1:
Make an initial guess of
Step 2:
Calculate the mechanical reinforcement degrees:
x f y,x y f y, y
x ; y
f c* f c*
where
1
f c* cc f c & cc min 40 MPa
1/3
fc
Here, cc is the brittleness factor as defined in clause 5.1.6(1) of prEN-1992-1-1:2021.
Step 3:
Calculate the shear capacity, u, the strut inclination, , and reinforcement strain, x, according to following
scheme:
IF: x y 1:
x f y,x
u f c* x y ; cot ; sx f y , x x
y Es
IF: x y 1:
*
f c y 1 y ; y 0.5
u
1
f c* ; y 0.5
2
1
y ; y 0.5
cot y
1 ; y 0.5
f c* cos 2 sx
sx x
x Es
Step 4:
1
1 1101
where:
L. C. Hoang 22.12.2022
For citations, copying and referencing see Note on Front page! page 617
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex G Background to FprEN 1992-1-1: 2023
1 x ( x 0,001)cot 2
Step 5:
Repeat steps 1 – 4 until the value in step 4 equals the value in step 1.
----------ooOOoo----------
Calculations have been carried out for benchmark test series published by Vecchio & Collins; Yamaguchi &
Naganuma; Vecchio et al.; Pang & Hsu and Zhang & Hsu [3 –7].
In order to evaluate Expression (G.9), only test specimens where the predicted strength is governed by
concrete crushing, combined with yielding of the reinforcement in the y-direction or combined with no rebar
yielding at all, are interesting. Specimens where the predicted strength is governed by yielding of reinforcement
in both directions (see solution above for the case x y 1 ) are not interesting, since this case is not
critical to concrete crushing.
The obtained results can be seen in Figures 6 – 10 where the ratio between tested and calculated shear
strength has been plotted versus strength parameters (y and fc) and calculated parameters (sx, cotand 1),
each one having and influence on the -value. Note that in the figures, the term “orthotropic reinforcement”
means x f y , x y f y, y While “isotropic reinforcement” means x f y , x y f y , y .For all tests, the mean
Figure 6: test /cal. vs. mechanical reinforcement degree, y, for the weak reinforcement direction
Figure 7: test /cal. vs. calculated level of tension stress in reinforcement along the x-axis (not yielding)
L. C. Hoang 22.12.2022
For citations, copying and referencing see Note on Front page! page 618
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex G Background to FprEN 1992-1-1: 2023
Figure 8: test /cal. vs. calculated 1st transverse strain in the concrete
L. C. Hoang 22.12.2022
For citations, copying and referencing see Note on Front page! page 619
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex G Background to FprEN 1992-1-1: 2023
Clause G.4(1)-(2)
To have an approach for design of shell elements, which is consistent with the provisions for membrane
elements, subsection G.4 adopt a classical sandwich model. This makes it possible to establish a set of
membrane stresses in the two layers of the sandwich model, which are statically equivalent with the in-plane
forces and the torsional and bending moments acting on the shell elements. Calculation of the reinforcement in
the two layers can then be performed according to subsection G.3.
Clause G.4(5)
For slabs, i.e. shells without in-plane forces, explicit design formulas are provided in Expressions (G.18-G.21)
as well as in Tables G.2 and G.3. The formulas and the rules for limitation of redistribution of moments are
analog to those applied to membrane elements. The theoretical background for these formulas may be found in
e.g. [2]. It is noted that the formulas in Tables G.2 and G.3 have been implemented in a number of commercial
finite element software packages. Finally, it is noted that there is a limitation on the torsional moment, |MEdxy | <
0,07d2fcd, which has to be fulfilled when using the Expressions (G.18 – G.21) to calculate the required
reinforcement. This is necessary because these expressions assume that the reinforcement required to carry
bending as well as torsion can be obtained from a classical sectional analysis. However, the torsional capacity
for slabs can be significantly lower than the flexural capacity if the reinforcement content is high. This is so
because there is a different in the internal lever arm for the two cases. Furthermore, in the case of torsion there
is a reduction of the concrete compression strength due to transverse cracking at the top as well as at the
bottom of the slab. The procedure to calculate the torsional capacity of slabs in a more accurate manner can be
found in e.g. Section 7.3 of Ref. [2]. Figure 11 shows calculated flexural capacity (mR) and calculated torsional
capacity (tR) of an isotropic reinforced slab element. As can be seen, for relatively low reinforcement content, mR
and tR are practically the same. This tendency continues up to a capacity of approximately mR = tR = 0,07d2fcd.
As the reinforcement content increases, much smaller torsional capacity is obtained compared to the
corresponding flexural capacity. Hence, for torsional moments larger than approximately 0,07d2fcd, it is unsafe to
use a flexural sectional analysis to calculate the reinforcement needed to carry combinations of bending and
torsion.
Figure 11: Calculated flexural capacity (mR) and torsional capacity (tR) of isotropic reinforced slab element. The
calculation of tR follows the procedure in [2], where As denotes the reinforcement area per unit length of the slab
cross section. The results are valid for = 0.6 and d/h = 0,84, where h is the total depth.
Clause G.5
Clause G.5 provides guidelines for calculation of crack widths in membrane elements with skew cracks. The
provisions apply when the skew cracks have an inclination, , which is larger than 15o with the reinforcement
axes.
L. C. Hoang 22.12.2022
For citations, copying and referencing see Note on Front page! page 620
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex G Background to FprEN 1992-1-1: 2023
Figure 12: Fully rotational model showing normal strains in reinforced membrane element (averaged over
distance containing several skew cracks) and normal strains in the concrete between two cracks.
The provisions of Clause G.5 are primarily based on the Cracked Membrane Model [8], where the following
main assumptions should be mentioned:
1) The skew cracks are the result of a stress state that includes a diagonal compression field in the
concrete and tension in the reinforcement bars.
2) The reinforcement content is sufficient to ensure a smeared crack system (i.e. formation of more than
just one skew crack)
3) The direction of the compression field coincides with the direction of the cracks (rotating crack model
with stress free cracks)
4) Average total strains in the cracked membrane element = sum of average strains in concrete and
average strains due to crack opening.
5) Effect of tension stiffening is estimated based on a rigid-plastic bond model (bond between rebars and
concrete)
Figure 12 illustrates a cracked membrane element, which fulfills the above-mentioned assumptions. Here, 1
and 2 are the average principal strains (total strains) of the cracked element when measured over several
cracks while c1 and c2 are the average principal concrete strains between cracks. Due to the assumption 3), 2
= c2. Based on assumptions 3) and 4) it is possible to estimate the crack width, w, from the relationship: w/s =
(1 - c1), where s is the spacing between the skew cracks. Transformation from mean crack width to calculated
crack width, wk,cal, takes place by means of the factor kw defined in Clause 9.2.3(2) of FprEN-1992-1-1:2023.
Annex G provides Equation (G.22) to estimate the mean spacing between the skew cracks. This equation is
identical to Equation (5.12) in [8] as well as the one introduced in the fib model code 2010 [9]. The above-
mentioned strain difference (1 - c1) is in Annex G denoted as (sm - cm) and given by Equation (G.23). It is
noted that the right-hand side of Equation (G.23), with input parameters given in Equations (G.24) – (G.26), has
been deliberately written in a format that as much as possible resembles the format of Equation (9.11) in FprEN-
1992-1-1:2023 for the case of uniaxial tension. However, these equations are in fact the result of a
decomposition and rearrangement of the Equation (5.22) in [8] derived from the Cracked Membrane Model, in
which the Poisson’s ratio of concrete is assumed to be zero and where (1- ) has been approximate with (1- )
~ 1 (assuming the reinforcement ratio to be much smaller than unity). For details, the reader is referred to [8]. It
should be noted that in many practical situations, the term |2| in (G.23) is small compared to the other two
terms and may therefore be neglected. Finally, it is noted that the upper limit (1-kt)s/Es in Equations (G.24-25)
has been introduced to achieve consistency with Equation (9.11) and with experimental results. However, unlike
the case of uniaxial tension (i.e. Equation (9.11)), where the effective tensile strength of concrete is taken as fctm,
Equations (G.24-25) assume fct,eff = 0,5fctm. This value has been recommended here because many membrane
shear test series (see e.g. [13]) have shown that transition from un-cracked to cracked element takes place
already when the major principal stress 1 is in the order of 0,5 – 0,6fctm. The introduction of fct,eff = 0,5fctm in
Annex G is consistent with Table 12.4, where the rules for minimum reinforcement in walls subjected to in-plane
normal and shear stresses are based on fct,eff = 0,5fctm.
The inclination,, of the skew cracks may be determined by use of Equation (G.27). This equation simply
expresses the compatibility condition of a fully cracked membrane element (concrete tensile strength neglected)
subjected in-plane stresses x, y and xy. The equation can e.g. be derived by minimizing the complementary
elastic energy of the membrane element.
Calculated crack widths according to Clause G.5 (and with kw = 1.7, see Equation (9.8)) have been compared
with the maximum observed crack widths in membrane tests extracted from refs. [6, 10 - 12]. The results are
shown in Figure 13. When assuming a lognormal distribution, the mean value of the test-to-calculation ratio is
0,86 with a standard deviation of 0,41.
L. C. Hoang 22.12.2022
For citations, copying and referencing see Note on Front page! page 621
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex G Background to FprEN 1992-1-1: 2023
Figure 13: Calculated crack width compared with observed maximum crack widths in selected membrane tests
References
[1] Muttoni A, Ruiz MF, Pejatovic M: Background document to subsection 8.2.3 – Members requiring
design shear reinforcement. CEN/TC250/SC2/WG1/TG4, February 2021.
[2] Nielsen, MP, and Hoang, LC.: Limit Analysis and Concrete Plasticity. Third edition, CRC Press, Boca
Raton, FL, 796 pp.
[3] Vecchio FJ, Collins MP: The response of reinforced concrete to in plane shear and normal stresses.
The Department of Civil Engineering, University of Toronto, Canada, 1982.
[4] Yamaguchi T, Naganuma K.: Experimental study on mechanical characteristics of reinforced
concrete panels subjected to in-plane shear force . Journal of Struct. Constr. Engng., AIJ, No. 419,
1991.
[5] Vecchio FJ, Collins MP, Aspiotos J: High-Strength Concrete Elements Subjected to Shear. ACI
Structural Journal, Vol. 91, No. 4, July - August 1994. pp. 423-433.
[6] Pang XB, Hsu TTC: Behaviour of Reinforced Concrete Membrane Elements in Shear. ACI Structural
Journal, Vol. 92, No. 6, Nov - December 1995. pp. 665-679.
[7] Zhang LX, Hsu TTC : Behaviour and Analysis of 100 MPa Concrete Membrane Elements. Journal of
Structural Engineering, ASCE, Vol. 124, No. 1, Jan. 1998, pp. 24 – 34
[8] Kaufmann W: Strength and deformations of structural concrete subjected to in-plane shear and
normal forces. Doctoral thesis, Swiss Federal Institute of Technology Zürich, Institute of Structural
Engineering, 1998.
[9] Federation international du beton: fib Model Code for Concrete Structures 2010. Ernst & Sohn, 2013.
[10] Proestos, GT: Influence of high-strength reinforcing bars on the behaviour of reinforced concrete
nuclear containment structures subjected to shear. MSc-thesis, Department of Civil Engineering,
University of Toronto; 2014.
[11] Marti P, Meyboom J: Response of prestressed concrete elements to in-plane shear forces. ACI Struct
J 1992;89(5):503–14.
[12] Dyngeland T: Behaviour of reinforced concrete panels. PhD-thesis, Department of Structural
Engineering. Trondheim, Norway: Norwegian University of Science and Technology; 1989.
[13] Beck A: Paradigms of shear in structural concrete: Theoretical and experimental investigation.
Doctoral thesis, Swiss Federal Institute of Technology Zürich, Institute of Structural Engineering,
2021.
L. C. Hoang 22.12.2022
For citations, copying and referencing see Note on Front page! page 622
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex G Background to FprEN 1992-1-1: 2023
L. C. Hoang 22.12.2022
For citations, copying and referencing see Note on Front page! page 623
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
For citations, copying and referencing see Note on Front page! page 624
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex H Background to FprEN 1992-1-1:2023
Introduction
The content of Annex H is mainly taken from EN 1992-3:2006. The content of this standard relating to water-
retaining structures has been incorporated into FprEN 1992-1-1:2023, part of in into Annex H, and part of it
into Section 9.2.3. In this background document this rearrangement of content will be documented. Parts of
EN 1992-3:2006 dealing with containment structures were discarded as CEN-TC250/SC2 considered that
this content was not concrete-specific and design issues of such structures built using concrete could be dealt
with using the content of the general part. For this reason, the scope of Annex H is limited to water-tightness
of concrete structures, although it may also be used for other liquids if it can be shown that their behaviour is
similar to that of water.
Some new content is provided by suggestion of a systematic review comment to EN 1992-1-1:2004 by the
Belgian Mirror Group. The suggestion was to reinstate content for ENV 1992-4:1998 regarding the calculation
of the leakage rate through cracks. The derivation of this proposal is also documented and discussed in this
BD.
In order to derive the proposed expression (for the basic equation see (Vennard, 1947)), consider the
equilibium condition of a mass of water centered around the crack axis, as shown in Fig. 1. h is the depth of
the wall, and p1 and p2 are the water pressures at either end of the wall.
2 p1lw, p y x 2 p2 l w , p y
y
h
y y
2 lw, p h 2plw, p y p p (1)
h h
This equation shows that the shear stress is proportional to the distance from the centre of the crack.
v=0
wk,cal,e vc
x
Fig. 2 Variation of velocity of flow as a function of the distance form an edge (taken from ref. [1])
Fig. 2 shows that the velocity of flow increases as the distance from an edge increases. At the edge the
velocity is zero. This variation in velocity is due to the existence of a shear stress between fluid layers. It has
been found, by experimental evidence, that this shear stress is proportional to the gradient of velocity in the
direction perpendicular to the flow. The constant of proportionality is known as the coefficient of viscosity (or
just viscosity for short) and is represented by symbol .
dv dv
(2)
dy dy
By combining expressions (1) and (2) and applying them to a thin rectangular pipe of width wk,cal,e the following
differential equation is obtained
y dv
p
h dy
y v
y p
h 0
p dy dv ydy dv
h vc
(3)
p y 2
vc v v vc Ky 2
h 2
wk ,cal ,e 4 y2
y v0K vc v vc 1 4 2
2 wk2,cal ,e wk ,cal ,e
The mean velocity of water in the pipe, can be determined from the definition of flow rate:
wk ,cal ,e
2
y2
q wk ,cal ,elw, p vm vdA lw, p vc 1 4 2 dy
wk ,cal ,e
A
wk ,cal ,e
2
wk ,cal ,e
4 y3 2 2 (4)
lw, p vc y 2 vc wk ,cal ,elw, p
3 wk ,cal ,e wk ,cal ,e 3
2
3 2
vc vm vm vc
2 3
From Eq. (3), by using the formula for the velocity at y=wk,cal,e/2, where v=0, an expression can be obtained
for the flow rate, accounting for the definition of q as the section area times the mean velocity:
2
p wk ,cal ,e 3 3 q
vm
h 8 2 2 l p , w wk ,cal ,e
(5)
1 l 1
q wk3,cal ,e p p , w
12 h
This equation, which is the classical equation for Poiseuille flow between two parallel planes, is similar to Eq.
(A107.1) of ENV 1992-4:1998. However, the coefficient proposed by this Standard is 1/50 (instead of 1/12),
with a value that can be increased from 1/50 to 0 in cracks with a calculated crack width smaller than 0.2 mm
where self-healing occurs.
The difference in the coefficient is in that the derivation does not account for the tortuosity and roughness of
the cracks. For small cracks, such roughness would represent a significant part of the channel width.
(Chivers, 2002) [1] presents a study of the influence of roughness of the crack on the flow rate. Based on
(Jones, Wooten, & Kaluza, T.J., 1988) [2], (Button, Grogan, Chivers, & Manning, 1978) [2] and (Gardiner &
Tyrrell, 1968) [4], (Chivers, 2002) fixes an upper bound for the friction coefficient (f) in Darcy’s equation for
laminar flow equal to 96/Re, where Re is Reynolds number.
Based on this value, an alternative derivation can be made directly from Darcy’s empirical equation, adapted
for the case of a thin rectangular section (see Eq. (6)).
p h vm2
hL f (6)
water g wk ,cal ,e g
where:
f is Darcy’s friction coefficient
g is the acceleration of gravity on earth
Reynolds’ number can be expressed as follows for the case of a narrow rectangular pipe:
water
Re v w (7)
m k ,cal ,e
Substituting the upper bound value of f in Eq.(6), and accounting for the expression of Reynold’s number, the
proposal of FprEN 1992-1-1:2023 can be readily derived, as shown in Eq. (8).
References Bibliografía
[1] Button, B., Grogan, A., Chivers, T., & Manning, P. (1978). Gas Flow Through Cracks. ASME J. Fluids
Engng, 100, 453-458.
[2] Chivers, T. (2002). The influence of surface roughness on the fluid growth trhough cracks. Fatigue Fract
Engng Mater Struct, 25, 1095-1102.
[3] Gardiner, G., & Tyrrell, R. (1968). The flow resistance of experimental models of naturally ocurring cracks.
Proc. Inst Mech Engrs., 200, No. C4.
[4] Jones, T., Wooten, S., & Kaluza, T.J. (1988). Single Phase Flow Through Cracks. 63rd Annual Technical
Conference of the Society of Petroleum Engineers (pág. Paper SPE 18175). Houston, USA: SPE.
doi:https://doi.org/10.2118/18175-MS
[5] Vennard, J. K. (1947). Elementary Fluid Dynamics. New York: John Wiley & Sons, Inc.
27.03.2023
For citations, copying and referencing see Note on Front page! page 629
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
27.03.2023
For citations, copying and referencing see Note on Front page! page 630
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
Introduction
When concrete strength is assessed in an existing structure, according to clause I.5.2.1(1) of FprEN 1992-1-
1:2023 the characteristic value 𝑓 , of the in-situ compressive strength of concrete cores should be evaluated
according to EN 13791:2019. However, in the main part of FprEN 1992-1-1:2023, all formulations that include
concrete strength are expressed in terms of 𝑓 . This is why there is the necessity to include a formula for the
evaluation of 𝑓 on the basis of the assessed 𝑓 , .
