0% found this document useful (0 votes)
19 views6 pages

Dyck Et Al SM 2024a

This study investigates the effects of boundary conditions on phase stabilities in the palladium-hydrogen system, focusing on how different mechanical constraints influence phase transformations. The research reveals that increased dimensionality of constraints leads to significant compressive stresses that destabilize the hydride phase, resulting in a reduced critical temperature for hydride formation. The findings suggest that under 2D constraints, the critical temperature for hydride formation is predicted to be 307 K, indicating a driving force for hydride formation at room temperature.

Uploaded by

Tarkes Dora
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
19 views6 pages

Dyck Et Al SM 2024a

This study investigates the effects of boundary conditions on phase stabilities in the palladium-hydrogen system, focusing on how different mechanical constraints influence phase transformations. The research reveals that increased dimensionality of constraints leads to significant compressive stresses that destabilize the hydride phase, resulting in a reduced critical temperature for hydride formation. The findings suggest that under 2D constraints, the critical temperature for hydride formation is predicted to be 307 K, indicating a driving force for hydride formation at room temperature.

Uploaded by

Tarkes Dora
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 6

Scripta Materialia 247 (2024) 116117

Contents lists available at ScienceDirect

Scripta Materialia
journal homepage: www.journals.elsevier.com/scripta-materialia

Phase transformation in the palladium hydrogen system: Effects of


boundary conditions on phase stabilities
Alexander Dyck a , Thomas Böhlke a , Astrid Pundt b , Stefan Wagner b,∗
a Institute of Engineering Mechanics - Chair for Continuum Mechanics, Karlsruhe Institute of Technology (KIT), Kaiserstraße 10, 76131 Karlsruhe, Germany
b Institute for Applied Materials – Materials Science and Engineering, Karlsruhe Institute of Technology (KIT), Engelbert-Arnold–Straße 4, 76131 Karlsruhe, Germany

A R T I C L E I N F O A B S T R A C T

Keywords: Different elastic constraint conditions affect the phase stabilities of metal-hydrogen systems. This is studied
Metal-hydrogen system considering the free energy density within a chemo-mechanically coupled approach with linear elastic
Thermodynamics deformations and homogeneous concentrations. Utilizing the palladium-hydrogen alloy as a model system, the
Mechanical stress
effects of various mechanical constraints in 1D, 2D and 3D are investigated. These constraints change occurring
Phase stability
Critical temperature
mechanical deformations and strongly influence the systems chemical potential compared to the unconstrained
system. With increasing dimensionality of the constraints, large compressive mechanical stresses occur, which
destabilize the hydride phase. This yields a reduced critical temperature of hydride formation. Spinodal and
equilibrium miscibility gaps of the system are deduced as a function of the boundary conditions. Notably,
the critical temperature of the ideal palladium-hydrogen system with 2D constraints is predicted to be 307 K ,
revealing a driving force for hydride formation at room temperature even under these constraints.

As recently reported for metal-hydrogen or Li-ion battery systems, lattice expansion is suppressed, depending on the actual boundary con-
in constrained alloy systems such as thin films adhered to rigid sub- ditions. Large compressive stresses of the order of several GPa result,
strates or clusters embedded in a rigid matrix, the thermodynamics de-stabilizing the hydride phase [9–13,24–26]. In the present paper, we
of structural phase transitions is altered by the mechanical stress state study the stability of the solid solution phase and the conditions for the
[1–6]. Structural phase transitions can both be related to different crys- onset of phase transformation applying linear elastic theory in terms
tal structures and different unit cell volumes of the phases. This results of a chemo-mechanically coupled approach. The palladium-hydrogen
in misfit conditions at coherent interfaces, imposing high mechanical system is utilized as model system. We aim at revealing the physical
stresses [2,3,7–11]. In total, a three-dimensional stress state results, boundary conditions leading to a mechanical stress-driven suppression
that increases the system’s elastic energy density and hence changes of the phase transformation for a given temperature. Subsequently,
the global chemical potential [1,12–14]. With respect to the free sys- these effects are discussed in detail by considering the free energy and
tem, this leads to modified thermodynamic conditions of two-phase the chemical potential determining the phase stability.
equilibria in constrained systems, unveiled in modified thermodynamic Phase equilibria in hydride forming metals can be determined as
stabilities of phases, modified critical temperatures of phase transfor- minima of the free energy density 𝜓 of the system consisting of the
mations, as well as in shifted terminal compositions of the phases metal and hydrogen, according to Gibb’s thermodynamics [27]. Coex-
[1,9–14]. In order to study contributions of mechanical stresses, their isting potential troughs of the free energy density describe the thermo-
effect on two-phase equilibria and on the coherency conditions at in- dynamic equilibrium of the phases via the double-tangent construction
terfaces, metal-hydrogen systems are archetypical model systems due [1]. In addition, in a chemo-mechanically coupled model approach, the
to the ease of their experimental handling and the well-known bulk free energy density serves as a potential for both stresses and the chem-
phase diagrams [1–4,9–11,15–21]. In metal-hydrogen systems, hydro- ical potential [28]. Thus, in order to study metal-hydrogen systems and
gen absorption and the formation of hydride phases result in a volume hydride formation, the free energy density has to be specified. In this
expansion of the host metal lattice scaling mainly linearly with the work, the independent variables determining the free energy density
global hydrogen concentration [8,22,23]. In constrained systems this are chosen to be the hydrogen concentration 𝑐 in moles per volume,

