Thermodynamics of Nanophase Manganese Oxides: Sodium, Potassium, and Calcium Birnessite and Cryptomelane
Thermodynamics of Nanophase Manganese Oxides: Sodium, Potassium, and Calcium Birnessite and Cryptomelane
net/publication/313014597
CITATIONS READS
15 255
2 authors:
Some of the authors of this publication are also working on these related projects:
All content following this page was uploaded by Alexandra Navrotsky on 13 June 2018.
Contributed by Alexandra Navrotsky, December 21, 2016 (sent for review October 10, 2016; reviewed by R. Lee Penn and Jeffrey Post)
Manganese oxides with layer and tunnel structures occur widely state are important for stabilizing these minerals and enhancing
in nature and inspire technological applications. Having variable their functionality.
compositions, these structures often are found as small particles There have been a number of qualitative observations of
(nanophases). This study explores, using experimental thermo- poorly understood phenomena concerning phase formation
chemistry, the role of composition, oxidation state, structure, and/or transformation among birnessite and cryptomelane. (i)
and surface energy in the their thermodynamic stability. The mea- All natural and synthetic birnessite and cryptomelane phases
sured surface energies of cryptomelane, sodium birnessite, potas- appear to have small particle size, less than 100 nm and often
sium birnessite and calcium birnessite are all significantly lower
much smaller, at their initial formation. However, on the geo-
logic time scale, these natural minerals are able to grow in size
than those of binary manganese oxides (Mn3O4, Mn2O3, and
(coarsen) to form particles of larger dimensions, particularly
MnO2), consistent with added stabilization of the layer and tunnel within the layers. Such growth often involves complex chemical
structures at the nanoscale. Surface energies generally decrease reactions with changes in cation content, hydration, and oxida-
with decreasing average manganese oxidation state. A stabilizing tion state. (ii) Mixed sodium–potassium birnessite often forms in
enthalpy contribution arises from increasing counter-cation con- synthesis from aqueous solution, even when one is trying to make
tent. The formation of cryptomelane from birnessite in contact the ion-exchanged potassium form. (iii) Synthesis of calcium
with aqueous solution is favored by the removal of ions from the birnessite most successfully begins with sodium birnessite as the
layered phase. At large surface area, surface-energy differences starting phase followed by exchange of calcium for sodium
make cryptomelane formation thermodynamically less favorable under aqueous conditions. (iv) Cryptomelane exists initially
than birnessite formation. In contrast, at small to moderate sur- only within a limited size range, with surface area ∼50–350 m2/g
face areas, bulk thermodynamics and the energetics of the aque- (2, 11), when synthesized in the laboratory under ambient
ous phase drive cryptomelane formation from birnessite, perhaps conditions. Again, given enough time and thermodynamic
aided by oxidation-state differences. Transformation among birnes-
driving forces, macroscale size may be achieved, especially in
natural materials. (v) The synthesis of cryptomelane in an
site phases of increasing surface area favors compositions with
aqueous medium appears to begin with potassium birnessite
lower surface energy. These quantitative thermodynamic find- formation and then requires leaching of potassium from the
ings explain and support qualitative observations of phase- birnessite layers before spontaneous formation of the tunnel
transformation patterns gathered from natural and synthetic structure (11).
manganese oxides. In addition to kinetic controls on reaction rates, these observa-
tions must have fundamental thermodynamic underpinning, and yet
manganese oxides | birnessite | cryptomelane | calorimetry | the thermodynamic properties of these phases as a function of
thermodynamics composition and particle size are poorly known. Some calorimetric
studies of enthalpies of formation of layer and tunnel manganese
oxides have been reported (12), although their compositions in
C omplex manganese oxides, highly reactive fine-grained ma-
terials ubiquitous in nature, have served in a number of
important capacities to benefit both the Earth and human society
terms of AOS of manganese, and therefore oxygen content, were
not measured and their particle sizes and surface areas were not
characterized. These factors have recently been shown to be im-
(1). These oxides have profoundly affected Earth’s critical zone portant in the thermodynamics of binary oxides in the Mn–O system
throughout geologic time, influencing the evolution of the at- (13) and probably play an equal or more important role in the
mosphere and the bioinorganic chemistry of organisms. They are complex layer and tunnel materials.