Definitions
𝑓 Compressive strength of concrete determined from samples of concrete taken in accordance with EN
12350-1:2019 made into cylinder or cube specimens and cured in accordance with EN 12390-2:2019
and tested in accordance with EN 12390-3:2019.
𝑓 Characteristic value of 𝑓 .
𝑓, Actual in-situ concrete compressive strength in the structure.
𝑓, Compressive strength of concrete cores taken at test location within a structural member, expressed
in terms of the strength of a 2:1 core of diameter ≥ 75 mm (5% fractile) and assessed according to
Section 8 of EN 13791:2019.
𝑓 , Characteristic value of 𝑓 , .
. ∙
,
𝑓 𝑓 , (2)
∙ .
In this scenario, in Annex I the formulation of EN 13791:2019 has been assumed if the cores are extracted
from the region governing for the verification (see case c) of table I.2).
If the cores are extracted from regions not necessarily representing the conditions of whole structural member
nor the region governing for the verification, the effect on 𝜂 of the position of the cores, within the structural
member, should be taken into account.
Figure 1, taken from Moccia et al. (2020) and based on several test results present in the literature, shows the
influence of the position of cores taken from columns on 𝜂 and in particular
top layer (upper 20% of the depth, red markers in the figure): 40 values, 𝜇 0.95, 𝑉 0.095;
bottom layer (bottom 20% of the depth, black markers in the figure): 23 values, 𝜇 1.05, 𝑉 0.105.
Thus, if the cores have been extracted from regions not necessarily representing the conditions of the whole
structural member nor the region governing for the verification, there is a risk that the region with the highest
μ is assessed (for instance μ 1.05), whereas the governing verification region could require e.g. μ
0.95. In this case, a safe estimate of 𝑓 from 𝑓 , would be
. , ,
𝑓 ∙ (4)
. . .
For citations, copying and referencing see Note on Front page! page 631
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
Figure 1. Coefficient 𝜂 in bottom and top layers of columns related to concrete cylinder strength 𝑓 ,
(Moccia et al. 2020)
References
Moccia, F., Kubski, X., Fernandez Ruiz, M., Muttoni, A. (2020). The influence of casting position and
disturbance induced by reinforcement on the structural concrete strength, Structural Concrete, 2020, 1-28.
For citations, copying and referencing see Note on Front page! page 632
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
27.03.2023
For citations, copying and referencing see Note on Front page! page 633
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
I.4.2.1(2) The approach proposed in this clause is consistent with that included in Annex A of FprEN 1992-1-1:2023. If
the coefficient of variation of concrete core strength is greater than the values given un table I.1, the relevant
Adjusted partial factors
partial factor for concrete c, evaluated by using the procedure given in Annex A, is greater than 1.5. This is
why, in this case, it is recommended to adjust c, according to the procedure given in Annex A. Similarly, If the
coefficient of variation of reinforcement yield strength is greater than the values given un table I.2, the relevant
partial factor for reinforcement s should be adjusted by using the procedure given in Annex A since it is greater
than 1.15.
I.5.3.(2) The formulation included in the clause is consistent with that given in D.7.2 of prEN 1990 draft April 2022, for
log-normal distribution. The clause includes also the table relevant to 10% characteristic value (not included in
Characteristic value of
prEN 1990:2022) for the evaluation of ductility properties of reinforcement.
the properties of
reinforcing steel Table I.5 has been obtained by using the so-called ‘prediction method’ that is the same as that used in Annex
D of prEN 1990:2020 (see Appendix C of Gulvanessian, Calgaro, Holický 2002, ‘Designers’ Guide to Eurocode:
Basis of Structural Design: EN 1990’, Thomas Telford Publishing, London).
I.8.3.2(3) The clause gives a conservative and simplified formulation for the evaluation of shear strength of beams and
slabs having transverse spacing of shear legs greater than that given (as maximum value) in clause 12 of
Shear strength in case
FprEN 1992-1-1:2023. The starting point of this clause is the experimental study by Lubell et al. 2009 (i.e.
of substandard
Lubell A.S., Bentz E.C., Collins M.P. 2009, ‘Shear Reinforcement Spacing in Wide Members’, ACI Structural
transverse spacing of
Journal, Vol. 106, No. 2).
shear legs
The formulation considers that, in the transverse direction, the concrete part of beam/slab that contributes to
shear strength is not greater, per each side of the leg, than half the maximum leg distance given in clause 12
of FprEN 1992-1-1:2023.
I.8.3.3 The clause gives a conservative and simplified formulation that follows the same approach as that of clause
I.8.3.2(3).
Shear at interfaces in
case of substandard
spacing between
interface reinforcement
in a composite slab
I.9.1(2) In the assessment, low values of the target reliability index 𝛽 for non-seismic ultimate limit states can be
allowed. In this scenario fulfilling ULS requirements can also imply that high stresses occur under service load
Limits on reinforcement
situations. This is why stresses in the serviceability limit state should be limited to avoid yielding of the
and concrete stress for
reinforcement and development of micro cracks and cracks in concrete.
low target reliability
index at ULS The stress limits in table I.6 are the same as those in tables 9.1 and 9.2 of FprEN 1992-1-1:2023. However,
explicit reference to these tables has not been made since the stress verifications in table 9.2 are limited to
some exposure classes while in I.9.1(2), for the above-mentioned reasons, they don’t depend on the exposure
classes.
For citations, copying and referencing see Note on Front page! page 634
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
27.03.2023
For citations, copying and referencing see Note on Front page! page 635
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
Introduction
This section provides the background to clauses (3)-(5) of I.11.4.1 and details the derivation of the closed-form
equation for the design anchorage and lap length of plain surface bars of round and square cross section.
. . .
PP 𝑓 4.13MPa ∙ ∙ ∙ (2)
where 𝑓 is the mean value of the axial stress 𝑓 (called anchorage/lap strength in the following) developed in
the bar by bond (MPa); 𝑙 is the anchorage/lap length; 𝑓 is the mean cylinder concrete compressive strength
(MPa); 𝑐 is the concrete cover; 𝜙 is the equivalent bar diameter assumed as the bar diameter for round section
bars and as
𝜙 𝐴 (3)
for square section bars (𝐴 is the area of the cross section of the square section bar).
The approach proposed by Palmisano et al. (2020) is based on the statistical evaluation of test results. To this
aim model uncertainties, that account for random effects that are neglected in the models and simplifications in
the mathematical relations, have to be considered. According to JCSS (2001), the most common way to introduce
model uncertainties in the calculation model is to use an additive equation or a multiplicative equation or a
combination of both. In this case the following multiplicative relationship has been chosen as proposed by Mancini
et al. (2018):
𝑦 𝜃 ∙𝑦 , (4)
where 𝜃 is the model uncertainty of the ith test, 𝑦 is the ith test result and 𝑦 , represents the relevant value
resulting from the assumed model.
Moreover, taking into account that the proposed equations have a nonlinear form, the coefficient of determination
𝑅 is evaluated by using the following equation:
∑ ,
𝑅 (5)
∑
For citations, copying and referencing see Note on Front page! page 636
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
(a)
400
Good bond
350 casting
position
300
250
fstm,TEST [MPa]
200
Round bars
150 Square bars
100
50
0
0 50 100 150 200 250 300 350 400
fstm,MOD [MPa]
(b)
400
Poor bond
350 casting
position
300
250
fstm,TEST [MPa]
100
50
0
0 50 100 150 200 250 300 350 400
fstm,MOD [MPa]
Figure 1. Comparison between bar stress estimated by equations (1) and (2) and that measured in tests. (a)
equation (1) for GP. (b) equation (2) for PP.
𝑓 𝜃∙𝑓 ∙𝐴 (6)
where A represents the multiplicative term relevant to all deterministic parameters of equations (1) and (2)
𝐴 𝛿 𝑀𝑃𝑎 ∙ ∙ ∙ (7)
and 𝛿 , 𝛿 , 𝛿 , 𝛿 are the coefficient/exponents of equations (1) and (2), summarised in table 2.
For citations, copying and referencing see Note on Front page! page 637
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
Casting
Equation 𝜹𝟏 𝜹𝟐 𝜹𝟑 𝜹𝟒
position
By applying the procedure shown in Taerwe (1993) to equation (6), the fractiles 𝑓 of the variable 𝑓 can be
evaluated:
𝑓 𝐴∙ ∙𝑓 2
∙ exp 𝛿2 ∙ ℎ ln 1 𝑉2 ℎ ln 1 𝑉 2
𝛿2 2 𝑙𝑛 1 𝑉2 (8)
1 2
where j = m, k, d with:
m = mean value;
k = characteristic value assumed as the 5% fractile;
d = design value relevant to a specific reliability index β;
and coefficients hj are:
hm = 0 for the mean value;
hk = 1.645 for the characteristic value;
hd = αRβ for the design value, where αR is the FORM (First-Order Reliability Method) correction factor
assumed equal to 0.8 for dominant resistance variables (JCSS 2001).
Equation (8) can be rewritten to express directly the design anchorage length 𝑙 as a function of the design axial
stress 𝜎 in the reinforcement and other relevant parameters in a format similar to that given FprEN 1992-1-
1:2023 for ribbed bars:
. ∙
𝛥 ∙ ∙ ∙ (9)
where
∙
𝛥 (10)
.
𝜁 ∙ exp 𝛿 ∙ ℎ ln 1 𝑉 ℎ ln 1 𝑉 𝛿 𝑙𝑛 1 𝑉 (11)
If hd,=3.8 is the value of hd evaluated for = 3.8, equation (9) can be rewritten, for practical application, as a function
of the concrete partial factor C:
. ∙
𝛥 , . ∙ ∙ ∙ ∙ (12)
.
∙ , .
𝛥 , . (13)
.
𝜁 , . ∙ exp 𝛿 ∙ ℎ ln 1 𝑉 ℎ , . ln 1 𝑉 𝛿 𝑙𝑛 1 𝑉 (14)
The exponent 5 has been obtained by evaluating lbd, by equation (9), for values of from 3.3 to 5.2 (i.e. from low
consequences of failure to a failure probability equal to 10-7 according to EN 1990:2002) and imposing a ratio of
equation (12) to equation (9) in the range 0.96-1.04 to make negligible the difference between the rigorous
approach (i.e. equation (9)) and the practical one (i.e. equation (12)). To this aim the formulation of C, included
in annex A of FprEN 1992-1-1:2023 (with the parameters included in table A.3), has been assumed.
By using this approach and the coefficients/exponents of equations (1) and (2), equation (12) yields equations
(15) and (16) for GP and PP, respectively.
. . . ∙ . .
132.9 ∙ ∙ ∙ ∙ (15)
.
. . . ∙ . .
414.8 ∙ ∙ ∙ ∙ (16)
.
For citations, copying and referencing see Note on Front page! page 638
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
As the original equations have a semi-empirical origin, it is reasonable to round the coefficient/exponents in order
to have the final design equations. Hence equations (15) and (16) can be rewritten and compacted in the final
equation (17) to be used for practical application.
. ∙
130 ∙ 𝜂 ∙ ∙ ∙ ∙ (17)
.
where
1 = 2 = 3 = 4 = 1 for GP;
1 = 3.1, 2 = 1.6, 3 = 0.9, 4 = 3/5 for PP.
The following limitations should be applied to equations (15)-(17):
the ratio 1.5eq/c should not be taken smaller than 0.5 in view of absence of test data in the reference
database beyond c/eq = 3.0;
the application of the equations should be limited to sd ≤ 300 MPa since in the reference database most
of the tests have fstm lower than 280 MPa with just a few having fstm of about 350-400 MPa;
the equations are valid for c/eq ≥ 1.0 in view of absence of test data in the reference database below
c/eq = 1.0;
the ratio should be taken not lower than 10, since the relevant minimum value of the reference test
database is about 12.
where
. /
𝑓 1.0 ∙ 𝜂 ∙ 𝜂 ∙ (19)
η2 is equal to 1.0 and 0.7 for good and poor bond casting position, respectively;
η3 = 1.0 for ϕ ≤ 32 mm.
In the following the differences between the proposed practical approach (i.e. equation (17)) and that of CEB-FIP
Model Code 90 are studied by analysing the ratio lbd/lbd,MC90 for c = 1.5.
Figures 2 and 3 show the influence of c/ for sd = 200 MPa on lbd/lbd,MC90 whilst figures 4 and 5 show the influence
of sd for fck = 25 MPa on lbd/lbd,MC90.
Figures 3 and 5 highlight that for PP the equation of CEB-FIP Model Code 90 significantly underestimates the
anchorage length. On the other side, for good bond casting position (see figures 2 and 4) this underestimation is
only relevant to low c/ (i.e. lower than 2) while for c/ > 2.0 the equation of CEB-FIP Model Code 90 is
conservative.
For citations, copying and referencing see Note on Front page! page 639
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
2.0 c/=
c/f=1 1.0
c/=
c/fi=1.5 1.5
lbd/lbd,MC90 c/= 2.0
c/fi=2
c/= 2.5
c/fi=2.5
1.0
0.5
0.0
0 10 20 30 40 50 60
fck [MPa]
Figure 2. Influence of c/ on lbd/lbd,MC90 in case of c = 1.5, sd = 200 MPa and GP.
5.5 c/=
c/f=1 1.0
lbd/lbd,MC90 c/=
c/fi=1.5 1.5
5.0
c/= 2.0
c/fi=2
4.5 c/= 2.5
c/fi=2.5
c/= 3.0
c/fi=3
4.0
3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
0 10 20 30 40 50 60
fck [MPa]
Figure 3. Influence of c/ on lbd/lbd,MC90 in case of c = 1.5, sd = 200 MPa and PP.
For citations, copying and referencing see Note on Front page! page 640
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
2.5 c/=
c/f=1 1.0
lbd/lbd,MC90 c/=
c/fi=1.5 1.5
c/= 2.0
c/fi=2
c/= 3.0
c/fi=3
1.5
1.0
0.5
0.0
0 50 100 150 200 250 300 350 400
sd [MPa]
5.0 c/=
c/f=1 1.0
lbd/lbd,MC90 c/=
c/fi=1.5 1.5
4.5
c/= 2.0
c/fi=2
3.0
2.5
2.0
1.5
1.0
0.5
0.0
0 50 100 150 200 250 300 350 400
sd [MPa]
References
Cairns, J., Feldman, L. (2018). Strength of laps and anchorages of plain bars. Structural Concrete, 19(6), 1782-
1791.
CEB-FIP Model Code 90. 1993, Lausanne.
EN 1990:2002. Eurocode: Basis of Structural Design.
EN 1992-1-1:2004. Eurocode 2: Design of Concrete Structures - Part 1-1: General Rules and Rules for Buildings.
Diamantidis, D., Holický, M., Sýkora, M. (2016). Reliability and risk acceptance criteria for civil engineering
structures. Transactions of the VŠB – Technical University of Ostrava, Civil Engineering Series, 16 (2),
1–10.
fib (2010). fib bulletin 52. Structural Concrete Textbook on behaviour, design and performance, Second edition
Volume 2: Basis of design. 2010. Lausanne: International Federation for Structural Concrete (fib).
JCSS (2001). Probabilistic Model Code. Lyngby, Denmark: Joint Committee on Structural Safety.
For citations, copying and referencing see Note on Front page! page 641
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
Mancini, G., Carbone, V.I., Bertagnoli, G., Gino, D. (2018). Reliability-based evaluation of bond strength for tensed
lapped joints and anchorages in new and existing reinforced concrete structures. Structural Concrete,19,
904–917.
Palmisano, F., Greco, R., Biasi, M., Tondolo, F., Cairns, J. (2020). Anchorage and laps of plain surface bars in
R.C. structures. Engineering Structures, Vol. 213, 110603/1-12.
Taerwe, R.L. (1993). Towards a consistent treatment of model uncertainties in reliability formats for concrete
structures. CEB Bulletin d’Information No 219, Safety and Performance concepts (pp. 5-61), Lausanne:
CEB.
For citations, copying and referencing see Note on Front page! page 642
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
27.03.2023
For citations, copying and referencing see Note on Front page! page 643
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
Introduction
This section provides the background to clause I.11.4.1(7) and details the derivation of parameters included in
table I.7. Further details are included in Palmisano et al. (2020a, 2021).
Figure 1. Reference STM for the evaluation of bond strength (fib 2014).
Figure 2. Reference STM for the evaluation of bond strength in case of concrete cover loss.
Figure 3. Definition of cs, cx, cy, corner bar (1), edge bar (2).
For citations, copying and referencing see Note on Front page! page 644
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
The lower limit of the concrete cover for the application of formulae 11.3 and I.26 of FprEN 1992-1-1:2023 without
any modification is chosen equal to for the following reasons:
to be consistent with the limit given in clause 6.5.2.3 of FprEN 1992-1-1:2023;
to be consistent with the minimum value of concrete cover in the test database for anchorage length of
plain bars assumed as reference for formula I.26 of prEN 1992-1-1:2023;
even if the minimum value of concrete cover c in the reference test database (fib 2014) for the anchorage
length of ribbed bars is 0.5, the number of tests performed with c = 0.5, is too low to have a statistical
validity; as stated in fib (2014), tests with c ≤ 0.5 were excluded from the statistical summary;
differently from other approaches present in the literature, the approach here proposed gives a general
procedure valid for all cases of low concrete cover thickness.
Tables 1 and 2 give the vales of k’lbs,c according to Palmisano et al. (2020a) for edge and corner bars, respectively,
as defined in figure 3.
𝑐
1 0.5 0.5
2𝜙
𝑐
2 0.5 0.5
𝜙
𝑐
3* 0.5 0.5
𝜙
For citations, copying and referencing see Note on Front page! page 645
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
𝑐
1 0.5 0.5
𝜙
𝑐
2 0.5 0.5
𝜙
3*
if cmin,xy/ < 1/3
,
0.1 0.7
𝑐
4 ** 0.5 0.5
𝜙
𝑐
5 ** 0.5 0.5
𝜙
For citations, copying and referencing see Note on Front page! page 646
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
The selected tests taken from Schenkel’s experimental program included a series of four lap splice tests. All
specimens had four pairs of 16 mm diameter ribbed bars, lapped in the same section, in bottom position, within
the constant moment zone (CMZ) of a four-point bending scheme. In these series, all beams had the same
dimensions and the bars were cast with a concrete cover equal to 20, 10, 5 or 0 mm.