* Corresponding author.
E-mail address: [email protected] (S. Wagner).

https://doi.org/10.1016/j.scriptamat.2024.116117
Received 7 February 2024; Received in revised form 13 March 2024; Accepted 3 April 2024
Available online 12 April 2024
1359-6462/© 2024 The Authors. Published by Elsevier Ltd on behalf of Acta Materialia Inc. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
A. Dyck, T. Böhlke, A. Pundt et al. Scripta Materialia 247 (2024) 116117

temperature 𝜃 and occurring strains 𝜺 [28]. Commonly, the total vol- concentration and studying the sign of the resulting function. The dif-
ume specific free energy density 𝜓 of an elastically deforming metal ferentiation results in
absorbing hydrogen is additively split in two parts [28] 𝜕2 𝜓 𝜕𝜇H 𝑅𝜃𝑟
𝑐max = = − 𝜈0 𝜂H 𝑚 − 𝐸HH , (7)
( ) 𝜕𝑐H2 𝜕𝑐H 𝑟𝑐H − 𝑐H2
𝜓 𝜺, 𝑐H = 𝜓c (𝑐H , 𝜃) + 𝜓e (𝜺, 𝑐H ), (1)
where
where the first part is chemical and depends on the temperature 𝜃
and the total dimensionless hydrogen concentration 𝑐H in H/Pd. This 𝜕 tr (𝝈)
𝑚= . (8)
dimensionless concentration is defined as the ratio between the total H- 𝜕𝑐H
concentration c and the maximum concentration 𝑐max , thus 𝑐H = 𝑐∕𝑐max . Thus, for given 𝜃 and 𝐸HH , the resulting stresses either suppress or al-
The concentration 𝑐max is the maximum hydrogen concentration achiev- low for a phase separation, i.e., hydride formation, to occur. We note
able from a structural point of view, with 𝑐max = 1∕𝜈0 and the volume that different from this approach Alefeld [30] considered an elastic
density 𝜈0 of interstitial sites in moles available for hydrogen in the reduction of the H-H interaction strength depending on the physical
metal lattice. This ansatz is based on the assumption, that elastic de- boundary conditions of the system, while we distinguish the strain-
formations do not alter the chemical properties of the metal-hydrogen induced H-H interaction and the mechanical stress impact on hydride
system and vice versa. The second part represents the elastic contribu- sp
formation. As long as there exists a spinodal composition 𝑐H ∈ [0, 𝑟], for
tion due to deformations 𝜺 and chemical strains of the metal. The elastic which
part is a quadratic function of the elastic strains within the sample and
𝜕𝜇H ||
given by [28] =0 (9)
𝜕𝑐H ||𝑐 sp
1 [ ] H
𝜓e (𝜺, 𝑐H ) = 𝜺e ⋅ ℂ 𝜺e , (2)
2 holds, phase separation is possible. At the spinodal composition, any
where the elastic strain is the difference of the total strain and the oc- concentration fluctuation in the system builds up. Related equilibrium
curring chemical strains, 𝜺e =[𝜺 −] 𝜺0 and ℂ is the stiffness tensor of the concentrations of the phases with the maximum solubility 𝑐𝛼max of the
solid solution phase and the minimum concentration 𝑐𝛽min of the hy-
[ ] The expression 𝜺[e ⋅]ℂ 𝜺e denotes the scalar product between
metal.
ℂ 𝜺e and 𝜺e , while ℂ 𝜺e denotes the linear mapping of the elastic dride phase can be deduced via Maxwell constructions of the chemical
strains via the stiffness tensor. The chemical or compositional strain potential. Additionally, by rearranging Eq. (7), a critical temperature
𝜺0 results from an isotropic hydrogen-induced lattice expansion of the 𝜃crit can be determined, above which no phase separation occurs for
metal, when it occupies interstitial lattice sites being linear in the con- given 𝐸HH . All specified concentrations are functions of the boundary
centration [25] conditions. In the following, stress states for varying constraint condi-
tions in the case of homogeneous concentrations are derived and phase
𝜺0 = 𝜂H 𝑐H 𝑰, (3) separation and its suppression in the palladium-hydrogen system with
different constraint conditions are studied.
with the lattice expansion coefficient 𝜂H and identity tensor 𝑰 . The
Palladium is capable of interstitially absorbing hydrogen in large
chemical part of the free energy density is given by [1,28]
amounts. The volume density of available octahedral interstitial sites
( ( ) ( )) for hydrogen absorption is 𝜈0 = 8.94 × 10−6 m3 ∕mol and 𝑟 = 0.62 H∕Pd
𝑟 𝑐H
𝜓c = 𝜇0 𝑐H 𝑐max + 𝑅𝜃𝑐max −𝑟ln + 𝑐H ln [1]. Since the parameter 𝑟 is mainly determined by electronic effects
𝑟 − 𝑐H 𝑟 − 𝑐H
𝑐max [29], it is assumed here not to change with H-induced stress. Beyond the
− 𝐸 𝑐2 . (4) solid solution limit and below the critical temperature 𝜃crit , phase trans-
2 HH H
formation with the emergence of a hydride phase can occur. At room
Here 𝜇0 is a reference chemical potential, 𝑅 is the gas constant and temperature, in the Pd-H bulk system the solid solution limit of the
𝐸HH is a long-range attractive hydrogen-hydrogen interaction energy. face-centered cubic 𝛼 phase is 𝑐𝛼max = 0.01 H∕Pd, which is in equilibrium
It results from the interaction of dissolved hydrogen atoms via the H- with the face-centered cubic 𝛽 ≈ 𝛼 ′ phase of 𝑐𝛼max = 0.6 H∕Pd composi-