effective in accumulation and recovery of economic ores and
have recently inspired development of novel catalysts for green Significance
chemistry and energy sustainability technologies. Minerals based
on MnO2 include multiple classes, two of which are well known
Manganese oxides with layer and tunnel structures occur widely
to geochemists: the 2 × 2 tunnel structure hollandite (the po-
in the natural environment and inspire technological applications.
tassium - bearing variety is cryptomelane) and the layered struc-
In addition to having variable compositions, these structures often
ture phyllomanganates (e.g., birnessite) (2, 3). Layered Mn oxides, are found as small particles (nanophases). This study explores,
such as the mineral birnessite, derived from MnO2 by inclusion of using experimental thermochemistry, the role of composition,
cations and water, with concomitant decrease in manganese av- oxidation state, structure, and surface energy in the thermody-
erage oxidation state (Mn AOS) from 4 to typical values between namic stability of synthetic birnessites and cryptomelane. These
3.5 and 3.9, with manganese in tetravalent, trivalent and some- quantitative thermodynamic findings help explain and support
times divalent oxidation states, are among the most important Mn qualitative observations of phase-transformation patterns gath-
oxides in nature (4). These “nanosheet” Mn oxides readily trans- ered over the past five decades.
form among phases, influencing crucial environmental and tech-
nological processes, including biogeochemical cycles (1, 4, 5) water Author contributions: N.B. and A.N. designed research; N.B. performed research; N.B. and
oxidation catalysis (6–8) and radionuclide confinement (9, 10). A A.N. analyzed data; and N.B. and A.N. wrote the paper.
recent thermochemical study of layered calcium manganese ox- Reviewers: R.L.P., University of Minnesota; and J.P., Smithsonian Institution.
ides (6) relevant to water oxidation catalysis (7, 8) suggests that The authors declare no conflict of interest.
low surface energy and energetically accessible changes in oxidation 1
To whom correspondence should be addressed. Email: [email protected].
Birnessite synthesis*
H+; HCl 1.0 × e−2 0.9 1 0.01 0.717 1.421
Cl−; HCl 1.0 × e−2 0.3 1 0.01 0.470 1.421
Na+; NaCl 3.015 0.4 1 3.015 0.536 1.421
K+; KCl 3.015 0.3 1 3.015 0.470 1.421
Cl−; KCl 3.015 0.3 1 3.015 0.470 1.421
Na+; NaMnO4 0.125 0.4 1 0.125 0.536 1.421
K+; KMnO4 0.125 0.3 1 0.125 0.470 1.421
MnO4− 0.125 0.35 1 0.125 0.505 1.421
Cryptomelane synthesis†
H+; H2SO4 1.0 0.9 1 1 0.707 3.105
HSO4−; H2SO4 1.0 0.4 4 4 0.067 3.105
K+; KCl 3.0 0.3 1 3 0.435 92.933
Cl−; KCl 3.0 0.3 1 3 0.435 3.105
K+; KMnO4 0.125 0.3 1 0.13 0.435 3.105
MnO4−; KMnO4 0.125 0.35 1 0.13 0.475 3.105
Reagent parameters to calculate total ionic strength include: concentration [Ci] in mol/L of the reagent ions
(i) with size parameter (di) and ionic charge (z). Debye–Hückel activity coefficient (γ) is computed from the
Debye–Hückel equation, and the Davies activity coefficient is computed from Davies equation.
*Final ionic strength: 1/2 Σz2c = 3.150.
CHEMISTRY
†
Final ionic strength: 1/2 Σz2c = 5.625.