A four-point bending scheme was also used in the two series of laps investigated by Regan and Reid (2009). The
first series included a pair of 25 mm corner bottom cast ribbed bars, lapped within the CMZ. In the second series,
there were two pairs of edge bottom cast bars. In the investigations, the concrete cover was varied between 33
to 0 mm; some cases with negative concrete equal to -8 mm were also tested. These last tests have been ignored
in the present note, because the influence of negative concrete cover (i.e. c/ < 0) on bond lies outside the scope
of this validation.
Cairns and Goodchild (2016) analysed four sets of tests. Each set comprised three individual specimens, one
with a concrete cover equal to 30 mm and two with concrete cover equal to 15 mm. Sets from 1 to 3 were
composed of two pairs of 25 mm diameter lapped bars. Set 4 was composed of two pairs of 16 mm diameter bars
lapped to 25 mm diameter bars. The latter test has been excluded from this study not only because relevant to
laps of different bar diameter (i.e. a situation, even if very common in practice, not consistent with the other tests
included in the reference database) but also because the concrete cover (i.e. 15 mm) was barely less than the
minimum bar diameter (i.e. 16 mm). Each set was relevant to bottom cast ribbed bars lapped, with different offsets,
in the CMZ of a four-point bending scheme.
Tests on plain bars have been taken from Darby et al. (2016) investigations. In particular, two series of beam-end
specimens, one with and the other without links, have been studied. Each series was composed of two specimens
of a 16 mm diameter bar in bottom cast position with a concrete cover of 37 or 0 mm.
To validate the equations proposed by Palmisano et al. (2020a), each test series has been considered individually.
Since some test results are in terms of fst (anchorage strength) while other are in terms of fb (bond strength), the
following equation is used to derive fb from fst:
𝑓 (1)
where:
c/ in formula (2) is relevant to that of the test where c/ > 1;
= 0.25 for ribbed bars according to the formulation of fstm included in fib (2014);
for plain bars = 0.77 applies according Palmisano et al. (2020b).
If in a set of investigations there is more than one test with c/ > 1, the mean value of the anchorage/lap strength
of the tests in that set having c/ > 1 is assumed as fst,c/ >1 in formula (2).
The parameter k is evaluated, by using the test results, with reference to both fst,c/ >1 and fst,c/ =1 according to the
following formulae (3) and (4).
,
𝑘 (3)
,
,
𝑘 , (4)
,
As above mentioned, the type of tests included in the reference database are relevant to case 2 of table 1 and to
case 1 of table 2; thus, the following formulation for kMOD according to Palmisano et al. (2020a) should be used:
𝑘 0.5 0.5 (5)
The influence of c/≤ 1 on kTEST, kTEST,rel and kMOD is plotted in the figures 4-6 for ribbed bars and in figure 7 for
plain bars.
Figures 4-7 clearly show that equation (5) is conservative for both ribbed and plain bars since the points relevant
to kTEST and kTEST,rel are always above the line of kMOD.
For citations, copying and referencing see Note on Front page! page 647
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
1.5 Cairns
and
Goodchild
(2016)
kTEST
Serie1
0.5 kTEST,rel
Serie2
kMOD
Serie3
0.0
0.0 0.2 0.4 0.6 0.8 1.0
c/φ
Figure 4. Influence of c/ on kTEST, kTEST,rel and kMOD for Cairns and Goodchild’s investigations (ribbed bars).
Regan
1.5 and Reid
(2009)
kTEST or kTEST, rel or kMOD
1.0
kTEST
Serie1
0.5 kTEST,rel
Serie2
kMOD
Serie3
0.0
0.0 0.2 0.4 0.6 0.8 1.0
c/φ
Figure 5. Influence of c/ on kTEST, kTEST,rel and kMOD for Regan and Reid’s investigations (ribbed bars).
For citations, copying and referencing see Note on Front page! page 648
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
Schenkel
1.5 (1998)
kTEST
Serie1
0.5 kTEST,rel
Serie2
kMOD
Serie3
0.0
0.0 0.2 0.4 0.6 0.8 1.0
c/φ
Figure 6. Influence of c/ on kTEST, kTEST,rel and kMOD for Schenkel’s investigations (ribbed bars).
Darby et
1.5 al. (2016)
kTEST or kTEST, rel or kMOD
1.0
kMOD 1
kMOD,
0.5 kTEST
kTEST
kTEST,rel
Ktest, rel 2
0.0
0.0 0.2 0.4 0.6 0.8 1.0
c/φ
Figure 7. Influence of c/on kTEST, kTEST,rel and kMOD for the investigations relevant to plain bars included in Darby
et al. (2016).
References
Cairns, J., Goodchild, C. (2016). Effect of reduced concrete cover on strength of lapped joints. Proceedings of the
Institution of Civil Engineers: Structures and Buildings, 169, 34-45.
Canbay, E., Frosch. R. (2005). Bond Strength of Lap-Spliced Bars. ACI Structural Journal, 102(4), 605-614.
Darby, A., Orr, J., Ibell, T. (2016). Half Joint Assessment Programme Research. Topic 1. Residual bond
characteristics of prestressing strand. Final Report. University of Bath, Department of Architecture and
Civil Engineering.
fib (2014). fib Bulletin 72. Bond and Anchorage of Embedded Reinforcement: Background to the fib Model Code
2010. Lausanne: International Federation for Structural Concrete (fib).
Metelli, G., Plizzari, G.A. (2014). Influence of the relative rib area on bond behaviour. Magazine of Concrete
Research, 66(6), 277-294.
For citations, copying and referencing see Note on Front page! page 649
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
Palmisano, F., Biasi, M., Greco, M., Marano, G.C. (2020a). Effect of low ribbing and concrete cover thickness on
the anchorage of bars in existing R.C. structures. Proc. of the IABSE 2020 Wroclaw Poland Symposium
– Synergy of Culture and Civil Engineering - History and Challenges, Wroclaw, Poland, 7-9 October 2020,
617-624.
Palmisano, F., Greco, R., Biasi, M., Tondolo, F., Cairns, J. (2020b). Anchorage and laps of plain surface bars in
R.C. structures”. Engineering Structures, Vol. 213, 110603/1-12.
Palmisano, F., Cairns, J., Menga, A. (2021). Anchorage/lap strength of bars in R.C. structures in case of low
concrete cover thickness. Proc. of the IABSE Congress Ghent 2021 – Structural Engineering for Future
Societal Needs, Ghent, Belgium, 22-24 September 2021, 1057-1065.
Regan, P.E. and Reid, I.L.K. (2009). Assessment of concrete structures affected by cover delamination. Part 1 –
Effect of bond loss. Graduate School in Concrete Structures – Fratelli Pesenti, Politecnico di Milano, Italy,
Studies and researches – V.29.
Schenkel, M. (1998). Zum Verbundverhalten von Bewehrung bei kleiner Betondeckung (On the bond behaviour
of reinforcement with little concrete cover), Report No. 237, Zurich: Swiss Federal Institute of Technology,
Institute of Structural Engineering IBK.
SIA (2003). SIA 262, Concrete structures. Zurich: Swiss Society of Engineers and Architects (SIA).
SIA (2011). SIA 269/2, Existing structures – concrete structures. Zurich: Swiss Society of Engineers and Architects
(SIA).
Zwicky, D. (2013). Bond and ductility: a theoretical study on the impact of construction details – part 2: structure-
specific features. Advances in Concrete Construction, 1(2), 137-149.
For citations, copying and referencing see Note on Front page! page 650
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
27.03.2023
For citations, copying and referencing see Note on Front page! page 651
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
I.11.2.(2) The minimum of bar spacing (𝑐 ) in 11.2(2) of FprEN 1992-1-1:2023 is a practical rule, the aim of which is
clearly explained in 11.2.(1). This limit of 𝑐 is not referred to bonding. In fact, considered the general model
Minimum bar spacing
for bond that is the basis of the anchorage length in FprEN 1992-1-1:2023 a limit equal to 𝑐 , should be also applied
to (i.e. for bonding should be not lower than the bar diameter). In this scenario I.11.2(2) has been added.
I.11.4.1(2) The minimum value for 𝑙 (i.e. 5𝜙) given in this clause is equal to the minimum value in the database
considered in fib bulletin 72 (see e.g. Fig. 3.8 of fib bulletin 72) that gives the expression for bond strength
Minimum value for lbd
assumed as basis of formula (11.3) of FprEN 1992-1-1:2023.
In particular, about 100 tests out of about 800 have a bond length lower than 15𝜙. However, as stated at page
14 of fib bulletin 72, tests with short bond length of less than 15𝜙 for laps or 10𝜙 for anchorages were excluded
from the statistical summary as unrepresentative of practice.
Clause I.11.4.2(2) includes the limit of 5𝜙 as the result of the discussion of Turin meeting (25th October 2018)
between TC 250/SC 2/SC2.T3, TC 250/SC 2/WG 1/AhG Detailing & fib TG 2.5. In that meeting, it appeared to
be a general consensus on the necessity to remove, in case of existing structures, practical limits of bond
lengths if there is evidence that the area where anchorage develops is uncracked.
I.11.4.2(1) The effect of bends and hooks has been taken from fib bulletin 72 (2014) par. 3.3. It can be used also for plain
bars according to par. 6.1.3.5 of MC 2010 (version of fib bulletin 65, 2012).
Anchorage of bars with
bends and hooks The effect of bar diameter and confinement has been neglected. Moreover, to have the design values, from
the mean value given in fib bulletin 72, by using the same approach described in the ‘Background document
to clauses I.11.4.1(3)-(5)’ (i.e. the statistical approach proposed by Taerwe 1993, ‘Towards a consistent
treatment of model uncertainties in reliability formats for concrete structures’, CEB Bulletin d’Information No.
219, Safety and Performance concepts, pp. 5-61, Lausanne: CEB), has been used by using the following
assumptions:
Mean value and coefficient of variation of model uncertainties for ribbed bars have been taken as 1.04
and 0.18, respectively, according to Mancini et al. 2018 (i.e. Mancini, G., Carbone, V.I., Bertagnoli, G.,
Gino, D. 2018, ‘Reliability-based evaluation of bond strength for tensed lapped joints and anchorages in
new and existing reinforced concrete structures’, Structural Concrete,19, 904–917).
Mean value and coefficient of variation of model uncertainties for plain bars in good bond casting position
have been taken as 1.00 and 0.18, respectively, according to the ‘Background document to clauses
I.11.4.1(3)-(5)’.
Mean value and coefficient of variation of model uncertainties for plain bars in poor bond casting position
have been taken as 1.00 and 0.36, respectively, according to the ‘Background document to clauses
I.11.4.1(3)-(5)’.
The first coefficient of formula of fib bulletin 72 has been multiplied by 0.8 and 0.5 for ribbed and plain
bars, respectively, to take account of poor bond casting condition. For ribbed bars this factor is constituent
with that of FprEN 1992-1-1:2023. For plain bars this factor is equal to the mean value of the ratio of the
anchorage strength of bars in poor bond casting position to that of bars in good bond casting position in
the reference test database referred in the ‘Background document to clauses I.11.4.1(3)-(5)’.
The contribution of c has been evaluated by using the same procedure followed in the ‘Background
document to clauses I.11.4.1(3)-(5)’ for the design anchorage/lap length of plain bars.
It is worth noting that 1 in formula I.30 is lower than 0.5 for plain bars because of the high coefficient of variation
of model uncertainties for plain bars in poor bond casting position.
I.12.(1) Clause 12.1(6) of FprEN 1992-1-1:2023 is a requirement (“shall” clause); it obliges the designer to consider a
member as plain or lightly reinforced concrete member if the longitudinal reinforcement is lower than 𝐴 , .
Members with less
longitudinal Many existing structures do not fulfil the minimum requirements on detailing of new structures. If these are
reinforcement than considered as lightly reinforced concrete structures most of the checks will not be satisfied. Thus, even if
As,min 12.1(6) of FprEN 1992-1-1:2023 is suitable for a new structure, it is very restrictive for an existing one. This is
why I.12(1) gives the possibility to ignore 12.1(6) of FprEN 1992-1-1:2023 for existing structures.
For citations, copying and referencing see Note on Front page! page 652
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
27.03.2023
For citations, copying and referencing see Note on Front page! page 653
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
Introduction
This section gives additional clauses and background on deteriorated concrete structures. This section is intended
only to provide information outlining what are the obvious signs and what should be taken into account in
assessment.
This section does not intend to provide models for predicting degradation of structures for the different
mechanisms that prevail.
This section is mainly based on the input received from CEN/TC 250/SC 2/WG 1/TG 3 (draft V15 February 2019).
A list of recent articles on corrosion is included in the last paragraph.
Reinforcement corrosion
General
(1) Structural damage in structures with corroded reinforcement, to be considered at ULS verifications, are due
to:
- reinforcement deterioration;
- concrete deterioration;
- bond deterioration.
The main consequences of corrosion are:
- loss of bar cross-section;
- loss of steel ductility;
- cracking of concrete cover with the consequential loss in concrete cross section;
- loss of bond between reinforcement and concrete.
(2) The relevant parameters for the assessment are:
- Mechanical strength of concrete.
- Crack width, crack pattern and delaminated areas.
- Concrete cover thickness.
- Mechanical properties of reinforcement.
- Residual cross-sectional area of reinforcement.
Note: In the case of presence of cracks parallel to the reinforcement, typical of the corrosion process, their pattern
is a qualitative indication of the degree of damage and of the affected surface. Its actual value is a key parameter
for the assessment. The cracks and delamination are necessary to estimate the residual concrete section.
Although the reinforcement maximum strength may remain unaffected, the corrosion may induce a reduction of
steel ductility and the embrittlement of the reinforcement core when the loss in cross section is higher than 15-
20%. This has to be taken into account in the case of seismic risks and other dynamic loads and when considering
non-linear or plastic analysis.
For citations, copying and referencing see Note on Front page! page 654
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
initial small diameters of stirrups and therefore relevant large cross-section loss;
lower cover for stirrups leading to earlier initiation of corrosion compared to main bars.
- Spalling of cover and reduction of the concrete compression zone.
- Loss of bond of longitudinal reinforcement due to corrosion.
- Induced stresses in stirrups due to corrosion of the main bars.
In members with corrosion at the tensile side and without stirrups, the deterioration of bond at longitudinal bars
can reduce shear strength because of the relevant reduction of dowelling action.
When evaluating the shear resistance of members requiring design shear reinforcement, a further limitation of the
minimal inclination of the compression field should be considered due to bond deterioration.
(5) In the assessment of the ultimate axial load of a column:
- the concrete section should be reduced in case of spalling;
- if the stirrups have been broken due to the corrosion, a reduction in the longitudinal bars subjected to
compression due to risk of buckling should be taken into account.
Loss of bond
(1) The effects of corrosion in hardened concrete differ from those associated with corrosion prior to concreting.
Small amounts of corrosion, up to the level required to induce longitudinal cracking, do not cause loss of bond
resistance, and can even augment bond strength by a modest degree, particularly where the bar is in a ‘poor’
bond casting position.
At greater levels of corrosion, residual bond strength is strongly influenced by the degree of confinement provided
by transverse reinforcement in the form of links and by the surrounding structure. Transverse pressure from
support reactions augments bond. Links play a valuable role in maintaining residual strength of anchorages and
lapped joints.
(2) The residual capacity of anchorages and lapped splices should be checked at the ultimate limit state at
locations of high reinforcement stress where longitudinal cracking develops. Away from anchorages and laps, a
substantial loss of bond may be tolerated without ultimate strength being affected.
(3) The magnitude of the reduction in residual bond strength is highly dependent on the confinement to the bar
and is also affected by concrete quality and environment. It may be taken into account by increasing the design
value of the anchorage and lap length.
Note: The increase of lbd can be evaluated by using Table 6.1-4 of fib MC 2010, for plain and ribbed bars, or
section F.2.3 of Contecvect Manual (‘CONTECVET. A validated Users Manual for assessing the residual service
life of concrete structures. Manual for assessing corrosion-affected concrete structures’, EC Innovation
Programme IN30902I, preparation of document led by Geocisa and Torroja Institue, 2000), for ribbed bars.
However, in the literature there are proposals by different researchers that can be taken as reference. A new
proposal will be included in the forthcoming fib MC 2020, chapter 20.
(4) In addition to (3), the effect on bond of concrete cover spalling should be considered and it may be taken into
account by using the reduced cover thickness in the formulae that give lbd.
For citations, copying and referencing see Note on Front page! page 655
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
Frost attack
(1) Frost attack is produced due to the fact that ice has a higher volume than liquid water. Not all the pore water
solution freezes at the same temperature because as smaller are the pore size lower temperature is needed to
produce ice. The expansion can induce surface delamination of concrete or an internal microcracking when the
whole section is frozen. Its progress depends on the degree of saturation of the concrete and on the temperature
and its cycling regime. Freezing temperatures, if they are maintained constant, do not necessarily impair the
concrete which has shown to last at these temperatures maintaining its bearing properties.
(2) The progress of the process has not been systematically recorded. It can be monitored by following the amount
of delaminated concrete or the crack pattern and width evolution.
(3) The structural consequences are:
- the external frost attack can affect the performance similarly as sulphate attack due to:
loss in cross section;
loss of strength due to the loss in aggregate/concrete bond;
- the internal frost attack:
reduces (locally) the strength due to the internal cracking;
causes loss on modulus of elasticity;
causes loss in strength;
causes loss in steel/concrete bond.
For citations, copying and referencing see Note on Front page! page 656
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
For citations, copying and referencing see Note on Front page! page 657
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex I Background to FprEN 1992-1-1:2023
For citations, copying and referencing see Note on Front page! page 658
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
For citations, copying and referencing see Note on Front page! page 659
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex J Background to FprEN 1992-1-1:2023
Authors
Konrad Zilch Technische Universität München (DE)
Craig Giaccio Arcadis Consulting (UK)
Roland Niedermeier Technische Universität München (DE)
Wolfgang Finckh Ostbayerische Technische Hochschule Regensburg (DE)
Anthony Darby Bath University (UK)
Renata Kotynia Lodz University of Technology (PL)
Eva Oller Universitat Politècnica de Catalunya BarcelonaTech (ESP)
Thorsten Leusmann Technische Universität Braunschweig (DE)
Rolf Alex Deutsches Institut für Bautechnik (DIBt) (DE)
Anett Ignatiadis Deutscher Ausschuss für Stahlbeton (DAfStb) (DE)
Introduction
This document has been prepared in order to explain the proposed provisions for strengthening of existing
concrete structures with CFRP in FprEN 1992-1-1, Annex J. The main basis for the proposed provisions were
national guidelines [1], [2], fib-bulletin 90 [3] and national approvals.