induced dilatation field of the metal lattice. Hence, 𝐸HH defines the tion at intermediate global H concentrations [22]. Hydrogen absorption
driving force for the initiation of the phase transition in the metal hy- in Pd results in an isotropic volume expansion of Δ𝑉 ∕𝑉0 = 0.19𝑐H in
drogen system [1]. The logarithmic part results from an ansatz for the the whole range of hydrogen concentrations [22]. Hence, the linear
mixture entropy of hydrogen with metal, with the maximum number 𝑟 H-induced expansion coefficient of palladium is 𝜂H ≈ 0.063. For the
of hydrogen atoms per metal atom in the hydride phase. The quantity 𝑟 palladium-hydrogen bulk system 𝐸HH = 36.8(5) kJ∕mol [1].
is mainly determined by electronic interactions of H and the metal [29]. To specify the chemical potential as a function of concentration, the
We here assume that it does not depend on the stress state of the metal. computation of the mechanical stresses is necessary, cf. Eq. (6). To easen
The ansatz for the entropy is equivalent to the classical mixture entropy the considerations and to enable comparison of the calculations with
for 𝑟 = 1. This implies, that Eq. (4) considers both the electronic inter- results for, e.g., Pd thin films that usually grow with [111] texture, we
action of hydrogen atoms with the metal and the maximum number of here consider a palladium single crystal with its [111] direction parallel
interstitial sites available. As in the considered palladium – hydrogen to the 𝒆z -axis of the sample coordinate system. With the elastic con-
system the electronic interaction dominates at large concentrations, the stants 𝐶11 = 224 GPa, 𝐶12 = 173 GPa and 𝐶44 = 71.6 GPa of palladium,
parameter 𝑟 determines the maximum hydrogen concentration in Pd. the stiffness in Voigt-Mandel notation is (for details see the supplemen-
From the free energy density, both the chemical potential 𝜇H and the tal material)
mechanical stresses 𝝈 can be derived according to [28]
( ) ⎛ 270 158 142 −31 0 0 ⎞
𝜕𝜓 1 𝑐H ⎜ 158 270 142 31 0 0 ⎟
𝜇H = = 𝜇0 + 𝑅𝜃 ln − 𝐸HH 𝑐H − 𝜂H 𝜈0 tr (𝝈) , (5) ⎜ ⎟
𝜕𝑐H 𝑐max 𝑟 − 𝑐H 142 142 285 0 0 0
ℂ = 𝑸 ⋆ ℂ0 =̂ ⎜ ⎟ GPa, (10)
𝜕𝜓 [ ] ⎜ −31 31 0 82 0 0 ⎟
𝝈= = ℂ 𝜺 − 𝜺c . (6) ⎜ 0 0 0 0 82 −43 ⎟
𝜕𝜺 ⎜ ⎟
Hydride formation is possible, when the chemical potential is a non- ⎝ 0 0 0 0 −43 112 ⎠
monotonous function of 𝑐H [27]. We emphasize that this is true when no where 𝑸 is the rotation matrix rotating the [111] to the [001] plane, cf.
non-linear plastic deformation occurs [1]. The monotony of the chem- the supplemental material for this publication.
ical potential can be investigated by differentiation with respect to the Now the stress state can be specified as a linear function of the hydro-