The present work concentrates on measurements of en- The measured enthalpy associated with dropping a sample
thalpies of formation and surface energies of synthetic birnessite from room temperature into molten oxide solvent at high tem-
and cryptomelane phases to explain these frequently observed perature, where it dissolves, is called the enthalpy of drop solu-
phase formation and transformation phenomena on the basis that tion, ΔHds. The difference between ΔHds of reactants and
they are thermodynamically driven. Because the initially formed products (of the same composition) gives the enthalpy of re-
materials are fine grained, indeed nanophase, surface energies action. Thus, the enthalpy associated with an increase in surface
play a large role in selective stabilization of different phases and area can be obtained by comparing ΔHds for samples of different
are emphasized in this study. surface areas. Additionally, because the nanophase samples have
different water contents, ΔHds is corrected to account for their
Results surface-associated water, yielding a corrected drop solution en-
Table 1 summarizes the synthesis parameters of the samples. thalpy, ΔHds,corr. The correction subtracts the heat content of n
Table 2 summarizes the sample compositions, the manganese Mn moles of bulk H2O (considered as physisorbed water) from the
AOS, and the surface areas. For each of the suites of nanophase observed enthalpy of drop solution. Values of the directly mea-
samples, the compositions were controlled remarkably well, with sured enthalpy of drop solution, ΔHds, and the drop solution
cation content and Mn AOS virtually constant. The actual values enthalpy values corrected for water content, ΔHds,corr, are listed
of these two parameters reflected the synthesis conditions and in Table 2. As discussed previously (18–25), the slope of the line
could neither be predicted beforehand nor easily explained af- relating the water-corrected drop solution enthalpy, ΔHds,corr,
terward. It is interesting that similar aqueous synthesis parameters (Table 2) and nanophase surface area (Table 2) yields the surface
(except for the type of cation present) at constant ionic strength enthalpy (SE) (Table 3), which is essentially equivalent to surface
energy (26, 27) for the hydrous surface (Fig. 2 A–D). Birnessite and
produced similar cation content within the sodium and calcium
cryptomelane have substantially lower SE (Fig. 2 A–D) than the bi-
birnessite sample suites, but exactly twice the cation content was
nary Mn oxide phases previously studied (13), hausmannite, (Mn3O4,
taken up by potassium birnessites. The reasons for this effect are
SE = 0.96 ± 0.08 J/m2), bixbyite (Mn2O3; SE = 1.29 ± 0.10 J/m2), and
unclear. Of further note is that the AOS in the birnessites is always pyrolusite (MnO2; SE = 1.64 ± 0.10 J/m2). The low SE values suggest
significantly lower (more reduction of Mn4+ to Mn3+ or possibly that water is not strongly bound to the surfaces, interlayers, or
Mn2+) than simply required for charge balancing the alkali or tunnels, because water-adsorption enthalpy scales with the SE for
alkaline earth ions. This observation implies the presence of other many oxides (18–25, 28). There appears to be a small decrease in
charge balance mechanisms, such as the incorporation of protons SE with decreasing Mn AOS (Fig. 3), that is, with an increase in
under the acidic conditions of synthesis. The observation also mixed valence manganese (Mn3+/4+).
implies close coupling between the solution ion content, the so- Enthalpies of formation (Table 2) were computed using
lution acidity, and the oxygen fugacity to determine composition thermochemical cycles and auxiliary data. For the nanophase
and thermodynamic properties of the solid materials formed. complex Mn oxides, enthalpy of formation from the oxides can
Thus, the present study covers only a small range of possible be calculated in two ways. The first (ΔHof-ox*) includes the en-
conditions and compositions. thalpy contribution from redox effects taking mAc+O, Mn2O3,
X-ray diffraction (XRD) patterns of cryptomelane samples H2O, and O2 as reactants. This formation reaction is Eq. 1:
(Fig. 1A) are consistent with previous reports for the potassium-
bearing 2 × 2 tunnel structure hollandite (14, 15). XRD patterns ½mAc+ O + ð0.5ÞMn2 O3 ðs,25°CÞ + ð0.25ÞyðO2 Þðg,25°CÞ + nH2 Oð1,25°CÞ
of birnessite samples (Fig. 1B), similar to hexagonal acid birnessite,
= ½mAc+ O · ð0.5ÞMn2 O3 · ð0.25ÞyO2 · nH2 Oðs,25°CÞ.