CEN/TC250/SC2/WG1/TG1 21.12.2022
For citations, copying and referencing see Note on Front page! page 660
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex J Background to FprEN 1992-1-1:2023
J.3 General
Annex J contains rules for strengthening of existing concrete structures with carbon fibres. Adhesively bonded
reinforcement (ABR) is used as umbrella term for different application methods and products. Externally bonded
reinforcement (EBR) is glued to the prepared concrete surface. Near surface mounted reinforcement (NSM) is
applied in slots cut in the concrete cover. Figure BD-J.1.1 illustrates the terms, the application methods and the
corresponding products as developed for this Eurocode. Figure BD-J.1.2 shows some examples of strengthening
methods considered in annex J.
CF sheet/pre‐
CFRP strips impregnated CFRP strips CFRP bars
laminates
Figure BD-J.1.2: Examples for strengthening: a) Ilustration EBR and NSM, b) Strengthening a beam for
bending, c) Strengthening of a column by wrapping
The design of FRP strengthened concrete members follow in principle the mechanics of reinforced concrete.
However, some peculiarities have to be taken into account and may lead to different limit states as explained in
the following chapters.
For example, the bond behaviour of strips applied in EBR and NSM shows a distinct difference (comp. Figure
BD-J.1.3).
Figure BD-J.1.3: Principle of the bond force transfer of adhesively bonded CFRP-strips [4]
In NSM the total force capacity of a CFRP strip can be transferred by bond on a single crack in an acceptable
bond length (as with ribbed reinforcing steel). The strength of the adhesive will be governing the design.
In EBR the total force of a strip cannot be transferred by bond on a single crack. To get acceptable utilization of
the strip capacity strength increase of the force over intermediate crack elements has to be considered! As the
tensile strength of the adhesive is in general much larger than the tensile surface strength of the concrete it is
assumed in the models for EBR that the concrete will fail first and the tension strength of the concrete is governing
bond design. So, the surface tension strength of the concrete will be a very important parameter in design.
Furthermore, it has to be considered that the strength parameters of the CFRP and adhesive may depend on the
temperature and environmental conditions.
Background
The safety concept is based on the regulations of prEN 1990:2020 [47]. According to section 8.3.5 the design
value of resistance 𝑅 for a specific design situation should be calculated from the equation (BD-J.4.1)
corresponding to equation (8.18) in [47]:
∙
𝑅 ∙𝑅 ;𝑎 ;∑𝐹 (BD-J.4.1)
where
𝛾 is a partial factor associated with the uncertainty of the resistance model, and for geometric
deviations, if these are not modelled explicitly (Remark: numerical values may be obtained
from e.g. [48]);
𝑅 … denotes the output of the resistance model;
𝜂 is a conversion factor accounting for scale effects, effects of moisture and temperature,
effects of ageing of materials, and any other relevant parameters;
𝑋 represents the characteristic values of material or product properties;
𝛾 is a partial factor for a material property accounting for:
- unfavourable deviation of the material or product properties from their characteristic
values;
- the random part of the conversion factor 𝜂;
𝑎 denotes the design values of geometrical property;
𝐹 denotes design values of actions used in the assessment of 𝐸 ;
When applying partial factors to resistance, its design value 𝑅 should be calculated from Formula (BD-J.4.3)
where partial factors 𝛾 and 𝛾 are combined into a single partial factor 𝛾 for resistance:
𝛾 𝛾 𝛾 ∙𝛾 (BD-J.4.2)
∙ ; ;∑
𝑅 ∙ (BD-J.4.3)
Where relevant for the application-specific failure modes, the conversion factors are explicitly stated. They have
to be determined by experimental testing in accordance with an European Technical Product Specification unless
otherwise noted. The effects of geometric deviations are limited by design specifications, e.g. regarding the
number of layers of bonded reinforcement or the slot geometry, and are taken into account in the partial factor
𝛾 with a specific magnitude depending on the procedure. For bonded reinforcement, a thickness of adhesive
layer between 1 and 5 mm is assumed.
For the determination of the partial safety factors given in Table J.1, products were considered that were also
used in the structural tests, on the basis of which the design approaches in sections J.8 to J.11 were determined.
In accordance with the specifications in section J.5 for the design, the axial tensile strength is considered. For the
prefabricated products CFRP-strips and bars, data from the factory production control and the external continuous
surveillance were made available for two representative product groups A and B, which had been fabricated with
different production processes.
It must be taken into account that the product groups considered, as it is the case for pre-stressing steel strands,
are not only due to various production parameters, such as fibre type or fibre content, but are also based on a
targeted allocation to manufacturer-specific product classes on the basis of the mechanical properties determined
in the factory production control (FPC). The data from the FPC therefore only represent a certain range of
parameters, so that a suitable distribution function cannot be determined. Furthermore, it must be taken into
account that the data from the FPC may also contain values that are not applicable to the design situation of the
component due to the type of fracture, but are nevertheless accepted within the scope of the FPC if the strength
is sufficient. The data of the external surveillance, on the other hand, representing the entire parameter range of
the respective manufacturer-specific product class, only contain tests with applicable fracture types and show a
normal distribution of the strength.
Table BD-J.4.1: Statistical data of tensile strength from FPC and the external continuous surveillance for two
representative product groups A and B of prefabricated CFRP strips (statistical data courtesy
of different German cerfication bodies)
Product
Data source Parameter A B
Number of numerical test results 𝑛 185 617
Factory production control Coefficient of variation 𝑉 0,07 0,05
Kolmogorov–Smirnov statistic for a normal
𝐷 0,09 0,11
distribution function
by notified Number of numerical test results 𝑛 77 122
product Coefficient of variation 𝑉 0,06 0,09
Audit-testing in certification Kolmogorov–Smirnov statistic for a normal
the context of body 𝐷 0,06 0,07
distribution function
continuous by external Number of numerical test results 𝑛 61
surveillance product Coefficient of variation 𝑉 0,07 -
certification Kolmogorov–Smirnov statistic for a normal
body 𝐷 0,12
distribution function
For CF sheets applied by the wet lay up method, the partial factor for tensile strength is based on data collected
by Occhiuzzi et al. [5]. They found a lognormal distribution of tensile strength for the single and three layer
application of a typical CF sheet based on the evaluation of tests carried out at eight university laboratories and
give the following parameters.
Table BD-J.4.2: Statistical data of tensile strength of CF sheets according to Occhiuzzi et al. [5]
Number of layers
Parameter 1 3
Number of numerical test results 𝑛 270 288
Mean value of logarithmised tensile strengths 𝜇 8,20 8,19
Standard deviation of logarithmised tensile strengths 𝜎 0,15 0,14
Kolmogorov–Smirnov statistic for a lognormal distribution function 𝐷 0,061 0,058
Coefficient of variation of tensile strengths 𝑉 0,15 0,14
If the maximum coefficients of variation included in tables BD-J.4.1 and BD-J.4.2 are considered representative
of the population, the following partial factors 𝛾 for the tensile strength of ABR are obtained.
For the CFRP strips, the partial factor 𝛾 is determined for a normal distribution with the reliability index 𝛽 3,8
for residential and office buildings and the sensitivity factor (FORM) 𝛼 0,8 for resistance as follows:
, ∙ , ∙ ,
𝛾 1,17 (BD-J.4.4)
∙ ∙ , ∙ , ∙ ,
For the CF sheets applied in wet lay-up procedure, the partial factor 𝛾 is determined for the lognormal distribution:
∙ ∙ , ∙ , ∙ , ∙ , , ∙ ,
𝛾 𝑒 𝑒 1,23 (BD-J.4.5)
The application of these values for the design implies a corresponding confirmation of the applied coefficients of
variation and distribution functions on the basis of experimental investigations of the tensile strength and the
constancy of performance according to a European Assessment Document.
The partial safety factor for resistance 𝛾 for persistent and transient design situations is a combination of 𝛾 and
the partial factor 𝛾 associated with the uncertainty of the resistance model and for geometric deviations:
𝛾 𝛾 ∙𝛾 (BD-J.4.6)
The index “f” is used in prEN1992-1-1 for the material (FRP), while according to EN 1990 𝛾 is a partial factor that
takes account of unfavourable deviation of an action from its representative value.
Depending on the application technique, the partial factor 𝛾 has different magnitudes for the persistent and
transient design situation, the following partial safety factors are obtained, which are determined for the other
design situations analogously to prEN 1992-1-1, Section 4.3.3, Table 4.3. Although the same models are used
for both applications, a higher 𝛾 is applicable for CF sheets amongst others because of higher degree of
geometric deviations with this application technique.
Table BD-J.4.3: Derivation of partial factors for materials
Design situation Partial factor CFRP strips and bars In-situ lay-up CF sheets
𝛾 1,17 1,23
𝛾 1,10 1,15
Persistent and transient 1,30 1,40
Accidental 1,10 1,15
𝛾
Serviceability 1,00 1,00
Fatigue 1,30 1,40
The safety concept for bond is based on design assisted by testing according to EN 1990, Annex D (cf. section
J.15) by assessment via the characteristic value with 𝛾 𝛾 . The conversion factors 𝜂 ∙ 𝑘 and 𝑘 for the
concrete strengths are considered according to FprEN 1992-1-1.
For the determination of the partial safety factor for bond 𝛾 , taken from fib Bulletin 90 [3], failure of near-surface
concrete layers or the adhesive failure was assumed. For the persistent and transient design situation, 𝛾
corresponds to the partial safety factor for the concrete bearing component 𝛾 for design of shear at interfaces
according to Section 8.2.6 (5) of the main part of FprEN 1992-1-1.
Table BD-J.4.4: Partial factors for bond strength
Partial factor 𝛾 for
Design situation
bond strength
Persistent and transient 1,50
Accidental 1,15
Serviceability 1,00
Fatigue 1,50
The design concept for near surface mounted CFRP reinforcement distinguishes between concrete failure and
adhesive failure (see section 11). For reasons of simplification, the same safety factor was used here. A closer
look could possibly apply a lower safety factor to the adhesive failure due to its probably lower scatter. However,
the long-term strength has to be considered.
The limits provided in this section have been selected to reflect to testing ranges used in the development of
J.5.1 (3)
relationships used in this annex. They also correlate with products widely available on the market.
Establishing a minimum value for the characteristic tensile strength of adhesive ensures that the basic assumption
of design of externally bonded reinforcement, namely that the concrete layer close to the surface will fail, has
been satisfied.
The material properties are product-specific and have to be given by the supplier. There are some standard test
J.5.2
procedures for the CFRP in ISO 10406 but not for all relevant properties and not for the entire strengthening
system. Further some of the design parameters may not be directly derived from the test results. In this case or
where test procedures are not standardized yet, recommended values are given in the corresponding sections of
Annex J.
In the product standard EN 1504, which provides with definitions, requirements, quality control and evaluation of
conformity, Part 2 handles surface protection systems for concrete and Part 4 give rules for structural bonding.
It has to be considered that all properties of the single components (e.g. CFRP strips, adhesive) have to be
ensured to be valid in the entire strengthening system.
For further information see “Requirements to Materials – CFRP reinforcement for strengthening” (Background
document to Annex J.15) [42].
According to the Construction Products Regulation (Regulation (EU) No 305/2011 (CPR)) on EU level test
procedures for the essential characteristics of construction products may be included in two types of technical
specifications for construction products:
- European harmonized product standards or
- European Assessment Documents (EAD).
At the time of this writing, none of the two types of technical specification was available for CFRP Strengthening
Systems. EADs are under preparation. That’s why [42] contains not only methods for determining the design
parameters for Annex J, but also the descriptions of possible test procedures, where necessary.
The factor 𝜂 is a reduction factor applied to the tensile strength of the ABR CFRP. It considers long-term effects
J.5.3 (1) and (2)
as creep under sustained load and influences of the alkaline environment and the temperature.
For further information see “Requirements to Materials – CFRP reinforcement for strengthening” (Background
document to Annex J.15) [42].
In approvals, where tests had been performed values between 0,65 and 1,0 can be found. That’s why the
recommended value 0,7 has been chosen. In approvals where no tests have been performed lower values may
be found.
The permissible environmental conditions as well as the resulting measures will be regulated in the European
J.6.3
Assessment Documents (or a subsequent execution standard) of the strengthening System. For further
information see Background document to Annex J.15 [42].
Presently, the supplied system dictates the exposure classification. This document was prepared on the basis
that the designer will specify an exposure classification as part of a design, along with the design life to enable a
suitable product to be supplied for the strengthening application.
The concrete cover must be measured (in execution) so that the reinforcement is not endangered when cutting
the slots. Also, the filling and protection must both ensure fire performance and bond of both existing reinforcement
and NSM reinforcement in slots is acceptable.
The flexural capacity of the strengthened member is assessed by means of traditional cross-sectional analysis,
J.8.1.1
based on strain compatibility assuming full composite action between the CFRP and the concrete substrate,
horizontal equilibrium of forces and moment equilibrium. The load level during strengthening is considered through
the acting strains in the cross section when the moment has come for ABR application.
Debonding checks are undertaken to validate this approach.
J.8.1.2 General
The provisions for the confining of concrete columns in Annex J consider only the confining effect due to the
adhesively bonded CFRP. The confining effect of the internal reinforcement has not been considered for reason
of simplification.
An approach for considering the internal reinforcement can be found in [1], [43], [44].
The application of FprEN 1992-1-1, Section 8.1.4 in addition to Annex J, J.8.1.2 is not on the safe side because
of different behaviour of the internal steel reinforcement and the strengthening material.
The diameter of a circular column, D, or effective diameter of a rectangular or square column, Deq ≥ 150 mm;
The majority of testing has been carried out on relatively small-scale columns, the vast majority with a
cross-sectional dimension of 150 mm. Some larger specimens have been tested to demonstrate
applicability of models and behaviour at more realistic scale (Darby et al. [7] for rectangular columns,
Youssef [52]). The models have therefore been verified for columns of effective diameter 150 mm and
above.
The first order eccentricity satisfies the condition e0/Deq < 0,20;
While there is some debate around an appropriate limit concerning eccentricity of loading, Song et al.
[6] show that for FRP confined square columns tested with eccentricities greater than e/D of 0,25, the
capacity increase is negligible compared to the unconfined equivalent. While there are a few other
studies on eccentrically loaded columns, there is limited data to support a less onerous condition or
modifications to the model.
The basis of the confinement model assumes confinement around the full column perimeter, full bond
to the substrate and failure governed by rupture of the FRP. However, where significantly eccentric
loads are applied a tensile stress in the FRP does not exist around the full perimeter, altering
confinement within the concrete. For more extreme eccentricities debonding of the FRP can occur due
to shear stresses at the FRP-concrete interface (particularly in rectangular columns), reducing
confinement stiffness. There is also a possibility that FRP rupture does not govern failure, but rather,
lateral deformations become gross, P-Delta effects start to influence behavior and steel can rupture in
tension (Darby et al. [7]). To avoid these issues, an e/D value of 0,20 ensures that the whole cross
section is in compression (varying linearly going from zero axial strain on one side to maximum axial
compressive strain on the opposite face) and thus some significant level of confinement exists around
the whole column perimeter.
model (e.g. Darby et al. [7], Binici [55]) is much simpler to apply, is conservative, and gives perfectly acceptable
results, transitioning at a strain of 0,00175.
A rectangular stress block is not suitable for confined concrete, due to the significant variation in additional
strength that can be provided.
The value of 𝜀 can be taken as 0,006 unless more accurate information is available.
Not only is ultimate strength enhanced by confinement, but ultimate strain is also significantly increased. The
ultimate compressive strain for FRP confined concrete varies significantly and models to predict this strain are
highly empirical with a large scatter. It is therefore suggested that a lower bound (based on the database of
experimental results used to validate the models) of 0,006 be used for the ultimate strain for the purposes of
developing a conservative stress-strain model.
The provisions in FprEN 1992-1-1, Annex J for the increase in compressive strength of concrete due to
J.8.1.2 (4)
confinement from FRP are adapted from fib-90 (2019) [3] as outlined below.
The aim of the amendments to fib-90 [3] provisions is to implement an increment increase in compressive strength
outlined in clause 8.1.4.
Rectangular Columns:
According to fib-90 [3], Equation (6.55):
, , ,
1 3,3 ∙ ∙𝛼 ∙ ∗
∙ 𝑓𝑜𝑟 ∙𝛼 ∙ ∗
∙ 0,07 (BD-J.8.1)
, ,
1 𝑓𝑜𝑟 .𝛼 . ∗
. 0,07 (BD-J.8.2)
Where bf, sf, and bf are defined in Figure J.2 (b), while h, b and rc are defined in Figure J.2 (a), in FprEN 1992-1-1.
For ease of use, it was noted that the 2nd – 4th components of af will not always be used in practice. These terms
only apply to helical or intermittently wrapped systems, therefore the terms was re-written as follows:
𝛼 𝑘 ∙𝑘 (BD-J.8.5)
where:
∙ ∙
𝑘 1 (BD-J.8.6)
∙ ∙
𝑘 1 1 ∙ (BD-J.8.7)
∙ ∙
Replacing fib 90 [3] symbols with FprEN1992-1-1 symbols and re-arranging, the following is obtained:
The term kh is not explicitly incorporated into the formulae as it is the exception rather than the rule. However
FprEN1992-1-1 allows 𝑓 to be factored by kh if helical/intermittent wrapping is used.
The intention in the provisions in FprEN1992-1-1 was to retain the same approach as Section 8.1.4 in the main
text, the term fcd was introduced such that:
𝑓 , 𝑓 ∆𝑓 (BD-J.8.10)
Substituting formula (BD-J.8.5) into formulae (BD-J.8.3) and (BD.J.8.4), and simplifying, the following Formulae
is used in FprEN 1992-1-1.
The value of 3,3 in the fib 90, equation (BD-J.8.1), has been substituted by kcc. Comparing the model with the
database of tests, a value of kcc of 1,5 is recommended as a safe upper limit. Due to the complexity of the model
and numerous factors involved, there is considerable scatter of results from the model compared to the measured
results as shown in the figure below, resulting in a high mean value for the ratio of measured/predicted strength
increase and a high standard deviation. A value of kcc = 1,5 ensures that there is 95% confidence in the
measured/predicted strength increase > 1 (with material partial safety factors set to unity) for tests which meet
the criteria in equation (BD-J.8.11).