2
A. Dyck, T. Böhlke, A. Pundt et al. Scripta Materialia 247 (2024) 116117

Fig. 1. Metal-hydrogen systems with different constraint conditions. Expansion of the sample cube is a) possible in all directions (0D constraint), b) suppressed in
𝒆𝑥 -direction, c) suppressed in 𝒆𝑥 - and 𝒆𝑦 -direction, and d) entirely suppressed by placing the cube in a rigid shell. The gray frames visually indicate the directions of
the constraint condition.

gen concentration depending on the considered constraint conditions, ⎛ 0 ⎞


⎜ 0 ⎟
which are summarized in Fig. 1 for a free system (0D constraint) as well ⎜ 2D ⎟
as systems with 1D, 2D and 3D constraints. We emphasize that for the 2D ⎜ 𝜀33 ⎟
𝜺 =̂ , (14)
0D, 1D and 2D constrained systems stresses evolving at coherent phase ⎜ 0 ⎟
⎜ 0 ⎟
interfaces are neglected. Hence, subsequently we focus on the driving ⎜ ⎟
force for the onset of phase transition in the homogeneous solid solution ⎝ 0 ⎠
phase with the specified rigid boundary conditions. Stresses evolving at since the system can expand freely in 𝒆𝑧 -direction. The strain 𝜀2D fol-
33
phase interfaces during phase transformation are more complex and
lows from the stress free condition 𝜎33
2D = 0, and the resulting stress state
will be targeted in a separate publication. is [1]

0D constraint For the unconstrained system of Fig. 1 a), the total strain ⎛ −18 ⎞
⎜ −18 ⎟
] ⎜ 0 ⎟
and stress follow as
[ 2D
𝝈 = ℂ 𝜺 − 𝜺0 =̂ ⎜
2D ⎟ 𝑐 GPa, 𝑚2D = −36 GPa. (15)
⎛0⎞ ⎜ 0 ⎟ H
⎜0⎟ ⎜ 0 ⎟
⎜ ⎟ ⎜ ⎟
[ ] 0 ⎝ 0 ⎠
𝜺0D = 𝜺0 , 𝝈 0D = ℂ 𝜺 − 𝜺0 =̂ ⎜ ⎟ (11)
⎜0⎟
⎜0⎟ 3D constraint Finally, for the 3D constrained system of Fig. 1 d) resem-
⎜ ⎟
⎝0⎠ bling e.g. a cluster in a surfactant shell or rigid matrix, the total strain
is
and hence the system is stress-free.
⎛0⎞
⎜0⎟
1D constraint For the system with 1D constraint in 𝒆𝑥 -direction in ⎜ ⎟
3D ⎜ 0 ⎟
Fig. 1 b), that resembles e.g. a narrow wire that is clamped in one di- 𝜺 =̂ . (16)
rection, the total strain is ⎜0⎟
⎜0⎟
⎜ ⎟
⎛ 0 ⎞ ⎝0⎠
⎜ 𝜀1D ⎟ The resulting stress state, with components increasing in comparison to
⎜ 22 ⎟
𝜀1D
𝜺1D =̂ ⎜ 33 ⎟ , (12) the 1D and 2D constraint due to Poisson’s effect, is
⎜ 0 ⎟
⎜ 0 ⎟ ⎛ −36 ⎞
⎜ ⎟ ⎜ −36 ⎟
⎝ 0 ⎠ ⎜ ⎟
[ ] −36 ⎟
since the system can expand freely in 𝒆𝑦 - and in 𝒆𝑧 -direction. Both 𝝈 3D = ℂ 𝜺3D − 𝜺0 =̂ ⎜ 𝑐 GPa, 𝑚3D = −108 GPa. (17)
⎜ 0 ⎟ H
strains follow from the stress free conditions 𝜎22
1D = 𝜎 1D = 0, and the ⎜ 0 ⎟
resulting stress state is
33
⎜ ⎟
⎝ 0 ⎠