demonstrate common features of birnessite with a lack of long-range
order between layers, peaks tailing toward higher angle two-theta, [1]
and asymmetric profiles (3, 16, 17). Because crystallography was not
the focus of the present work, a further detailed structural study was Coefficient “m,” defined as m = c+/2, where the cation charge, c+,
not pursued. accounts for moles of cation as well as the number of oxygen
Birkner and Navrotsky PNAS | Published online January 27, 2017 | E1047
Table 2. Composition, surface area, Mn AOS, wt% water loss, drop solution enthalpy, water-corrected drop solution enthalpy,
formation enthalpy from the oxides (without and with oxidation reaction, and from the elements), and standard entropy for
the nanophase samples
BET surface area,† m2/g TGA weight ΔHds,corr,¶ ΔH°f-ox,# ΔH°f-ox*,,jj ΔH°f,** S°298,††
Composition* (m2/mol) Mn AOS loss,‡ [wt %] ΔHds,§ kJ/mol kJ/mol kJ/mol kJ/mol kJ/mol J/mol·K
Na0.09MnO1.815·0.64H2O 94.64 ± 0.69 (9,295.45) 3.54 ± 0.01 12.52 193.91 ± 2.01 (10) 149.68 ± 2.65 −56.01 ± 2.11 −79.08 ± 3.82 −703.29 102.82 ± 0.09
Na0.08MnO1.82·0.42H2O 58.72 ± 0.18 (5,526.83) 3.56 ± 0.01 8.59 179.79 ± 1.85 (8) 150.71 ± 2.51 −55.09 ± 1.96 −78.92 ± 3.64 −639.19 86.97 ± 0.08
Na0.08MnO1.825·0.54H2O 28.30 ± 0.20 (2,726.47) 3.57 ± 0.01 10.81 191.65 ± 1.66 (8) 154.53 ± 2.21 −58.19 ± 1.78 −82.66 ± 3.34 −673.10 95.33 ± 0.09
Na0.09MnO1.82·0.47H2O 18.52 ± 0.16 (1,768.78) 3.55 ± 0.01 9.69 187.09 ± 1.35 (8) 154.35 ± 1.13 −60.45 ± 1.49 −83.69 ± 2.40 −656.17 90.88 ± 0.08
K0.21MnO1.87·0.49H2O 34.63 ± 0.44 (3,595.24) 3.53 ± 0.01 11.35 154.25 ± 1.93 (10) 120.47 ± 2.73 −51.23 ± 2.05 −74.40 ± 3.88 −679.98 98.86 ± 0.08
K0.21MnO1.90·0.47H2O 28.74 ± 0.21 (2,991.87) 3.59 ± 0.01 11.55 154.06 ± 1.90 (8) 121.29 ± 2.69 −49.57 ± 2.03 −75.27 ± 3.84 −678.34 97.21 ± 0.08
K0.21MnO1.85·0.23H2O 18.50 ± 0.17 (1,823.77) 3.49 ± 0.01 6.25 137.18 ± 1.88 (8) 121.33 ± 2.66 −54.00 ± 1.99 −74.95 ± 3.81 −603.22 80.82 ± 0.07
K0.20MnO1.84·0.13H2O 15.41 ± 0.10 (1,493.37) 3.48 ± 0.01 4.42 130.86 ± 1.39 (8) 121.91 ± 1.97 −53.46 ± 1.53 −76.23 ± 3.11 −574.89 73.39 ± 0.07