0,5 ∙ ∙ 2 for 𝑟 50 mm
𝑘 (BD-J.8.13)
0,5 for 𝑟 50 mm
accounts for the reduction in strength observed for FRP wrapped around corners. This is consistent with fib 90 [3]
and is used for both confinement and shear strengthening. Corner radius above 50 mm is rare in tests, due to the
need to maintain concrete cover to internal reinforcement, but the upper limit on kr is 0,5. This is consistent with
the strength reduction factor for the FRP for confinement of circular columns, based on a lower limit observed in
tests.
Circular Columns
Similar substitutions outlined for rectangular columns were applied to arrive at provisions for circular columns in
FprEN 1992-1-1. Resulting in the following formulae:
∙ ∙ , ∙
∆𝑓 0 for 0,07 (BD-J.8.14)
∙
∙ ∙ ∙ , ∙
∆𝑓 𝑘cc ∙ ∙ 0,5 ∙ 𝑓 for 0,07 (BD-J.8.15)
∙
The term kr is omitted as it is always taken as 0,5 (i.e. radius > 50mm) and, as such, is directly included in the
formulae.
The term kh for helical/intermittent FRP confinement of circular columns is not provided in fib 90, but has been
derived on the same basis as the equation for kh given for rectangular columns and is applied in the same way.
𝑘 1 ∙ (BD-J.8.16)
∙
Due to the relative simplicity of the model, there is considerably less scatter in results when comparing the
measured strength increase from the experimental database with the predictions from the model, as shown in the
figure below.
As a result of the lower mean and standard deviation for this model, the factor kcc = 2,5 ensures that there is 95%
confidence that the measured/predicted strength increase > 1 (with material partial safety factors set to unity) for
tests which meet the criteria in formula (BD-J.8.15). This is lower than the 3,3 suggested in fib 90 [3] but gives a
safe upper bound to the value.
Shear strengthening for columns contains the only fully closed systems (Fig. BD-J.8.1).
(a) (b)
Figure BD-J.8.1: Shear strengthening of columns with
(a) closed – continuous application of the strengthening system and
(b) closed – discrete application of the strengthening system. [3]
Shear strengthening is required if the average shear strength over the cross-section Ed (Eq. (8.18)) is larger than
Rd,c (Eq. (8.27)) in case the unstrengthened element does not have shear reinforcement, or when the average
shear strength Ed is larger than Rd,sy (Eq. (8.42)) or Eq. (8.44) is not fulfilled if the unstrengthened element has
shear reinforcement. That is, in cases where the internal steel reinforcement is inadequate following the detailing
rules of EN 1992-1-1 the difference may be covered using ABR CFRP in accordance with J.8.2.3. The method
described in Annex J allows superposition with the basic section shear capacity and is similar to that given in [3].
𝜏 , 𝜏 𝜏 , (BD-J.8.17)
Where:
, 𝜌 𝑓 cot 𝜃 cot sin (BD-J.8.18)
𝜌 (BD-J.8.19)
∙
Unless more rigorous analysis is undertaken, should be taken as 45 degrees for the calculation of Rd and Rd,f.
If the structure has not the shear strengthening, at least a fraction equal to 3 = 0,5 of the necessary shear
reinforcement should be covered by stirrups. If this condition is not satisfied, the difference in the shear strength
shall be covered by the externally applied CFRP reinforcement.
For the effective thickness of CF sheets Eq. (J.1) has to be considered, so the reduced effectiveness of a jacket
with many layers is taken into account.
In members with flexural strengthening in accordance with J.8.1 and with Ed > Rd,c , the shear verification of
J.8.2.3 is applicable.
Additional Information on shear strengthening in relation to flexural debonding can be found in J.11.1.2.3.
0,5 2 𝑅 50 𝑚𝑚
(BD-J.8.23)
0,5 𝑅 50 𝑚𝑚
𝑓 can be obtained from Eq. (J.2)
Figure BD-J.8.2: Strips crossed by the shear crack. (le corresponds to lbf in FprEN1992-1-1)
Figure BD-J.8.3: Generic bilinear b-s constitutive law for bond in EBR FRP systems.
(a) for ℎ /sin 𝑙 , , and 𝑙 , , 𝑠 / cot𝜃 cot sin ℎ /sin , if all strips are intersected by the
shear crack the bond length is ≥ 𝑙 , , and 𝑓 is calculated according to:
𝑓 , (BD-J.8.25)
(b) for ℎ /sin 𝑙 , , and 𝑠 / cot𝜃 cot sin 𝑙 , , , if all strips are intersected by the shear crack,
the bond length is < 𝑙 , , , and 𝑓 is calculated according to:
𝑓 , 1 1 (BD-J.8.26)
(c) for ℎ /𝑠𝑖𝑛𝛼 𝑙 , , and 𝑠 / 𝑐𝑜𝑡𝜃 𝑐𝑜𝑡𝛼 𝑠𝑖𝑛𝛼 ℎ /𝑠𝑖𝑛𝛼, if all strips are intersected by the shear crack,
the bond length < 𝑙 , , , and 𝑓 is calculated according to:
𝑓 , (BD-J.8.27)
, ,
⎧ ∙ 2 𝑓𝑜𝑟 𝑠 𝑙 , ,
𝑓 𝑠 (BD-J.8.29)
⎨ 𝑓𝑜𝑟 𝑠 𝑙 , ,
⎩
The characteristic values of the shear strength 𝜏 and the ultimate slip 𝑠 can be taken from Table BD-J.8.1.
Table BD-J.8.1 - Parameters of the bilinear 𝜏 - s constitutive law for bond.
Type τb1 s1 (mm) s0 (mm)
Mean value 0.53 fcm fctm 0.0063 0.21
CFRP strips
5% characteristic value 0.37 fcm fctm - 0.20
Mean value 0.72 fcm fctm 0.0107 0.24
CFRP sheets
5% characteristic value 0.44 fcm fctm - 0.23
⁄
For ℎ /𝑠𝑖𝑛𝛼 𝑙 , , : 𝑓 , (BD-J.8.31)
, ,
with 𝑘 ≤ 0,9.
If no anchors are used 𝑘 𝑓 , /𝑓 , .
(a) (b)
Figure BD-J.8.4: Anchorage of three-sided CFRPs:
(a) Use of spike anchors in combination with CFRP sheets;
At the serviceability limit state, the strain in the strengthening system must be limited for reasons of the durability
J.9
of the bond. Here is good experience with a strain limitation for rare load combinations of f < 2 mm/m for members
with existing reinforcing steel of fyk = 500 MPa. This corresponds to the more general proposal for a stress limit in
dependence of the existing reinforcing steel strength of:
𝜎 0,8𝑓 (BD-J.9.1)
The intention behind limiting strain resp. stress for rare load combinations is to prevent significant damage to the
bond of the externally bonded reinforcement. Furthermore, yielding of the reinforcing steel under rare load
combinations is ruled out in the main part, likewise to prevent high bond stresses and irreversible deformations of
the structure.
Limiting the reinforcing steel stresses alone is not sufficient, as members without reinforcing steel reinforcement
can also be strengthened.
Figure BD-J.10.1: Goodman projection for determining the load range S0,i
Figure BD-J.10.2: S-N curve and experimental data from Bizindavyi [9], Carloni [10], Budelmann [11], Leusmann
[12], Dai [13] and Ferrier [14]
Table BD-J.10.2 : Experimental data from Carloni [10], Budelmann [11], Leusmann [12], Dai [13] and Ferrier
[14].
Description
When checking fatigue for non-static loads, the Annex J can be used to verify the bond of flexural strengthening
in the form of near-surface-mounted CFRP strips. As the carbon fibres exhibit virtually no signs of fatigue, only
the bond needs to be checked for fatigue when using CFRP strips. Besides the fatigue of the strengthening
system, the concrete, reinforcing steel and prestressing steel must also be checked according to EN 1992.
In contrast to externally bonded CFRP strips, however, there is no comprehensive analysis concept available for
near-surface-mounted strips. Owing to the low number of fatigue tests involving near-surface-mounted CFRP
strips (see [16]), a quasi-fatigue strength analysis is the only option here. With so few test results available, it is
not possible to specify an S-N curve for near-surface-mounted reinforcement. And as an S-N curve is unavailable,
it is not possible to extrapolate for a number of load cycles greater than that given in the test results. Therefore,
the analysis can only assume sufficient fatigue resistance for max. 2ˑ106 load cycles. Design methods for numbers
of load cycles > 2ˑ106 are not covered in the DAfStb guideline [1].
In this analysis, adequate resistance to fatigue for near-surface-mounted CFRP strips may be assumed for up to
2 ꞏ 106 load cycles provided the end anchorage force for a frequent cyclic action to EN 1992 and taking into
account the “shift rule”, does not exceed the value 0,6 𝑓 and the strip stress range does not exceed a value
given by Eq. (J.32). The strip thickness 𝑡 in mm should be used here so that the result is an admissible stress
range in N/mm².
Background
The analysis of the fatigue resistance of the bond of near surface mounted CFRP strips is based on a small
number of tests described in [16] and [3]. Hence it is not possible to provide a S-N-curve for the bond of near
surface mounted CFRP strips. Owing to the lack of a S-N-curve it is not possible to extrapolate to a greater number
of cycles than investigated in the tests. In this analysis it can therefore only be assumed that adequate fatigue
resistance is provided up to 2 ∙ 106 load cycles ).
The few fatigue tests were carried out on individual products. It should therefore be additionally checked for each
product that the following conditions are fulfilled:
The upper load should be less than 60% of the composite failure load at the point of adhesive cohesion
failure (transition to friction plateau):
𝐹, , 0,6 ⋅ 𝑓 ⋅ 𝑏 ⋅ 𝑡 with 𝑓 according to Formula (J.56)
The amplitude should be less than 15% of the composite failure load at the point of adhesive cohesion
failure (transition to friction plateau):
Δσ 0,15 ⋅ 𝑓 with 𝑓 according to Formula (J.56)
The fctm,surf formulation in FprEN 1992-1-1, Eq. (J.34) is based on the lower bound of the possible formulations.
J.11.1.1.1
Other possible formulations taking into consideration the dependence of concreting position on the surface tensile
strength by the factor k are possible [18], [19], [20]. Alternatively, usual ranges are shown in figure BD-J.11.3 for
the different concreting positions so that the designer gets some feeling for possible problems. The values
included in this diagram are based on the arithmetic mean from at least four pull-off tests according to EN 1542
which had been performed on concrete specimens made from globular limestone aggregates from the Munich
area.
The Simplified method is based on the following conservative simplification of the refined method in Chapter
J.11.1.1.3.
Equation J.38 can be reformulated as follows
⋅ ⋅
𝑓 ⋅𝛽 ⋅𝑓 , (BD-J.11.1)
Where:
2 1 if 𝑙 𝑙 , ,
𝛽 , , , , (BD-J.11.2)
1 if 𝑙 𝑙 , ,
Equation J.40 can be reformulated with the default values of J.11.1.1.3 𝜏 0,37 ⋅ 1,0 ⋅ 𝑓 , ⋅𝑓 and ∙
𝑠 0,2 (ksys = 1,0) as follows
𝑓 , 0.272 ⋅ 𝑓 , ⋅𝑓 (BD-J.11.3)
The front part of the equation (BD-J.11.1) can be conservative reformulated with ktc = 0.85 , kcc=0.7 and 𝜂
/
0.93
.
𝑓 ⋅𝛽 ⋅𝑓 , (BD-J.11.4)
.
𝑓 ⋅ 𝛽 ⋅ 0.272 ⋅ 𝑓 , ⋅𝑓 (BD-J.11.5)
Equation J.39 can be reformulated with the default values of J.11.1.1.3 𝜏 0,37 ⋅ 1,0 ⋅ 𝑓 , ⋅𝑓 and ∙
𝑠 0,2 (ksys = 1,0) as follows
∙ ∙ ,
𝑙 , ∙ (BD-J.11.6)
, ⋅ , ⋅
𝐸𝑓 ∙𝑡𝑓
𝑙 , 𝑙𝑏𝑓,𝑚𝑎𝑥 1,5 ∙ (BD-J.1175)
𝑓𝑐𝑡𝑚,𝑠𝑢𝑟𝑓 ⋅𝑓𝑐𝑚
Figure BD-J.11.2 Determination of the bond forces at the end anchorage using the differential equation for bond
and the bilinear bond approach, based on [18], [21], [22]
The basic anchorage resistance - CFRP to concrete corresponds to the tensile stress that can be anchored by
the EBR CFRP at a single crack. The refined method for determining this tensile stress is based on a slip-bond
stress relationship that can be derived from force and slip measurements in bond tests with EBR CFRP. The slip-
bond stress relationship of EBR CFRP can be idealised using a bilinear approach as shown in Figure BD-J.11.2.
Up the maximum bond stress f1, a linear relationship between the slip and the bond stress applies. After
exceeding the maximum bond stress, progressive softening occurs until complete debonding of the EBR CFRP
occurs when the critical slip sf0 is reached. If the friction occurring in practical applications at even larger slip values
after debonding is neglected, the idealised bilinear bond approach provides a good approximate
phenomenological description of the complex processes in the composite system. The size of the area under the
bilinear slip bond stress curve is of the unit Nmm/mm2 and corresponds to an areic energy Gf, the bond fracture
energy which must be expended to completely detach a unit area. From this, it can be seen that a fracture energy
limitation must be considered when anchoring EBR CFRP at a single crack. An increase in the tensile stress
prevailing in the crack cross-section is only possible up to a certain maximum effective bond length lbf,max. Larger
bond lengths do not lead to higher tensile stresses that can be anchored at the single crack due to the fracture-
energy limitation. Therefore, additional cracks are required to build up higher tensile stresses, since the bond
fracture energy defined by the bilinear bond approach can then be activated at each intermediate crack element
located between adjacent cracks.
However, for the differential equation that describes the relationship between relative displacement (slip) sf x and
stress f x , neglecting the concrete deformations for low concrete stress c, no analytical solution suitable for
manual calculation can be found for the section-wise defined bilinear bond approach with the corresponding
boundary and transition conditions at a single crack that allows a determination of the tensile stress that can be
anchored as a function of the bond length lbf. In contrast, a simplified solution of the differential equation can be
obtained by the substitute approach of a linear, plastically softening slip-composite stress relation that results from
the bilinear approach with sf1 = 0 and has the same magnitude of composite fracture energy.
1 ∙𝜏 𝑤ℎ𝑒𝑟𝑒 𝑠 𝑠
𝜏 𝑠 (BD-J.11.10)
𝑤ℎ𝑒𝑟𝑒 𝑠 𝑠
0
For this simplified slip-bond stress relationship, the tensile stress f that can be anchored at the single crack may
be determined as a function of the bond length x = lbf:
∙ ∙
𝜎 𝑙 ∙ sin ∙𝑙 (BD-J.11.11)
∙ ∙
On basis of this equation the maximum anchorable tensile stress fbf,max is determined for:
∙ ∙
𝑙 , ∙ . (BD-J.11.12)
∙ ∙
𝑓 , (BD-J.11.13)
However, since the linear elastic branch of the bilinear bond approach was neglected for the approximate solution,
equation (BD-J.11.12) determines anchorage lengths that are too short compared to calculations that consider
both the linear elastic and the plastic softening branch. For a design approach, equation (BD-J.11.12) must
therefore be multiplied by the factor 4⁄ 𝜋 ∙ 𝑘 , , which leads to equation (BD-J.11.14) and (J.33).
∙ ∙
𝑙 , ∙ . (BD-J.11.14)
,
If the sinusoidal course of the tensile stress 𝜎 𝑙 that can be anchored as a function of the bond length 𝑙 is
replaced by a quadratic parabola, which approximately also describes the exact course on the basis of the bilinear
bond approach with sufficient accuracy, the following equation is obtained
𝑓 , ∙ ∙ 2 𝑤ℎ𝑒𝑟𝑒 𝑙 𝑙 ,
, ,
𝜎 𝑙 (BD-J.11.15)
𝑓 , 𝑤ℎ𝑒𝑟𝑒 𝑙 𝑙 ,
For the design at ultimate limit state, the characteristic values of the variables are applied and the equations are
amended with the partial safety factor BA for bond of adhesively bonded CFRP reinforcement. In addition, with
regard to the concrete strengths, the factor cc to account for the difference between the undisturbed compressive
strength of a cylinder and the effective compressive strength that can be developed in the structural member as
well as the factors ktc and ktt considering the effect of high sustained loads and of time of loading on are added to
the equations.
Figure BD-J.11.4: Examples of failure modes for a reinforced concrete beam with flexural strengthening in the
form of externally bonded reinforcement [23]
First of all, modes of failure related to the function of the CFRP strip can be added to those familiar from
conventional reinforced concrete:
Failure of concrete in compression zone
Yielding of internal reinforcement followed by concrete failure
Yielding of internal reinforcement followed by failure of the adhesively bonded strip
Shear failure
Yielding of externally bonded steel plate
Besides these modes of failure well known from conventional reinforced concrete and relatively easy to describe,
there are other modes specific to strengthening measures with externally bonded reinforcement. The first of these
that should be mentioned is concrete cover separation failure, where the concrete cover becomes detached at
the end of a strip. This occurs due to the additional, vertical offset between shear link and strip because the tensile
stresses from the strip cannot be tied back to the compression zone of the beam. This mode of failure therefore
corresponds to a horizontal shear failure in the area between the externally bonded reinforcement and the internal
reinforcement. (End Cover Separation as described in J.11.1.2.4)
The bond between the adhesive and the concrete often fails when using externally bonded reinforcement. In such
a bond failure the layers of concrete near the surface break away once the tensile strength of the concrete has
been exceeded. Owing to the only moderate tensile strength of the layers of concrete near the surface, following
local debonding of externally bonded reinforcement, the result is mostly a total failure of the bond between the
externally bonded reinforcement and the concrete as the load rises further because the forces involved cannot
normally be carried by any remaining areas of intact bonding (unzipping effect) This behaviour means it is
necessary to consider the bond of externally bonded reinforcement very carefully.
In conventional reinforced concrete usually, bond checks are done by end anchorage verifications, which are
based on bond values from pull-out tests. If we introduce such verifications in similar form to structural elements
with externally bonded reinforcement, the full tensile strength of the CFRP-strip cannot be anchored, because
after a specific length, the bond forces cannot be increased anymore. (See also Figure BD-J.11.5). However, the
full-scale tests have shown that much higher forces in the externally bonded reinforcement are reached at the
place of the maximum bending moment, as it would be possible looking only at the end anchorage.