⎛ −10 ⎞ The stresses of the constrained palladium-hydrogen system will strongly


⎜ 0 ⎟ affect the stabilities of the solid solution phase and the hydride phase.
[ 1D ] ⎜ 0 ⎟ This becomes apparent by re-writing Eq. (5) as
𝝈 = ℂ 𝜺 − 𝜺0 =̂ ⎜
1D ⎟ 𝑐 GPa, 𝑚1D = −10 GPa. (13) ( )
⎜ 3 ⎟ H 𝑐H ( )
⎜ 0 ⎟ 𝜇H = 𝜇0 + 𝑅𝜃 ln − 𝜂H 𝜈0 𝑚XD + 𝐸HH 𝑐H . (18)
⎜ ⎟ 𝑟 − 𝑐H
⎝ 0 ⎠
Hence, the monotony of the chemical potential and thus the stability
The non-vanishing shear stress is a result of the cubic stiffness, cf.
ranges of the phases in the (𝜃, 𝑐H )-space will be determined by the
Eq. (10), and the non-spherical total strain. However, as the chemical
relation of 𝐸HH and the stress-dependent parameter 𝑚XD . This param-
potential is only modified by the normal stresses, it does not influence
eter describes the stress-induced de-stabilization of the hydride phase.
the following discussions.
This is illustrated in Fig. 2 for the chemical potential of the palladium-
hydrogen system with the different constraint conditions at 𝜃 = 300 K .
2D constraint For the 2D constrained system in Fig. 1 c), with con- According to Eq. (7), as long as 𝜕𝜇H ∕𝜕𝑐H ≤ 0, there exists a critical tem-
straints in 𝒆𝑥 - and 𝒆𝑦 -direction resembling e.g. a thin film with height perature 𝜃crit , below which the chemical potential is a non-monotonic
much smaller than the lateral dimension, that adheres to a rigid sub- function in increasing 𝑐H . Then a thermodynamic driving force for the
strate, the total strain is formation of the hydride phase exists. This statement is equivalent to

3
A. Dyck, T. Böhlke, A. Pundt et al. Scripta Materialia 247 (2024) 116117

Fig. 2. a) Chemical potential of the linear-elastic palladium-hydrogen system with different constraint conditions as a function of the hydrogen concentration 𝑐H at
𝜃 = 300 K . For the free system (0D constraint) and the systems with 1D and 2D constraints, at 300 K , the chemical potential is a non-monotonous function, and hence
there is a driving force for hydride formation. For 3D constraint, on the other hand, the solid solution phase is stable in the whole range of hydrogen concentrations.
In detail, the overswing of the chemical potential becomes smaller with increasing constraints, revealing the stress-induced de-stabilization of the hydride phase in
constrained systems. b) The derivative of the chemical potential with respect to the hydrogen concentration 𝑐H , indicating the overswing of the chemical potential,
even in the 2D constrained system.