Ca0.12MnO1.825·0.81H2O 97.76 ± 0.87 (10,383.96) 3.53 ± 0.01 15.62 181.29 ± 1.88 (10) 125.71 ± 2.66 −33.63 ± 1.99 −55.44 ± 3.81 −817.68 115.95 ± 0.11
Ca0.11MnO1.79·0.73H2O 55.42 ± 0.51 (5,653.72) 3.47 ± 0.01 11.92 177.40 ± 1.73 (8) 127.04 ± 2.45 −37.21 ± 1.83 −57.08 ± 3.58 −777.27 110.21 ± 0.11
Ca0.12MnO1.82·0.66H2O 34.63 ± 0.44 (3,552.24) 3.52 ± 0.01 12.10 174.49 ± 1.29 (8) 128.80 ± 1.83 −37.66 ± 1.44 −58.86 ± 2.98 −770.26 105.49 ± 0.10
Ca0.12MnO1.80·0.67H2O 18.50 ± 0.17 (1,877.27) 3.48 ± 0.01 11.10 175.30 ± 1.19 (8) 128.88 ± 1.69 −39.68 ± 1.33 −59.49 ± 2.85 −765.85 106.35 ± 0.10
K0.10MnO1.885·0.43H2O 234.02 ± 0.77(22,852.59) 3.67 ± 0.01 11.88 124.58 ± 1.88 (10) 94.77 ± 2.66 −1.56 ± 2.02 −29.78 ± 3.81 −648.54 88.91 ± 0.07
K0.11MnO1.955·0.41H2O 150.63 ± 0.16 (14,842.18) 3.80 ± 0.01 12.59 132.71 ± 1.75 (8) 104.38 ± 2.48 −6.49 ± 1.94 −39.39 ± 3.61 −648.44 87.45 ± 0.07
K0.11MnO1.96·0.36H2O 128.05 ± 0.14 (12,503.15) 3.81 ± 0.01 11.69 129.34 ± 1.17 (8) 104.68 ± 1.66 −6.09 ± 1.45 −39.69 ± 2.83 −633.54 83.90 ± 0.07
K0.10MnO1.965·0.28H2O 58.00 ± 0.25 (5,581.71) 3.83 ± 0.01 10.38 128.04 ± 1.01 (8) 109.02 ± 1.44 −7.77 ± 1.34 −43.50 ± 2.64 −610.76 77.75 ± 0.07
atoms entering into the reaction. Coefficient “n” corresponds to thermodynamic components thermodynamic components, which
moles of water, and “y” corresponds to moles of molecular oxy- can be taken as alkali/alkaline earth oxide, Mn 2 O 3 , MnO 2 ,
gen incorporating into the solid phase as a result of oxidation. and H 2 O or alkali/alkaline earth oxide, Mn 2 O 3 , O 2 , and
The second method of calculation eliminates the redox effects H 2 O, although the number of distinct species present can be
(ΔHof-ox) by taking a mixture of Mn2O3 and MnO2 (and no O2) much greater.
as reactants (e.g., mAc+O, Mn2O3, MnO2, and H2O). This for- Standard formation enthalpies from the elements (ΔH°f)
mation reaction is Eq. 2: (Table 2) were calculated from values of formation enthalpy
from the oxides by adding the standard formation enthalpies of
½mAc+ O + ð0.5Þð1-yÞMn2 O3 + yMnO2 ðs,25°CÞ + nH2 Oð1,25°CÞ binary (simple) oxides and water (29) in the appropriate reaction
[2] stoichiometry for each of the nanophase samples.
= ½mAc+ O · ð0.5Þð1-yÞMn2 O3 · yMnO2 · nH2 Oðs,25°CÞ.