On CFRP strips with their high tensile strength, the sole consideration of the end anchorage would be highly
uneconomical. Thus, the bond forces of the CFRP-strips must be transferred at the point where the forces occur.
For this reason, two areas are distinguished for bond verifications, the end anchorage zone and the rest of the
structural element.
At the end anchorage zone, the forces of the CFRP-strip must be anchored, which occur at outermost bending
crack. The bearable bond forces at the end anchorage zone can be determined by so-called idealized end
anchorage tests, in which the externally bonded reinforcement will be pulled off in the longitudinal direction.
(End Anchorage as described in J.11.1.2.2)
In the remaining area of the structural element, the bond force can be transferred by elements, which are
separated by bending cracks, the so-called intermediate crack element.
On such an intermediate crack there is always a basic force in the strip at lower stressed crack and at the higher
stressed crack there is this basic force plus an additional force. This additional force must be transferred by bond
to the structural element. (Intermediate Crack Debonding as described in J.11.1.2.3)
Figure BD-J.11.5: Principle of the bond force transfer of externally bonded CFRP-strips [24]
In addition to the failure modes described above a premature uncoupling of the externally bonded reinforcement
at shear cracks due to high shear force may occur. (End Anchorage as described in J.11.1.2.2)
The analysis at the flexural crack closest to the point of contraflexure represents the standard case. In
this case the FRP force acting at this flexural crack must be lower than the resistance of the cross-
section taking into account the “shift rule”
Where strengthening is done in a small region of the member, the CFRP strengthening shall a distance
of 𝑙bf + ℎ beyond the area of need.
strip force which does not depend on the base force in the strip. The change in the tensile force of the strip is
verified along the member. As the analysis does not depend on the base force in the strip, it can be performed on
superimposed load combinations, unlike the detailed analysis. Several boundary conditions were specified in [26]
with the aim of simplifying the detailed analysis. In addition to the boundary conditions listed in Equation (J.50), it
was also necessary to specify an ultimate strain of 10 mm/m for the strips which must not be exceeded at any
point on the member. It is therefore specified in FprEN1992-1-1, J.11.1.2.3 (3) that the ultimate strain and the
design tensile strength of the strips must not be exceeded.
In [1],[3],[22],[23],[34] the equation for the simplified analysis is given below:
, ⋅ . ⋅√ , ⋅ , ⋅ ⋅
𝛥𝐹 ⋅𝑏 (BD-J.11.16)
, ⋅ . ⋅ , , ⋅ . ⋅ , ⋅ ,
𝛥𝑓 (BD-J.11.17)
⋅
,
. ⋅ , ⋅ ⋅ , ⋅ . ⋅ , . ⋅ ⋅ . ⋅ , ⋅ ,
𝛥𝑓 (BD-J.11.18)
⋅
, , ,
, ⋅ , ⋅ ⋅ , ⋅ ⋅ ⋅
𝛥𝑓 (BD-J.11.19)
, , ,
𝛥𝑓 (BD-J.11.20)
For durability reasons the adhesive bond resistance is reduced as in Equation (J.38)
⋅ ⋅ ∙ , , ,
𝛥𝑓 (BD-J.11.21)
with:
√𝑠
𝛥𝑓 , 0,85 ⋅ 𝑘 , ⋅ 𝑓 ⋅𝑓 , ⋅
𝑡
, 𝑠 ,
𝛥𝑓 , 𝑓 ⋅
𝑡
𝜅 𝑠 ,
𝛥𝑓 , ⋅
ℎ 𝑡
Crack Pattern
The proposal is based on the DAfStb Guideline “Strengthening of Concrete Members with Adhesively Bonded
Reinforcement” [1] and the fib-bulletin 90 [3].
The crack spacing is determined on the safe side over 1,5 times the insertion length of the inner reinforcement.
Thus, an initial crack pattern is assumed there and only the already existing reinforcing steel is considered.
In Equation (J.45) 𝜙 is the diameter of a conventional (unstressed) steel reinforcement bar and 𝑛 , is the
number of bars of equal diameter in the tension zone of a cracked section. The summation refer to addition of the
term 𝜙 ⋅ 𝑛 , ⋅ 𝑓 for each individual bar in the tension zone of a crack concrete section.
More accurate and complex calculations are required for prestressed concrete.
supports, which are frequently encountered in precast concrete construction. The acting forces can be determined
approximately with a truss model, as shown in figure BD-J.11.7.
For the usual components of building construction, k = 1+ (200/d)0,5 can be set approximately to 1,5 on the safe
side. From this follows approximately when neglecting the normal force:
,
1
𝑉Rd,cfE 0,75 ⋅ 1 19,6 ⋅ , ⋅ 0,1 ⋅ 1,5 ⋅ 100𝜌 𝑓ck ⋅ 𝑏w ⋅ 𝑑 (BD-J.11.27)
Where a shear strap is necessary, it should be designed according to Figure BD-J.11.8 for the design value of
the acting tensile force according to Eq. (J.52).
Figure BD-J.11.8 Calculation of force in shear wrapping due to tensile force in bending reinforcement
When using externally bonded reinforcement, an approach that lies on the safe side is to design the force at the
end strap for the maximum force in the flexural strengthening that can be accommodated by the end anchorage
according to Eq. (J.40). This is because a larger force cannot occur in the strip at the end strap except when the
shear strap counts towards increasing the bond force of the flexural strengthening. The force in the shear strap
is, however, superposed on the maximum force that can be accommodated at the end anchorage according to
J.11.1.1.3.
Figure BD-J.11.9 Movement of the crack edges due to the shear force [21]
Beyond a certain level of shear stress, adhesively bonded stirrups are also required in order to ensure that the
tensile forces from the adhesively bonded reinforcement can also be transferred to the flexural compression zone
of the member by the truss shown in Figure BD-J.11.10. The limit 02 according to DIN 1045:1988 [36] has proven
to be a suitable parameter in such cases (cf. Z-36.12-73:2009 [37]). Equation (J.54) describes a similar limit [33].
Figure BD-J.11.10 Mechanism for transferring the tensile forces from the adhesively bonded reinforcement to
the flexural compression zone of the member using the truss model [21]
The design value of the applied shear force of the externally bonded shear reinforcement is determined as the
maximum of two components as it is in existing approvals. The first component in Equation (J.55) is obtained by
distributing the design value of the total applied shear force over the elastic stiffnesses and the second component
is the difference between the design value of the total applied shear force and the design value of the shear
resistance of the steel reinforcing links.
J.11.1.3 Basic anchorage resistance – CFRP to concrete for NSM CFRP strengthening
The proposal for the basic anchorage resistance – CFRP to concrete for NSM CFRP strengthening is based on
the DAfStb Guideline “Strengthening of Concrete Members with Adhesively Bonded Reinforcement” [1] and the
fib-bulletin 90 [3].
Essentially, the concept proposed by Blaschko [16], [17] is used for verifying the bond. In this concept it is
assumed that the CFRP strip makes a full contribution and there is good composite action between strip and
concrete, with the strength of the adhesive usually governing this composite action. As the composite action very
effective, the full tensile strength of the CFRP strip can be anchored within a very short length – similar to
conventional steel reinforcing bars. It is therefore sufficient to check the end anchorage at the point at which the
strip is no longer required for the load-carrying capacity, very similar to anchoring steel reinforcing bars. This
concept to describe the way in which strips in slits work has proved worthwhile over the past 10 years in the former
national technical approvals and was therefore included in the DAfStb guideline [1].
As with conventional reinforced concrete construction with curtailed reinforcing bars, checking the bond requires
verification of the curtailment taking into account the end anchorage of the CFRP strip. This involves verifying that
the design value of the member resistance is greater than the design value of the acting internal forces in the
strengthened condition for every cross-section of the strengthened member. The partial tensile forces assigned
to the lines of reinforcement can be determined in a simplified way by assuming a planar strain distribution. Figure
BD-J.11.11 provides an overview of the curtailment verification.
reduction factor to take account of long-term effects of the adhesive bond, bA, was introduced in Equation (J.59).
A value of 0,5 was proposed for this factor although a different value may be specified in the system approval.
Failure will occur in the concrete if the concrete strength is very low, i.e. below C20/25. In this case, a concrete
wedge will break away on either one side or both sides of the strip, with the base of the wedge in the concrete
member corresponding to the base of the slit. This can be calculated with Equation (J.61). In this Equation this
bond strength is calculated from the square root of the concrete compressive strength and a calibration factor kbck.
The system coefficient for the bond failure of the concrete can be taken from the national technical approval for
the system. Tests carried out at the Technische Universität München (Technical University of Munich) established
a characteristic value kbck = 4,5 [40], [41].
The factors bC is also introduced into Eq. (J.59) to take account of the long-term durability behaviour of the
materials involved. As these are also coefficients specific to particular products, they can again be obtained from
the national technical approvals.
Range of application
The concept in J.11.1.3 is based on bond test with CFRP-strips. For this concept the CFRP-strips width shall be
between 10 mm and 30 mm and the CFRP-strips thickness shall be between 1 mm and 3 mm.
For other CFRP elements such as round or square bars the bond behaviour is different. So for the determination
of FbfRd further considerations are necessary. This may be bond tests to determine FbfRd directly or to recalibrate
the factor 0,6 in equation (J.60) and the factor 4,5 in equation (J.61).
The limits for the centre-to-centre spacing are based on experiences. The limit of sf ≤ 3 times the slab thickness
≤ 400 mm is in line with the provisions for reinforcing steel
The slot width and depth limits ensure that there is 1 to 3 mm of adhesive between the NSM bar or strip and the
concrete. This ensures that there is full bond around the perimeter of the slot. The 3 mm upper limit is consistent
with adhesive thickness limit for external FRP plate bonding and the range between 1 mm and 3 mm is in line
with testing which has been carried out on NSM flexural strengthening of beams.
The centre to centre spacing ensures that the width of concrete between slots is at least equal to the length of the
sides of each slot so that there is not a propensity for concrete cover separation across the width of the beam in
preference to debonding around the concrete slot perimeter. This also prevents overlapping of confining stress
zones around each NSM bar, maximising bond stresses ([50],[51].
The minimum edge distance equal to 4 times the depth of the slot is based on findings of Hassan and Rizkalla
[50], for round bars, and Rashid, et al. [51] for strips. This is to prevent detachment of the concrete cover along
the edges of the beam, and the NSM reinforcement being able to carry the same force as a bar/strip far from the
edges.
The NSM technique can be used for the flexural and shear strengthening of deficient reinforced concrete members
for the following reasons: better bond characteristics; activation of controlled debonding that increases the ductility
of strengthened members; mechanical response to make the structure stiffer under serviceability loads; limited
damage by peeling-off failure due to flexural cracks; excellent fatigue resistance; reduced site work (due to not
necessary plaster removal, unevenness and irregularities of the concrete surface and removal of the weak laitance
layer on the concrete surface); easier anchorage to adjacent members to prevent debonding, where the maximum
moments typically occur at the ends of the member – flexural strengthening of columns); protection from
mechanical damage, accidental impact and vandalism; this aspect makes this technology particularly suitable for
the strengthening of negative moment regions, unchanged aesthetics [3].
The most commonly used CFRP strips have a thickness of 1,4 – 3,0 mm and a width of 15 – 30 mm, while FRP
bars of different diameters from 2 – 20 mm, which are bonded into the concrete grooves. As it was mentioned in
[45] two of the most efficient parameters guaranteeing the best bond behaviour are: the width of the strip and the
strips installation as deepest as possible into the slits providing the largest pull-out load and energy dissipation
(Fig. BD-J.12.1a). The research carried out by [46] indicated very promising conclusion that cutting of the bottom
horizontal arm of the steel stirrups in order to apply CFRP strips of an width larger than the concrete cover (Fig.
BD-J.12.1b) does not influence bending and shear capacity, because the bottom arms of the stirrups are safety
anchored. This solution in Fig. BD-J.12.1b) is not covered by Annex J. Another important issue mentioned in [45]
pointed an importance to reduce the number of NSM strips that make an increase in the distance between CFRP
strips and effects the interaction between steel reinforcement and strips by reducing the stiffness under the
longitudinal steel bars (Fig. BD-J.12.1c).
CEN/TC250/SC2/WG1/TG1 21.12.2022
For citations, copying and referencing see Note on Front page! page 690
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex J Background to FprEN 1992-1-1:2023
of concrete members with adhesively bonded reinforcement” with Examples], Schriftenreihe des
DAfStb No. 595, Beuth, Berlin, 2013.
[22] Zilch, K., Niedermeier, R., Finckh, W.: Praxisgerechte Bemessungsansätze für das wirtschaftliche
Verstärken von Betonbauteilen mit geklebter Bewehrung – Verbundtragfähigkeit unter statischer
Belastung [practical design methods for economic strengthening of concrete members with externally
bonded reinforcement – bond strength under static loading]. Schriftenreihe des DAfStb No. 592, Beuth,
Berlin, 2012.
[23] Zilch, K.; Finckh, W.; Niedermeier, R.: Strengthening of Concrete Structures with Adhesive Bonded
Reinforcement, Design and Dimensioning of CFRP Laminates and Steel Plates, Ernst und Sohn
Verlag, 2014, ISBN: 978-3-433-03086-8
[24] Finckh, W.; Zilch, K.: Strengthening and Rehabilitation of Reinforced Concrete Slabs with Carbon-Fiber
Reinforced Polymers Using a Refined Bond Model, Computer-Aided Civil and Infrastructure
Engineering 27 (5), pp. 333-346, 2012.
[25] Zehetmaier, G.: Zusammenwirken einbetonierter Bewehrung mit Klebearmierung bei verstärkten
Betonbauteilen [interaction of embedded reinforcement and externally bonded reinforcement in
strengthened concrete members]. Dissertation, Munich TU, Department of Concrete Structures, 2006
[26] Zehetmaier, G.; Zilch, K.: Interaction between internal bars and externally FRP Reinforcement in RC
Members, Proceedings of the sixth international Symposium on FRP reinforcement for concrete
structures (FRPRCS-6), pp.397-406, Singapore, 2003
[27] Husemann, U.: Erhöhung der Verbundtragfähigkeit von nachträglich aufgeklebten Lamellen durch
Bügelumschließungen [increasing the bond strength of retrofitted externally bonded strips by means of
shear wrapping]. Dissertation, Braunschweig TU. Institute of Building Materials, Concrete Construction
& Fire Protection, 2009
[28] Husemann, U., Budelmann, H: Increase of the Bond Capacitiy of Externally Bonded CFRP-Plates on
RC-Structures Due to Self-Induced Contact Pressure, In: Proceedings of the 9th International
Symposium on Fibre Reinforced Polymers in Reinforced Concrete Structures (FRPRCS 9), Sydney,
Australia, 2009
[29] Z-36.12-86: Bausatz StoCretec zum Verstärken von Stahl- und Spannbetonbauteilen durch schubfest
aufgeklebte CFK-Lamellen nach der DAfStb-Verstärkungs-Richtlinie, Deutsches Institut für
Bautechnik; 2021
[30] Jansze, W.: Strengthening of reinforced concrete members in bending by externally bonded steel
plates. Design for beam shear and plate anchorage. Dissertation, Delft University of Technology, 1997.
[31] DIN EN 1992-1-1 (2011): Eurocode 2: Design of concrete structures - Part 1-1: General rules and rules
for buildings.
[32] DIN EN 1992-1-1/NA: National Annex – Nationally determined parameters – Eurocode 2: Design of
concrete structures - Part 1-1: General rules and rules for buildings.
[33] Zilch, K., Niedermeier, R., Finckh, W.: Praxisgerechte Bemessungsansätze für das wirtschaftliche
Verstärken von Betonbauteilen mit geklebter Bewehrung – Querkrafttragfähigkeit [practical design
methods for economic strengthening of concrete members with externally bonded reinforcement –
shear strength]. Schriftenreihe des DAfStb No. 594, Beuth, Berlin, 2012.
[34] Finckh, W.: Einfluss bauteilspezifischer Effekte auf die Bemessung von mit CFK-Lamellen verstärkten
Stahlbetonbauteilen [influence of member-specific effects on the design of RC members strengthened
with CFRP strips]. Dissertation, Munich TU, Department of Concrete Structures, 2012.
[35] Finckh, W.; Zilch, K.: Influence of shear crack offsets on the bond behavior of EBR at the intermediate
crack element. In: The 6th International Conference on FRP Composites in Civil Engineering - CICE
2012, Rom, 2012
[36] DIN 1045: Structural use of concrete – Design and construction. Deutsches Institut für Normung. Beuth,
Berlin, 1988.
[37] Z-36.12-73: Verstärken von Stahlbetonbauteilen durch in Schlitze verklebte Kohlefaserlamellen
Carboplus nach DIN 1045-1:2008-08 [strengthening of RC members with near-surface-mounted
Carboplus carbon fibre strips to DIN 1045-1]. Deutsches Institut für Bautechnik, 2009.
[38] Borchert, K.: Verbundverhalten von Klebebewehrung unter Betriebsbedingungen [bond behaviour of
externally bonded reinforcement under operating conditions]. Schriftenreihe des DAfStb No. 575,
Beuth, Berlin, 2009.
[39] Zilch, K.; Borchert, K.: Bond behaviour of NSM FRP strips in service, structural concrete (9), pp.127-
142, 2008
[40] Finckh, W.; Niedermeier, R.: Bericht über Ausziehversuche an Probekörpern aus Normalbeton
niedriger Festigkeit mit in Schlitze verklebten CFK-Lamellen, 2012
[41] Zilch, K.; Finckh, W.: Gutachterliche Stellungnahme zur Eignung von in Schlitze verklebten S&P-
Lamellen zur Verstärkung von niedrigfesten Betonen, 2013
CEN/TC250/SC2/WG1/TG1 21.12.2022
For citations, copying and referencing see Note on Front page! page 691
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
Annex J Background to FprEN 1992-1-1:2023
[42] Alex, R.; Niedermeier, R.; Finckh, W.; Leusmann, T.: Requirements to Materials – CFRP reinforcement
for strengthening (Background document to Annex J.15), December 2022.
[43] Niedermeier, R.: Verstärkung von Stahlbetondruckgliedern durch Umschnürung. Habilitationschrift,
Technische Universität München, 2009.