Fig. 3. a) Chemical potential of the linear elastic palladium-hydrogen system with 2D constraints, hydrogen-hydrogen interaction strength parameter 𝐸HH =
36.8 kJ∕mol and stress-contribution 𝜂H 𝜈0 𝑚2D = −20.28 kJ∕mol in Eq. (18), plotted for different temperatures. A critical temperature 𝜃crit = 307 K exists, below which
the chemical potential becomes a non-convex function, yielding a thermodynamic driving force for hydride formation. b) Derivative of the chemical potential as a
function of concentration. (For interpretation of the colors in the figure(s), the reader is referred to the web version of this article.)

sp sp
Fig. 4. a) Chemical potential as a function of the hydrogen concentration and definition of the spinodal concentrations 𝑐𝛼 and 𝑐𝛼 ′ as well as the related equilibrium
concentrations 𝑐𝛼max and 𝑐𝛼min
′ of the phases. b) Two-phase fields of the Pd-H system for different constraint conditions. The spinodal (dotted lines) and the equilib-
rium (solid lines) miscibility gaps are plotted. With increasing dimensionality of the constraint conditions the miscibility gaps shrink on the temperature and the
concentration axes.

the free energy density being non-convex with respect to the concentra- to Fig. 2, at room temperature even in a 2D elastically constrained
tion. palladium-hydrogen system hydride formation is predicted. This de-
For the different constraint conditions the parameters 𝑚XD and the picts the ideal condition of thin films. These results distinguish the
resulting critical temperatures 𝜃crit are summarized in Table 1. Appar- palladium-hydrogen system from the other widely studied model sys-
ently, up to the 2D constraint it follows 𝜃crit > 𝜃room and, according tem of niobium-hydrogen, where the stress-impact upon 2D constraints

4
A. Dyck, T. Böhlke, A. Pundt et al. Scripta Materialia 247 (2024) 116117

Table 1
H-H interaction parameter 𝐸HH initiating hydride formation, parameter 𝑚 describing the
stress-induced de-stabilization of the hydride phase, and resulting critical temperature 𝜃crit
for the ideal palladium-hydrogen system with linear-elastic material behavior and varying
constraints.

Constraint 𝐸HH in kJ∕mol 𝑚 in GPa 𝜂H 𝜈0 𝑚 in kJ∕mol 𝜃crit in K


0D 36.8 0 0 686
1D 36.8 −10 −5.63 581
2D 36.8 −36 −20.28 307
3D 36.8 −108 −60.84 no hydride formation possible