For a given cation, an increase in the cation content (with
concomitant decrease in AOS) has a thermodynamically stabi-
Here, the coefficient y, normalized to one mole of manga-
nese, accounts for the Mn(III) and Mn(IV) stoichiometry reflected lizing influence on the enthalpy of formation. This influence is
in the Mn oxidation state speciation in the samples (e.g., coeffi- seen most clearly for the potassium birnessites, where the values
cients, “1-y” and y). The enthalpies of formation from the oxides, of ΔH°f-ox in kilojoules per formula containing 1 mol of man-
ΔHf-ox* and ΔHf-ox, are strongly exothermic, which reflects the ganese are −36.7 for x = 0.125 (12), −52.3 for x = 0.21 (this
nature of strongly basic alkali and alkaline earth metal oxides work), and −60.6 for x = 0.29 (12). The bulk sodium birnessite
combining with relatively acidic manganese (III,IV) oxide species. with x = 0.09 has more negative ΔH°f-ox* and ΔH°f-ox than the bulk
ΔH°f-ox is less exothermic than ΔH°f-ox* because the latter contains potassium birnessite with x = 0.21 (Table 4). Both alkali birnessites are
an intrinsically exothermic contribution from oxidation of Mn3+ more energetically stable than the calcium birnessite with x = 0.12
to Mn4+. (Table 4). The thermodynamic effect of the cation contribution is
Some birnessites also may contain divalent manganese relatively small, as seen in relatively similar values of ΔH°f-ox,
(Mn 2+ ). Because the AOS is between 3 and 4 in all of our where oxidation is not involved. In contrast, the ΔH°f-ox* values are
samples, from a thermodynamic point of view, all reactions can much more exothermic and vary somewhat more strongly with
be balanced using only Mn3+ and Mn4+ phases as written cation type and content, which reflects the contribution of the
above. Although mass and charge balance are essential for oxidation enthalpy. Thus, in considering formation and trans-
proper thermodynamic analysis, further details of speciation formation reactions in both geological and technological settings,
and site occupancy are not necessary because a thermody- the redox energetics can easily compete with the energetic con-
namic description can be fully defined in terms of the bulk tribution made by the cations. Having more thermodynamically
composition of the system and intensive parameters such favorable standard formation enthalpies (Table 2), enthalpies of
as temperature, pressure, and oxygen fugacity. Similarly, al- formation from the oxides (Table 2), and lower SE (Table 3), the
though knowing the exact water content is crucial for mass birnessites (layer structure) are more stable relative to binary
balance, its speciation (molecular H2O, H+, and OH−) and lo- oxides than the cryptomelane (tunnel structure). Fritsch et al. (12)
cation need not be considered for the present thermodynamic also observed that birnessites are more thermodynamically stable
analysis. Thus, the solid phase is completely defined by four than tunnel structure Mn oxides.
Discussion
As observed for other layered Mn oxides (6), the very low SE
CHEMISTRY
values are not dramatically different between birnessite type
phases, but these differences are enough to influence the
overall phase-transformation energetics as the surface area
dramatically increases or decreases. All synthetic birnessite and
cryptomelane phases and natural forms that are present in re-
active environments (e.g., soils and sediments) have small particle
sizes (are not single crystals and may often be considered as-
semblages of nanoparticles). This finding is reflected in the
current study in these oxides’ synthesis by precipitation in an
aqueous environment near room temperature and by their
rather small surface energies, which provide little thermody-
namic driving force for coarsening. This result indicates that
additional thermodynamic perturbations are needed for
coarsening (crystal growth) to occur and that kinetics also
plays a role in their growth. To be clear, however, initial for-
mations and transformations of these phases are thermody-
namically controlled and occur among phases of small particle
Fig. 1. XRD powder patterns of the tunnel structure phase, cryptomelane size according to this study. Although the thermochemical
(K-cryptomelane) (A), and the birnessites (B), which are labeled as follows: A, data do not permit complete calculations of equilibria in-
sodium birnessite (Na-Birnessite); B, potassium birnessite (K-Birnessite); and volving different cation content, Mn AOS, and particle size, a
C, calcium birnessite (Ca-Birnessite). The birnessite powder patterns are few illustrative examples using the current data are useful.
y-offset to improve readability. arb, arbitrary units. Synthesis of pure phase calcium birnessite (CaBi) successfully
begins with sodium birnessite (NaBi) as the starting phase and
then exchanging calcium for sodium by shaking (or stirring) NaBi
The standard entropies (S°298) of the complex manganese in a concentrated solution of aqueous CaCl2 (this work and refs.
oxides are not generally available. As discussed previously (30), 14 and 31). This method reflects the lower stability of the calcium
it is reasonable to approximate them as the weighted sum of form, which may hinder its initial formation relative to other
standard entropies of binary oxides. Then the entropy change phases, but, once the structure forms, ion exchange is possible.