[44] Niedermeier, R.: German Design Approach for Concrete Columns Strengthened with CFRP
Confinement. Proceedings of the 7th International Conference on FRP Composites in Civil Engineering
(CICE 2014), Paper 050, Vancouver, BC, Canada, 2014
[45] Barros, J.; Kotynia, R.: Possibilities and challenges of NSM for the flexural strengthening of RC
structures, Fourth International Conference on FRP Composites in Civil Engineering CICE2008, 2008,
Zurich, Switzerland.
[46] Kotynia, R.: Analysis of reinforced concrete beams strengthened with near surface mounted FRP
reinforcement. Arch. Civil Mech. Eng., 2006, LII(2), 305–17.
[47] prEN 1990:2020: Eurocode – Basis of structural and geotechnical design
[48] fédération internationale du béton (fib): fib Model Code for Concrete Structures 2010, Ernst & Sohn,
2013.
[49] Bilotta, A., Ceroni, F., Nigro, E. And Pecce, M. (2011): Design by testing procedure of debonding load
for RC elements strengthened with EBR FRP materials, in Proceedings of 10th FRPRCS International
Symposium, ACI SP-275 Fiber-Reinforced Polymer Reinforcement for Concrete Structures, R. Sen, R.
Seracino, C. Shield and W. Gold (eds), Tampa, Florida.
[50] Hassan, T. and Rizkalla, S. (2004), Bond mechanism of near-surface-mounted fiber-reinforced polymer
bars for flexural strengthening of concrete structures’, ACI Struct J, 101 (6), pp. 830-839.
[51] Rashid, R.; Ohelers, D.J. and Seracino, R. (2008), IC Debonding of FRP NSM and EB Retrofitted
Concrete: Plate and Cover Interaction Tests, J. Compos. Constr., 12(2), pp. 160-167.
[52] Youssef, M. N.; Mosallam, A. S. and Feng; M. Q. (2006): "Experimental Investigation on Large-Scale
FRP-Confined Axial Members," In: Proceedings of the International Conference on Civil Engineering
Infrastructure Systems (CEIS 2006) American University of Beirut(AUB), Beirut-Lebanon, June 12-14.
[53] Lam, L.; Teng, J. G. (2003a): Design-oriented stress-strain model for FRP-confined concrete.
Construction and Building Materials, 17, 471-489.
[54] Lam, L.; Teng, J. G. (2003b): Design-oriented stress-strain model for FRP-confined concrete in
rectangular columns. Journal of Reinforced Plastics and Composites, 22(13), 1149-1186.
[55] Binici, B. (2008): Design of FRPs in circular bridge column retrofits for ductility enhancenment.
Engineering Structures, Volume 30, Issue 3, pp. 766-776.
CEN/TC250/SC2/WG1/TG1 21.12.2022
For citations, copying and referencing see Note on Front page! page 692
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
For citations, copying and referencing see Note on Front page! page 693
CEN/TC 250/SC 2 N2087
CEN/TC 250/SC 2/WG 1 N 1314
J.15 Background to FprEN 1992-1-1:2023
Figure 2.1: Principle sketches of reinforced concrete sections with subsequent strengthening by externally
bonded CFRP strips
The product-specific design parameters for kits of EBR CFRP included in Annex J of FprEN 1992-1-1:2023 (E)
are the following:
1)
determined by test procedure
2)
declared by the manufacturer and confirmed by test procedure
3)
declared by the manufacturer
The number of tests given in Annex BD-A, clauses A.2.1 to A.2.3 is the minimum number which is necessary.
Each applicant can expand it according to their needs.
2.3.1 The creep reduction factor ηf of CFRP strips derived from the tests according to Annex BD-A,
A.2.2.5 and A.2.2.6
The factor 𝜂 is the minimum of the factors 𝑅 from the test A.2.2.5 and 𝑅 from the test A.2.2.6
(𝜂 Min 𝑅 , 𝑅 . It is given separately for all alkaline solutions used in A.2.2.5 and A.2.2.6 (pH 9 and/or 11
and/or 13,7).
2.3.2 Parameters of the bilinear bond approach ksys,b1, ksys,b2 and ksys,b3 derived from the tests
according to Annex BD-A, A.2.3.5
According to the studies by Kuntz [2] and Holzenkämpfer [3], the bilinear approach for the bond law according to
Figure 2.2 is well suited for the evaluation of the bond tests according to A.2.3.5. Therefore, only the determining
parameters of this approach have to be determined for each individual test by the evaluation.
These are the maximum bond stress 𝜏 , the associated elastic limit deformation 𝑠 and the limit value of the
relative displacement 𝑠 .
The following procedure is used for this:
The differential formula for the bond of externally bonded reinforcement given in the Figure 2.3 can be completely
solved for the two branches of the bilinear bond approach. With an adjustment to the boundary conditions of the
bond test specimen according to Figure A.15, the following formulae are obtained, which describe the relationship
between the relative displacement 𝑠 at the load-side beginning of the bond length 𝑙 and the tensile force 𝐹 . For
𝑠 𝑠 , only elastic bond stresses occur within the total bond length 𝑙 . After exceeding the elastic limit
displacement (𝑠 𝑠 ), additional plastically softened bond areas occur.
𝐹 𝐸 ∙ 𝑏 ∙ 𝑡 ∙ 𝑠 ∙ 𝜔 ∙ tanh 𝜔 ∙ 𝑙 for 𝑠 𝑠
∙ ∙ ∙
𝐹 𝐸 ∙𝑏 ∙𝑡 ∙𝜔∙𝜆∙ for 𝑠 𝑠
∙ ∙
With:
𝜏
𝜔
𝑠 ∙𝐸 ∙𝑡
𝑠
𝜆
𝑠 𝑠
The length 𝑎 of the elastically deformed bond region enters into the second formula and can be determined with
the following expression. Because of the simultaneous occurrence of hyperbolic and trigonometric functions, 𝑎
cannot be determined in complete form but only iteratively.
𝜆 𝑠 𝑠
tanh 𝜔 ∙ 𝑎 ∙ 𝑠 𝑠 ∙ cot 𝜔 ∙ 𝜆 ∙ 𝑙 𝑎
𝑠 sin 𝜔 ∙ 𝜆 ∙ 𝑙 𝑎
The parameters 𝑠 , 𝜏 , and 𝑠 of the bilinear composite approach (cf. Figure 2.2) are determined using the least
squares method on the experimentally determined 𝑠 -𝜀 curves (cf. Figure 2.3) for each of the two sides of a
specimen. In view of the scope of the calculations to be performed, a limited number of supporting points is
recommended for the computational determination of the parameters.
𝜏 𝑓 𝑓 ∙𝑓 , 𝑘 , ∙ 0,37 ∙ 𝑓 ∙𝑓 ,
𝑘 , ∙ 0,37 ∙ 𝑓 ∙𝑓 , ∙𝑘 , ∙ 0,2
𝐺 𝑓 𝑓 ∙𝑓 ,
2
𝑠 𝑘 , ∙ 0,2
The region before the maximum bond length is described in formula (J.38) via a parabola. However, the exact
solution from the bilinear composite approach is not exactly parabolic.
After the test interpretation, a better agreement of the approximation from the parabolic approach with the test
values can be achieved with the factor 𝑘 , .
For many known reinforcement systems with CFRP strips and cold-curing, quartz sand-filled epoxy resin
adhesives, 𝑘 , , 𝑘 , and 𝑘 , are equal 1,0.
2.3.3 Reduction factor for fatigue derived from the tests according to Annex BD-A, A.2.3.6
The fatigue reduction factor 𝛼 used in formula (J.27) in the draft standard FprEN 1992-1-1:2023 (E), Annex J
contains information for an S-N curve, which is verified by means of fatigue tests. The tests are performed
according to the test procedure in Annex BD-A, Chapter A.2.3.6. The following load levels are checked:
Load level Lower load 𝐹bLi Upper load 𝐹bLi 1st test Upper load 𝐹bLi 2nd test
A 0,15 𝐹 0,45 𝐹 0,49 𝐹
B 0,3 𝐹 0,55 𝐹 0,58 𝐹
C 0,45 𝐹 0,64 𝐹 0,67 𝐹
D 0,6 𝐹 0,74 𝐹 0,76 𝐹
The load levels are characterised by the average value of the anchorage failure load 𝐹 determined for the same
configuration of specimen under A.2.3.5. During the test the number of load cycles is counted until a separated
length of 30 mm is reached, see Chapter A.2.3.6.
If at least 2 million load cycles are achieved at all load levels A to D in the first test and at least 90.000 load cycles
in the second test, the S-N-curve defined in FprEN1992-1-1:2023 (E), Annex J, formulae (J.29) and (J.30) can be
used.
Figure 3.1: Principle sketches of reinforced concrete sections with subsequent strengthening by externally
bonded CFRP sheets
The product-specific design parameters for kits of EBR CFRP sheets included in Annex J of FprEN 1992-1-1:2023
(E) are the following:
Table 3.2: Sources of design and application parameters from test procedures according to Annex BD-B for
kits of externally bonded CFRP sheets
3)
𝑇 minimum temperature of the
strengthening kit at intended use
𝑇 maximum temperature of the B.2.3.32)
strengthening kit at intended use
𝐶 highest concrete strength class of the B.2.32)
tests
𝐶 lowest concrete strength class of the tests B.2.32)
1)
determined by test procedure
2)
declared by the manufacturer and confirmed by test procedure
3)
declared by the manufacturer
The number of tests given in Annex BD-B, clauses B.2.1 to B.2.3 is the minimum number which is necessary.
Each applicant can expand it according to their needs.
3.3.1 The creep reduction factor ηf of CFRP sheets derived from the tests according
to Annex BD-B, B.2.2.6 and B.2.2.7
The factor 𝜂 is the minimum of the factors 𝑅 from the test B.2.2.6 and 𝑅 from the test B.2.2.7
(𝜂 Min 𝑅 , 𝑅 . It is given separately for all alkaline solutions used in B.2.2.6 and B.2.2.7 (pH 9 and/or 11
and/or 13,7).
3.3.2 Parameters of the bilinear bond approach ksys,b1 and ksys,b2 and ksys,b3 derived from the tests
according to Annex BD-B, B.2.3.5
The differential formula for the bond of externally bonded reinforcement given in the Figure 2.3 can be completely
solved for the two branches of the bilinear bond approach. With an adjustment to the boundary conditions of the
bond test specimen according to Figure B.15, the following formulae are obtained, which describe the relationship
between the relative displacement sf at the load-side beginning of the bond length 𝑙 and the tensile force 𝐹 . For
s 𝑠 , only elastic bond stresses occur within the total bond length 𝑙 . After exceeding the elastic limit
displacement (𝑠 𝑠 ), additional plastically softened bond areas occur.
𝐹 𝐸 ∙ 𝑏 ∙ 𝑡 ∙ 𝑠 ∙ 𝜔 ∙ tanh 𝜔 ∙ 𝑙 for 𝑠 𝑠
∙ ∙ ∙
F 𝐸 ∙𝑏 ∙𝑡 ∙𝜔∙𝜆∙ for 𝑠 𝑠
∙ ∙
With:
𝜏
𝜔
𝑠 ∙𝐸 ∙𝑡
𝑠
𝜆
𝑠 𝑠
The length 𝑎 of the elastically deformed bond region enters into the second formula and can be determined with
the following expression. Because of the simultaneous occurrence of hyperbolic and trigonometric functions, 𝑎
cannot be determined in complete form but only iteratively.
𝜆 𝑠 𝑠
tanh 𝜔 ∙ 𝑎 ∙ s 𝑠 ∙ cot 𝜔 ∙ 𝜆 ∙ 𝑙 𝑎
𝑠 sin 𝜔 ∙ 𝜆 ∙ 𝑙 𝑎
The parameters 𝑠 , 𝜏 , and 𝑠 of the bilinear composite approach (cf. Figure 2.2) are determined using the least
squares method on the experimentally determined 𝑠 -𝜀 curves (cf. Figure 2.3) for each of the two sides of a
specimen. In view of the scope of the calculations to be performed, a limited number of supporting points is
recommended for the computational determination of the parameters.
Subsequently, system-specific coefficients 𝑘 , and 𝑘 , can be determined in regression analyses according
to EN 1990, Annex D.8.
∙
For this purpose, the correlations 𝜏 0 ⇔𝑓 𝑓 , 0 and 𝐺 0 ⇔𝑓 𝑓 , 0 can be
used in addition. For strengthening systems with externally bonded reinforcement, in which the adhesive
(component B2) does not lead to a hardening of the concrete layers near the surface, the following the following
simple functions 𝑓 and 𝑓 of the term 𝑓 ∙ 𝑓 , can be assumed for this purpose,
𝜏 𝑓 𝑓 ∙𝑓 , 𝑘 , ∙ 0,37 ∙ 𝑓 ∙𝑓 ,
𝑘 , ∙ 0,37 ∙ 𝑓 ∙𝑓 , ∙𝑘 , ∙ 0,2
𝐺 𝑓 𝑓 ∙𝑓 ,
2
𝑠 𝑘 , ∙ 0,2
For many known reinforcement systems with CFRP sheets and cold-curing epoxy resin adhesives, 𝑘 , and
𝑘 , are greater than 1,0.
The region before the maximum bond length is described in formula (J.38) via a parabola. However, the exact
solution from the bilinear composite approach is not exactly parabolic.
After the test interpretation, a better agreement of the approximation from the parabolic approach with the test
values can be achieved with the factor 𝑘 , .
If, on the other hand, a low viscosity adhesive or primer is used that penetrates the concrete matrix and leads to
a hardening of the concrete layers close to the surface, a more attenuated relationship between the target values
and the concrete strengths is to be assumed, e.g.
𝜏 𝑓 𝑓 ∙𝑓 ,
𝐺 𝑓 𝑓 ∙𝑓 ,
3.3.3 Reduction factor for fatigue derived from the tests according to Annex BD-B, B.2.3.6
The fatigue reduction factor 𝛼 used FprEN 1992-1-1:2023 (E), Annex J, formula (J.27) contains information for
an S-N-curve, which is verified by means of fatigue tests. The tests are performed according to the test procedure
in Annex BD-B, Clause B.2.3.6. The following load levels are checked:
The load levels are characterised by the average value of the anchorage failure load 𝐹 determined for the same
configuration of specimen under B.2.3.5. During the test the number of load cycles is counted until a separated
length of 30 mm is reached, see Clause B.2.3.6.
If at least 2 million load cycles are achieved at all load levels A to D in the first test and at least 90.000 load cycles
in the second test, the S-N-curve defined in FprEN 1992-1-1:2023 (E), Annex J, formulae (J.29) and (J.30) can
be used.
4 Product specific design and application parameters for strengthening kits of near
surfaced mounted CFRP strips (NSM)
Figure 4.1: Principle sketches of reinforced concrete sections with subsequent strengthening by near surface
mounted CFRP strips
The product-specific design parameters for strengthening kits of NSM CFRP included in Annex J of FprEN 1992-
1-1:2023 (E) are the following:
1)
determined by test procedure
2)
declared by the manufacturer and confirmed by test procedure
3)
declared by the manufacturer
The number of tests given in Annex BD-A, clauses A.2.1 and A.2.2 and in Annex BD-C, clause C.2.3 is the
minimum number which is necessary. Each applicant can expand it according to their needs.
4.3.1 The creep reduction factor ηf of CFRP strips derived from the tests according to
Annex BD-A, A.2.2.5 and A.2.2.6
The factor 𝜂 is the minimum of the factors 𝑅 from the test A.2.2.5 and 𝑅 from the test A.2.2.6
(𝜂 𝑀𝑖𝑛 𝑅 , 𝑅 . It is given separately for all alkaline solutions used in 2.2.5 and 2.2.6 (pH 9 and/or 11 and/or
13,7).
References 5 References
[1] EN 1504-4:2004: Products and systems for the protection and repair of concrete structures. Definitions,
requirements, quality control and evaluation of conformity. Part 4: Structural bonding.
[2] Kuntz, S.: Verbundverhalten von laschenverstärkten Stahlbetonzuggliedern: Diplomarbeit, Institut für
Baustoffe, Massivbau und Brandschutz, TU Braunschweig; 1993.
[3] Holzenkämpfer, P.: Ingenieurmodelle des Verbundes geklebter Bewehrung für Betonbauteile;
Dissertation, TU Braunschweig; 1994.
[4] Niedermeier, R.: Zugkraftdeckung von klebearmierten Bauteilen; Dissertation, TU München; 2001.
A.1.8 Minimum and maximum thickness of layers of bonding agent (component B) and repair
mortar (component C and D)
The default minimum and maximum thicknesses
of the layers of bonding agent (component B) are 𝑑 , = 1 mm and 𝑑 , = 5 mm and
of the repair mortar (component C and D) 𝑑 , = 5 mm and 𝑑 , = 30 mm.
The default values are only used if the manufacturer's instructions and safety information do not contain
any other information.
A.1.9 Minimum curing time of the bonding agent (component B) and repair mortar (component C
and D) at standard environmental conditions (A.1.4) and at minimum temperature of
installation (see A.1.5)
The default values for the minimum curing time of the bonding agent (component B) at standard
environmental conditions (A.1.4) 𝑡 are 7 d and at minimum temperature of installation (A.1.5) 𝑡
28 d. The default values are only used if the manufacturer's instructions and safety information do not
contain any other information.
The default values for the minimum curing time of the repair mortar (component C and D) at standard
environmental conditions (A.1.4) 𝑡 are 7d and at minimum temperature of installation (A.1.5) 𝑡
28 d. The default values are only used if the manufacturer's instructions and safety information do not
contain any other information.
A.2.1 Tests to determine the product-specific properties of the adhesive (structural bonding agent)
and repair mortar
A.2.1.1 Glass transition temperature of the structural bonding agent (component B) and the repair
mortar (component C)
Purpose of the assessment
To identify the maximum temperature of intended use (see A.1.6) of the kit is necessary to determine the glass
transition temperature of the structural bonding agent (component B) 𝑇 and the repair mortar (component C)
𝑇 .
Assessment method
Basis for the tests is EN 12614. The glass transition temperature shall be determined on specimens of cured
structural bonding agent (component B) and repair mortar (component C) from one batch of resin and hardener.
The glass transition temperature must only be measured during the first heating cycle; the heating speed must
be 10°C/min. The preparatory cycle of temperature according to EN 12614, clause 6.2 shall be omitted if it causes
a change of the glass transition temperature.