suppresses hydride formation at room temperature. This is considered CRediT authorship contribution statement
in detail in a separate publication.
For the 2D constrained palladium-hydrogen system the chemical poten- Alexander Dyck Conceptualization, Methodology, Formal analysis,
tial and its derivative with 𝑐H are shown in Fig. 3 for 𝜃 = 200 K, 𝜃 = 𝜃crit Investigation, Visualization, Writing – original draft.
and 𝜃 = 400 K . Apparently, for 𝜃 = 200 K (blue curve) the chemical po- Thomas Böhlke: Discussion, Resources, Funding acquisition, Writing –
tential is a non-convex function of hydrogen concentration, enabling review & editing.
Astrid Pundt: Discussion, Resources, Funding acquisition, Writing – re-
two-phase equilibrium. For the critical temperature 𝜃crit = 307 K , the
view & editing.
chemical potential becomes a monotonically increasing function of 𝑐H
Stefan Wagner: Conceptualization, Methodology, Validation, Formal
(red curve) with positive derivative, cf. Fig. 3 b).
analysis, Writing – original draft.
For the palladium-hydrogen system with 3D constraints, on the con-
trary, evaluating Eq. (7) yields 𝜃crit < 0 K and thus hydride formation Declaration of competing interest
is suppressed at any temperature. This resembles the result of Ale-
feld et al. [30], yielding the absence of phase transformation in a 3D The authors declare that they have no known competing financial
constrained palladium-hydrogen system. We note that the calculated interests or personal relationships that could have appeared to influence
critical temperature of the unconstrained Pd-H system of 686 K is larger the work reported in this paper.
than the experimental value of 563 K [22]. This is a known issue of the
theory [1], that is often regarded for by an artificial reduction of the pa- Acknowledgements
rameter 𝑟 in the classical version of Eq. (18) with 𝑚 = 0. However, this
result primarily reveals the frontiers of the models describing the ther- AD and TB gratefully acknowledges partial funding by The Karlsruhe
modynamics of metal-hydrogen systems. It might be more constructive Institute of Technology (KIT) within the EXU funding “KIT Future Field-
s”, Grant ACDC.
to address possible H-concentration dependencies of the elastic and the
electronic interactions.
sp Appendix A. Supplementary material
For 𝜃 ≤ 𝜃crit critical spinodal concentrations 𝑐H result, where phase
transformation is feasible from a thermodynamic point of view accord- Supplementary material related to this article can be found online
ing to Eq. (5) and Fig. 3 b). As stated above, Maxwell-constructions of at https://doi.org/10.1016/j.scriptamat.2024.116117.
the equilibrium chemical potentials for the respective conditions can
then be used to determine the equilibrium concentrations of the co- References
existing solid-solution 𝑐𝛼max and the hydride phase 𝑐𝛼min ′ . The spinodal
and the equilibrium two-phase fields of the systems with 0D, 1D and [1] S. Wagner, A. Pundt, Quasi-thermodynamic model on hydride formation in
palladium–hydrogen thin films: impact of elastic and microstructural constraints,
2D constraints are shown in Fig. 4 b). Apparently, the miscibility gap
Int. J. Hydrog. Energy 41 (4) (2016) 2727–2738.
areas shrink with increasing dimensionality of the constraints. [2] S. Wagner, T. Kramer, H. Uchida, P. Dobron, J. Cizek, A. Pundt, Mechanical stress
To conclude, in this work the stability of the solid solution phase of and stress release channels in 10–350 nm palladium hydrogen thin films with dif-
ferent micro-structures, Acta Mater. 114 (2016) 116–125.
the palladium-hydrogen system with different elastic constraint condi-
[3] R. Schwarz, A. Khachaturyan, Thermodynamics of open two-phase systems with
tions and the resulting de-stabilization of the hydride phase are system- coherent interfaces, Phys. Rev. Lett. 74 (13) (1995) 2523.
atically considered within a chemo-mechanical approach. It is shown [4] A. Baldi, M. Gonzalez-Silveira, V. Palmisano, B. Dam, R. Griessen, Destabilization
that the driving force for hydride formation decreases with progres- of the Mg-H system through elastic constraints, Phys. Rev. Lett. 102 (22) (2009)
226102.
sively increasing dimensionality of the constraints. Concomitantly, the [5] E. Arzt, Size effects in materials due to microstructural and dimensional constraints:
widths of the spinodal and the equilibrium two-phase coexistence re- a comparative review, Acta Mater. 46 (16) (1998) 5611–5626.
gions shrink, with increasing solid solution limit, decreasing upper limit [6] M. Armand, J.-M. Tarascon, Building better batteries, Nature 451 (7179) (2008)
652–657.
of the two-phase field and reduced critical temperature. The model pre- [7] W. Nix, B. Clemens, Crystallite coalescence: a mechanism for intrinsic tensile stresses
dicts a driving force for hydride formation for the free system as well in thin films, J. Mater. Res. 14 (8) (1999) 3467–3473.
as for systems with 1D and 2D constraints at room temperature. For 3D [8] J. Eshelby, The continuum theory of lattice defects, Solid State Phys. 3 (1956)
79–144, Elsevier.
constraints suppressing the H-induced volume expansion of palladium
[9] V. Burlaka, S. Wagner, M. Hamm, A. Pundt, Suppression of phase transformation in
in all directions, hydride formation is entirely suppressed. Nb–H thin films below switchover thickness, Nano Lett. 16 (10) (2016) 6207–6212.
Different from the ideal linear-elastic systems stress-relaxation is often [10] R. Griessen, N. Strohfeldt, H. Giessen, Thermodynamics of the hybrid interaction of
hydrogen with palladium nanoparticles, Nat. Mater. 15 (3) (2016) 311–317.
observed in constrained systems in experiments above critical hydrogen
[11] S. Wagner, H. Uchida, V. Burlaka, M. Vlach, M. Vlcek, F. Lukac, J. Cizek, C. Baehtz,
concentrations [1,2,15,16,25,31]. This can be supported by hydrogen A. Bell, A. Pundt, Achieving coherent phase transition in palladium–hydrogen thin
acting as a defactant, reducing the formation energy of defects such as films, Scr. Mater. 64 (10) (2011) 978–981.
[12] J. Weissmuller, C. Lemier, On the size dependence of the critical point of nanoscale
grain boundaries, dislocations and vacancies [32]. Stress relaxation and
interstitial solid solutions, Philos. Mag. Lett. 80 (6) (2000) 411–418.
its impact on the systems thermodynamics will be considered in a sep- [13] F. Larche, J.W. Cahn, The interactions of composition and stress in crystalline solids,
arate publication. J. Res. Natl. Bur. Stand. 89 (6) (1984) 467.