Cryptomelane: CR2, NaBi, KBi, and CaBi (Bi, birnessite) are from the present work.
*Surface enthalpies are for the hydrous surfaces [SE(hyd)].
†
Samples of Ca–Mn–oxide nanosheets (6).
Birkner and Navrotsky PNAS | Published online January 27, 2017 | E1049
NaBi = CaBiðsolid oxidesÞ : Na0.09 · MnO1.815 · 0.64H2 OðsÞ
+ 0.12 CaOðsÞ + 0.17 H2 OðliqÞ
= Ca0.12 · MnO1.825 · 0.81H2 OðsÞ + ð0.5Þ0.09 Na2 OðsÞ
+ 0.0325 O2ðgÞ
CR2 K0.11·MnO1.94 113.75 ± 3.00 −43.70 ± 3.16 −15.18 ± 3.42 −529.24 ± 0.94 58.32 ± 0.21 −540.31 ± 0.63
NaBi Na0.09·MnO1.82 155.64 ± 3.47 −82.47 ± 3.67 −61.99 ± 4.49 −522.92 ± 0.71 64.13 ± 0.12 −556.00 ± 0.77
KBi K0.21·MnO1.87 122.65 ± 3.58 −71.34 ± 3.78 −52.32 ± 3.42 −536.86 ± 1.14 58.28 ± 0.17 −580.64 ± 0.68
CaBi Ca0.12·MnO1.81 129.80 ± 3.10 −58.22 ± 3.21 −33.86 ± 2.30 −563.26 ± 0.76 58.36 ± 0.11 −546.65 ± 0.80
Cryptomelane: CR2, NaBi, KBi, and CaBi (Bi, birnessite). *Extrapolated drop solution enthalpy for bulk scale samples (no surface hydration and x = 0 surface
area). Estimated standard enthalpy of formation for bulk phases were computed from both oxides, including †redox enthalpy (ΔH°f-ox*) and ‡without redox
energetics (ΔH°f-ox), as well as from the §elements [ΔH°f(bulk)]. Estimated ¶So298 (standard entropy) for bulk phases. #Estimated bulk-scale reaction Gibbs free
energy of formation was computed as G°f(bulk) = ΔH°f(bulk) – TΔS°298 using the estimated standard entropies (S°298).
are present in a birnessite synthesis in the laboratory as pre- the mixed cation content of the aqueous solutions from which
viously reported (32). they form, but the thermodynamic factors discussed above,
especially SE differences, also play a role in the thermody-
namic equilibria. Specifically, although the ΔH°f-ox values
NaBi = KBiðsolid oxidesÞ : Na0.08 MnO1.825 · 0.54H2 OðsÞ become less exothermic in the order Na, K, Ca, the surface
+ ð0.5Þ0.21 K 2 OðsÞ + 0.005 O2ðgÞ energies also decrease in the same order. This finding suggests
a competition between bulk thermodynamics and surface en-
= K 0.21 MnO1.90 · 0.47H2 OðsÞ + ð0.5Þ0.08 Na2 OðsÞ
CHEMISTRY
ergy effects in determining the specific composition formed
+ 0.07 H2 OðliqÞ under a given set of external conditions. Transformation re-
action enthalpy for the bulk phases is computed using the
NaBi = KBiðaqueous ion exchangeÞ : Na0.08 · H0.35 MnO2 · 0.365H2 OðsÞ following equations and the results are shown in Table 5.
+ 0.21 K+ðaqÞ + 0.005 H2 OðliqÞ ΔH°bulk-pt = ΔH°f ,nano-pt − ΔHðsurface termÞ
= K0.21 H0.20 MnO2 · 0.37H2 OðsÞ + 0.08 Na+ðaqÞ + 0.13 H+ðaqÞ .
ΔHðsurface termÞ = ðSE product − SE reactantÞ
Although the exact nature and content of the counter cat- × SSA m2 mol × 1 kJ=1,000 J
ions found in natural birnessites surely depend upon the types
and amounts of cations present in the local environment, this
work demonstrates the thermodynamic stability of pure phase ΔG°bulk-pt = ΔH°nano-pt − ΔHðsurface termÞ − TΔS°298 .