The glass transition temperature shall be determined on at least two specimens of bonding agent (component B)
which have cured at least for 𝑡 (see A.1.9) and on at least two specimens of repair mortar (component C) which
have cured at least for 𝑡 (see A.1.9) under standard environmental conditions (see A.1.4).
Expression of results
Glass transition temperatures of the structural bonding agent (component B) 𝑇 together with the curing time 𝑡
(see A.1.9) and the repair mortar (component C) 𝑇 together with the curing time 𝑡 (see A.1.9) shall be given
in the test report.
A.2.1.2 Flexural and compressive strength of the structural bonding agent (component B) and the
repair mortar (component C)
Purpose of the assessment
These tests are intended to determine the relevant mechanical properties of the structural bonding agent
(component B) and the repair mortar (component C) after curing under standard environmental conditions (see
A.1.4). This includes the flexural and compressive strength.
Assessment method
The flexural and compressive strength according to EN 196-1 or EN 1015-11 of at least 6 specimens 160 mm x
40 mm x 40 mm (length x height x width) of cured structural bonding agent and repair mortar from one batch of
resin and hardener shall be determined. The specimens shall cure at standard environmental conditions (see
A.1.4). At least 3 tests of the flexural strength according to EN 196-1, 9.1 or EN 1015-11, clauses 8 and 6 tests of
the compressive strength according to EN 196-1, clause 9.2 or EN 1015-11, clause 9 shall be carried out at
standard environmental conditions (see A.1.4).
At least tests after curing of the bonding agent (component B) for 𝑡 (see A.1.9) and of the repair mortar
(component C) for 𝑡 (see A.1.9) shall be carried out.
Expression of results
From the failure loads of the bending tests with the specimens of structural bonding agent 𝐹Gfl, and the specimens
of repair mortar 𝐹Mfl, , the bending tensile strengths 𝑓Gfli and 𝑓Mfli shall be determined according to EN 196-1, clause
9.1 or EN 1015-11, clause 8. From these values, the mean values 𝑓Gflm and 𝑓 as well as the 5% quantile value
according to EN 1990, Table D.1 (Vx unknown) 𝑓Gflk and 𝑓 shall be determined and given in the test report.
From the failure loads of the compressive tests with the specimens of structural bonding agent 𝐹Gc, and the
specimens of repair mortar, 𝐹Mc, the compressive strengths 𝑓 and 𝑓Mci shall be determined according to EN 196-
1, clause 9.2 or EN 1015-11, clause 9. From these values, the mean value 𝑓Gcm and 𝑓Mcm as well as the 5% quantile
value according to EN 1990, Table D.1 (Vx unknown) 𝑓Gck and 𝑓Mck shall be determined and given in the test report.
A.2.1.3 Flexural and compressive strength of the structural bonding agent (component B) and the
repair mortar (component C) cured at minimum temperature of installation in dependence of
curing time
Purpose of the assessment
These tests are intended to determine the flexural and compressive strength of the structural bonding agent
(component B) and the repair mortar (component C) at minimum temperature of installation (see A.1.5) in
dependence of curing time.
Assessment method
For different curing conditions regarding to temperature and time the flexural and compression strength of cured
specimens of bonding agent and repair mortar shall be determined according to A.2.1.2. At least specimens of
structural bonding agent (component B) cured at minimum temperature of installation 𝑇 , °C (see A.1.5) and
specimens of repair mortar (component C) cured at minimum temperature of installation 𝑇 , °C (see A.1.5)
shall be tested.
For each material at least 3 tests shall be carried out. The bending and compression test shall be carried out
immediately after the curing time ending. For the bonding agent (component B) at least tests after curing for tGTmin
and 2 tGTmin (see A.1.9) and for the repair mortar (component C and D) at least tests after curing for tMTmin and 2
tMTmin (see A.1.9) at the minimum temperature of installation 𝑇 , and 𝑇 , (see A.1.6) shall be carried out.
Expression of results
From the failure loads of the bending tests "i" for the curing time "j" of structural bonding agent 𝐹Gfl,ij and the
specimens of repair mortar 𝐹Mfl,ij the bending tensile strengths 𝑓Gflij and 𝑓Mtflj shall be determined according to
EN 196-1, clause 9.1 or EN 1015-11, clause 8. From these values, the mean value 𝑓Gtmj and 𝑓Mtmj as well as the
5% quantile value according to EN 1990, Table D.1 (Vx unknown) 𝑓Gflkj and 𝑓Mflkj shall be given in the test report
in conjunction with the used curing temperature and time.
From the failure loads of the compression tests "i" for the curing time "j" of the structural bonding agent 𝐹 , and
the specimens of repair mortar 𝐹 , the bending compression strengths 𝑓Gcij and 𝑓Mcij shall be determined
according to EN 196-1, clause 9.1 or EN 1015-11, clause 9. From these values, the mean value 𝑓Gcmj and 𝑓Mcmj
as well as the 5% quantile value according to EN 1990, Table D.1 (Vx unknown) 𝑓Gckj and 𝑓Mckj shall be given in
the test report in conjunction with the used curing temperature and time.
A.2.2.4 Modulus of elasticity, tensile strength and strain at failure of CFRP strips (component A)
Purpose of the test
These tests are intended to determine the relevant mechanical properties of the CFRP strips (component A). This
includes the modulus of elasticity and the tensile strength, both related to the section of the CFRP strip as well as
the ultimate strain of CFRP strips.
Assessment method
The tensile load at failure 𝐹 (called 𝑃 in EN 2561, 7.2.7), the strains at 0,1 𝐹 as well as at 0,5 𝐹 and the strain
at failure 𝜀 (called 𝜀 in EN 2561, clause 7.2.7) of the CFRP strips shall be determined according to EN 2561,
Annex A, with specimens type B. In deviation from EN 2561, Annex A, Table A.1 the thickness of the test
specimens may correspond to the thickness of the CFRP strips. Tests shall be carried out at least on 5 specimens
of the significant cross sections (see A.1.1) of every type of CFRP strip (see A.1.3).
Expression of results
According to EN 2561 it is only possible to measure failure loads 𝐹 , the strains 𝜀 )A at 0,1 𝐹 (called 𝜀 )A in
EN 2561, clauses 8.3 and 8.4), the strains 𝜀 )B at 0,5 𝐹 (called 𝜀 )B in EN 2561, clauses 8.3 and 8.4) and
the strains at failure 𝜀 (called (11)R in EN 2561, clause 8.6). To determine strengths and secant moduli of
elasticity from these test results there are two possibilities. The strengths related to the sections of the CFRP
strips (called in EN 2561, clause 8.1) and the strengths related to the sections of all fibres of the sections of
CFRP strips (called in EN 2561, clause 8.2) as well as the moduli of elasticity related to the sections of the
CFRP strips (called 𝐸 in EN 2561, clause 8.3) and the moduli of elasticity related to the sections of all fibres of
the sections of CFRP strips (called 𝐸 in EN 2561, clause 8.4). With CFRP strips, it is customary to refer to the
cross sections of the strips. Therefore, the tensile strength and the moduli of elasticity shall only be determined
related to the section of the CFRP strips according to EN 2561, clause 8.1 and EN 2561, clause 8.3.
From the single values 𝐸 , 𝑓 , 𝜀 according to EN 2561, 8.3, 8.1 and 7.2.7 he average values of the secant
modulus of elasticity 𝐸 , of the tensile strengths 𝑓 and of the ultimate strain 𝜀 and the 5% quantile value
according to EN 1990, Table D.1 (𝑉 unknown) of tensile strengths 𝑓 shall be given in the test report for all types
of CFRP strips (see A.1.3).
A.2.2.5 Resistance of CFRP strips (component A) after storage in alkaline environment at maximum
temperature at intended use (see A.1.6)
Purpose of the assessment
These tests are intended to determine the change of the mechanical properties of the CFRP strips (component
A) according to A.2.2.4 after long-time storage in alkaline environment at maximum temperature at intended use
(see A.1.6). This includes the change of the secant modulus of elasticity, tensile strength and ultimate strain of
the CFRP strips.
Assessment method
At least in one of the following alkaline solutions the CFRP strips shall be stored at maximum temperature at
intended use (see A.1.6) for at least 1800 h:
alkaline solution pH 9,0,
alkaline solution pH 11,0,
alkaline solution pH 13,7.
After this storage the specimens are washed and dried for 24 h at standard environmental conditions (see A.1.4).
Afterwards the tensile load at failure 𝐹 (called 𝑃 in EN 2561, clause 7.2.7), the strains at 0,1 𝐹 as well as at
0,5 𝐹 and the strain at failure 𝜀 (called 𝜀 in EN 2561, clause 7.2.7) of the CFRP strips shall be determined
according to EN 2561, Annex A, with specimens type B. In deviation from EN 2561, Annex A, Table A.1 the
thickness of the test specimens may correspond to the thickness of the CFRP strips.
Tests shall be carried out at least on 5 specimens of CFRP strips for at least one thickness per type of CFRP
strips (see A.1.3). Preferably it should be the section with the highest ratio between circumference and cross-
section area among all the sections of this type of CFRP strips (see A.1.3).
Expression of results
According to EN 2561 it is only possible to measure failure loads 𝐹 , the strains 𝜀 )A at 0,1 𝐹 (called 𝜀 )A in
EN 2561, clauses 8.3 and 8.4), the strains 𝜀 )B at 0,5 𝐹 (called 𝜀 )B in EN 2561, clauses 8.3 and 8.4) and
the strains at failure 𝜀 (called 𝜀 in EN 2561, clause 8.6). To determine strengths and secant moduli of
elasticity from these test results there are two possibilities according to EN 2561. In accordance to A.2.2.4 it is
only necessary to determine the strengths related to the sections of the CFRP strips 𝑓 (called in EN 2561,
clause 8.1) as well as the moduli of elasticity related to the sections of the CFRP strips 𝐸 (called 𝐸 in EN
2561, clause 8.3).
From the single values 𝐸 , 𝑓 and 𝜀 according to EN 2561, 8.3, 8.1 and 7.2.7 the average values 𝐸 ,𝑓
and 𝜀 as well as the ratios 𝑅 , 𝑅 and 𝑅 shall be determined as follows:
𝑅 𝑅 𝑅
where:
𝐸 ,𝑓 ,𝜀 are the average Modulus of elasticity, tensile strength and ultimate strain after
storage in alkaline solution
𝐸 , 𝑓 , 𝜀, are the average Modulus of elasticity, tensile strength and ultimate strain
according to A.2.2.4
The values 𝑅 , 𝑅 and 𝑅 shall be given in the test report for all types (see A.1.3) of CFRP strips separately
for all applied alkaline solutions (pH 9 and/or 11 and/or 13.7) used for the tests.
A.2.2.6 Resistance of CFRP strips (component A) in alkaline environment, under long-term load at
maximum temperature at maximum temperature at intended use (see A.1.6)
Purpose of the test
These tests are intended to determine the change of the mechanical properties of the CFRP strips (component
A) according to A.2.2.4 after long-time loading in alkaline environment at maximum temperature at intended use
(see A.1.6). This includes the change of the secant modulus of elasticity, of the tensile strength and of the ultimate
strain of the CFRP strips. In comparison to A.2.2.4 the secant modulus of elasticity and the tensile strength only
needs to be related to the cross section of the CFRP strips and not to the cross section of all fibres of the CFRP
strips.
Assessment method
At least in one of the following alkaline solutions the CFRP strips shall be stored at maximum temperature at
intended use (see A.1.6) and stressed with 50% of the characteristic value of tensile strength according to A.2.2.4
for at least 1800 h:
alkaline solution pH 9,0,
alkaline solution pH 11,0,
alkaline solution pH 13,7.
After this storage the specimens are washed and dried for 24 h at standard environmental conditions (see A.1.4).
Afterwards the tensile load at failure 𝐹 (called 𝑃 in EN 2561, 7.2.7), the strains at 0,1 𝐹 as well as at 0,5 𝐹
and the strain at failure 𝜀 (called 𝜀 in EN 2561, 7.2.7) of the CFRP strips shall be determined according to
EN 2561, Annex A, with specimens type B. In deviation from EN 2561, Annex A, Table A.1 the thickness of the
test specimens may correspond to the thickness of the CFRP strips.
Tests shall be carried out at least on 5 specimens of CFRP strips for at least one thickness per type of CFRP
strips (see A.1.3). Preferably it should be the section with the highest ratio between circumference and cross-
section area among all the sections of this type of CFRP strips (see A.1.3).
Expression of results
According to EN 2561 it is only possible to measure failure loads 𝐹 , the strains 𝜀 at 0,1 𝐹 (called 𝜀
in EN 2561, clauses 8.3 and 8.4), the strains 𝜀 at 0,5 𝐹 (called 𝜀 in EN 2561, clauses 8.3 and 8.4) and
the strains at failure 𝜀 (called 𝜀 in EN 2561, clause 8.6). To determine strengths and secant moduli of
elasticity from these test results there are two possibilities according to EN 2561. In accordance to 2.2.4 it is only
necessary to determine the strengths related to the cross section of the CFRP strips 𝑓 (called in EN 2561,
clause 8.1) as well as the moduli of elasticity related to the sections of the CFRP strips 𝐸 (called 𝐸 in EN
2561, clause 8.3) and not to the cross section of all fibres of the CFRP strips.
From the single values 𝐸 , 𝑓 and 𝜀 according to EN 2561, 8.3, 8.1 and 7.2.7 the average values 𝐸 ,𝑓
and 𝜀 as well as the ratios 𝑅 , 𝑅 and 𝑅 shall be determined as follows:
𝑅 𝑅 𝑅
where:
𝐸 ,𝑓 ,𝜀 are the average Modulus of elasticity, tensile strength and ultimate strain
after storage in alkaline solution
𝐸 ,𝑓 ,𝜀 are the average Modulus of elasticity, tensile strength and ultimate strain
according to A.2.2.4
The values 𝑅 , 𝑅 and 𝑅 shall be reported in the test report for all types (see A.1.3) of CFRP strips
separately for all applied alkaline solutions (pH 9 and/or 11 and/or 13.7) used for the tests.
A.2.3 Tests to verify the product-specific properties TGmin,a, TMmin,a, Tmax and Cmin, Cmax as well as to
determine the product-specific parameters of the bilinear bond approach ksys,b1 and ksys,b2 and
the reduction factor for fatigue loading fat2
A.2.3.1 Bond strength of specimens cured at standard environmental conditions (see A.1.4)
Purpose of the assessment
The tests are performed to determine the bond strength of the following specimens:
specimens with layers of concrete, repair mortar (component C and D), bonding agent (component B)
and CFRP strips (component A),
specimens with layers of concrete, bonding agent (component B) and CFRP strips (component A).
The bond strength is determined by pull-off tests according to EN 1542 after curing under standard conditions
(see A.1.4).
Assessment method
The test is based on EN 1542. The main differences to EN 1542 are the following:
only steel dollies according to EN 1542, 4.7 are used, not dollies of aluminium,
the bonding agent (Component B) is used also for the bonding of the steel dollies,
the thickness of the layers of bonding agent (also for the dollies) is 𝑑 , (see A.1.8),
The preparation of the specimens with repair mortar shall be done as follows:
On at least one specimen according to EN 1542, clause 4.12, with the maximum strength class of the
concrete substrate (see A.1.10) a layer of repair mortar (component C and D) with the maximum
thickness 𝑑 , 1 (see A.1.8) shall be installed according to the manufacturer’s instructions and safety
information (see Figure A.3a).
The preparation of the surface according to EN 1542, clause 5.1, shall be done for the use in vertical
position.
On the repair mortar (component C and D) a layer of CFRP strips bonded by a layer of bonding agent
(component B) with the minimum thickness 𝑑 , 2 (see A.1.8) shall be installed according to the
manufacturer’s instructions and safety information (see Figure A.3b). Preferably the width of the CFRP
strips should be at least 50 mm, to allow the drilling of the annular grooves without joints between CFRP
strips.
After curing of the repair mortar (component C and D) and the bonding agent (component B) according
to the manufacturer's instructions and safety information at standard environmental conditions (see A.1.4)
on each specimen 5 annular grooves of 50 mm diameter shall be drilled through the multilayer specimen
until a depth of 15 ± 5 mm in the reference concrete is reached (see Figure A.3). The layers of CFRP
strips inside the annular grooves preferably should not contain joints between CFRP strips.
1 attached formwork
2 reference body of concrete according to EN 1542
3 layer of repair mortar (component C and D) with minimum thickness 𝑑 ,
4 layer of bonding agent (component B) with minimum thickness 𝑑 ,
5 CFRP strip (component A) externally bonded to the repair mortar (component C and D)
Figure A.3: Cross section through the test specimen with repair mortar (component C and D)
a) Specimen with repair mortar (component C and D)
b) Specimen with repair mortar (component C and D), bonding agent (component B) and externally
bonded CFRP strips (component A)
1
At least one test series with the maximum layer thickness of repair mortar is necessary.
2
At least one test series with the minimum layer thickness of bonding agent is necessary.
After curing of the bonding agent (component B) according to the manufacturer's instructions and safety
information at standard environmental conditions (see A.1.4) on each specimen 5 annular grooves of
50 mm diameter shall be drilled through the multilayer specimen until a depth of 15 ± 5 mm in the
reference concrete is reached. The layers of CFRP strips inside the annular grooves preferably should
not contain joints between CFRP strips.
1 attached formwork
2 reference body of concrete according to EN 1542
4 layer of bonding agent (component B) with minimum thickness 𝑑 , (see A.1.9)
5 CFRP strip (component A) externally bonded to the concrete
Figure A.5: Cross section through the test specimen without repair mortar
a) Specimen without repair mortar
b) Specimen with bonding agent (component B) and externally bonded CFRP strips (component A)
The further treatment is the same for the test specimens with and without repair mortar:
On each annular groove a dolly according to EN 1542, clause 7.2, shall be glued on the circular surfaces
with the structural bonding agent (component B). The minimum thickness of the layer 𝑑 , (see A.1.8)
of bonding agent shall be used.
The structural bonding agent (component B) of the dollies shall cure according to the manufacturer's
instructions and safety information at standard environmental conditions (see A.1.4) before testing.
Afterwards the bond strength 𝑓cti,surf shall be determined on each annular groove of the specimens with
and without repair mortar according to EN 1542, 7.4 at standard environmental conditions (see A.1.4).
The bond strength shall be determined on at least 5 annular grooves with and without repair mortar (component
C and D) for every type of CFRP strip.