5
A. Dyck, T. Böhlke, A. Pundt et al. Scripta Materialia 247 (2024) 116117

[14] C. Lemier, J. Weissmüller, Grain boundary segregation, stress and stretch: effects [23] Y. Fukai, The Metal-Hydrogen System: Basic Bulk Properties, vol. 21, Springer Sci-
on hydrogen absorption in nanocrystalline palladium, Acta Mater. 55 (4) (2007) ence & Business Media, 2006.
1241–1254. [24] S. Wagner, P. Klose, V. Burlaka, K. Nörthemann, M. Hamm, A. Pundt, Structural
[15] R. Gremaud, M. Gonzalez-Silveira, Y. Pivak, S. De Man, M. Slaman, H. Schreuders, phase transitions in niobium hydrogen thin films: mechanical stress, phase equilibria
B. Dam, R. Griessen, Hydrogenography of PdHx thin films: influence of H-induced and critical temperatures, ChemPhysChem 20 (14) (2019) 1890–1904.
stress relaxation processes, Acta Mater. 57 (4) (2009) 1209–1219. [25] K. Nörthemann, A. Pundt, Coherent-to-semi-coherent transition of precipitates in
[16] S. Wagner, A. Pundt, Mechanical stress impact on thin Pd1-xFex film thermody- niobium-hydrogen thin films, Phys. Rev. B 78 (1) (2008) 014105.
namic properties, Appl. Phys. Lett. 92 (5) (2008). [26] J. Cahn, F. Larché, A simple model for coherent equilibrium, Acta Metall. 32 (11)
[17] S. Wagner, A. Pundt, Combined impact of microstructure and mechanical stress (1984) 1915–1923.
on the electrical resistivity of PdHc thin films, Acta Mater. 59 (5) (2011) [27] K. Binder, Theory of first-order phase transitions, Rep. Prog. Phys. 50 (7) (1987)
1862–1870. 783.
[18] L. Mooij, B. Dam, Hysteresis and the role of nucleation and growth in the hydro- [28] M.E. Gurtin, E. Fried, L. Anand, The Mechanics and Thermodynamics of Continua,
genation of Mg nanolayers, Phys. Chem. Chem. Phys. 15 (8) (2013) 2782–2792. Cambridge University Press, 2010.
[19] A. Baldi, L. Mooij, V. Palmisano, H. Schreuders, G. Krishnan, B.J. Kooi, B. Dam, R. [29] E. Wicke, H. Brodowsky, H. Züchner, Hydrogen in palladium and palladium alloys,
Griessen, Elastic versus alloying effects in Mg-based hydride films, Phys. Rev. Lett. in: Hydrogen in Metals II: Application-Oriented Properties, 2005, pp. 73–155.
121 (25) (2018) 255503. [30] G. Alefeld, Phase transitions of hydrogen in metals due to elastic interaction, Ber.
[20] N. Patelli, M. Calizzi, L. Pasquini, Interface enthalpy-entropy competition in Bunsenges. Phys. Chem. 76 (8) (1972) 746–755.
nanoscale metal hydrides, Inorganics 6 (1) (2018) 13. [31] Y. Pivak, H. Schreuders, M. Slaman, R. Griessen, B. Dam, Thermodynamics, stress
[21] Y. Manassen, H. Realpe, D. Schweke, Dynamics of H in a thin Gd film: evidence of release and hysteresis behavior in highly adhesive Pd–H films, Int. J. Hydrog. Energy
spinodal decomposition, J. Phys. Chem. C 123 (18) (2019) 11933–11938. 36 (6) (2011) 4056–4067.
[22] H. Peisl, Lattice strains due to hydrogen in metals, in: Hydrogen in Metals I: Basic [32] R. Kirchheim, On the solute-defect interaction in the framework of a defactant con-
Properties, 2005, pp. 53–74. cept, Int. J. Mater. Res. 100 (4) (2009) 483–487.

You might also like