Na-, K-, and Ca-birnessite to lend insight into observed trends
found in the laboratory and the environment. Because we do The phase-transformation enthalpies and free energies among
not have complete thermochemical data for the Na and K nanophase birnessites appear to depend only weakly on Mn AOS
birnessites as a function of alkali content, and because the Mn and cation content. This result is similar to observations for layer
AOS also probably varies as a function of alkali content, structure Ca–Mn–oxide nanosheets which have a constant internal
we cannot completely quantify the thermodynamics of ion enthalpy of oxidation independent of Ca:Mn ratio and Mn AOS (6).
exchange or oxides reacting or separate ion exchange from A comparison of the enthalpies of birnessite transformation reactions
redox reactions. The thermodynamic studies here however from solid oxides and ion-exchange reactions indicates that the latter
support observations that mixed cation (Na–K–Ca) birnessite are more exothermic. Thus, it is likely that such transformations
[e.g., (Na0.3Ca0.1K0.1)(Mn4+,Mn3+)2O4·1.5H2O] is observable occur via the aqueous phase rather than by solid state reaction, being
under laboratory synthesis conditions as well as in environ- favored by both thermodynamic and probably kinetic factors. This
mental conditions (33–35). These synthesis conditions reflect conclusion, that aqueous ion exchange is the prevalent mechanism of
Standard enthalpies, entropies, and free energies for nanophase and bulk transformations involving solid oxides are ΔH°f-ox,pt, ΔS°298(ox), and ΔG°f-ox,pt,
whereas those involving aqueous ion-exchange reactions are ΔH°f,pt, ΔS°298(aq), and ΔG°f,pt.
Birkner and Navrotsky PNAS | Published online January 27, 2017 | E1051
transformation, is consistent with previous birnessite phase-trans- similar AOS, but different particle size controlled by the aging time. Solid
formation studies performed using synchrotron XRD studies (34). compositions were controlled using high concentrations of the desirable
Because all of the birnessites have roughly similar AOS, the counter cation provided by alkali and alkaline earth chlorides. Total ionic
thermodynamic contribution of changing Mn AOS is really only strength was computed using concepts and values obtained from Tissue (37),
important for transformation of birnessites to cryptomelane with Stumm and Morgan (38), and Kielland (39). Ion activities in solution were used
the higher AOS likely the major driving influence. Cryptomelane to calculate the total activity for the concentrations of ionic reagents based on
the Davies equation (40). For the reaction matrices for birnessite synthesis, the
surface energy is relatively low, providing little driving force for
computed total ionic strength was 3.150 and for cryptomelane it was 5.625
coarsening, but it is higher than that of the corresponding K birnessite,
(higher than for birnessite due to the presence of sulfate). Syntheses were
favoring the latter at increasing surface area. These transformations
carried out in covered glass reaction vessels, ionic strengths were held constant,
are represented by the reactions below: and water volume monitored. Synthesis conditions are summarized in Table 1.
All phases were confirmed by powder XRD using a Bruker D8 Advance
KBi = CR2ðsolid oxidesÞ : K 0.21 MnO1.90 · 0.47H2 OðsÞ X-ray diffractometer (Cu Kα radiation; 45 kV and 40 mA) at 25 °C with step
+ 0.06 − O2ðgÞ size 0.02 °2θ and 10-s dwell time. Comparison of reference powder patterns
to profile fits of the samples was made using Jade software (version 6.11;
= K 0.10 MnO1.965 · 0.28H2 OðsÞ + ð0.5Þ0.11 K 2 OðsÞ 2002) and International Center for Diffraction Data.
+ 0.19 H2 OðliqÞ Adsorption isotherms of nitrogen at −196 °C (liquid nitrogen tempera-
ture) were measured using a Micromeritics ASAP 2020 apparatus in the P/P0
range of 0.05–0.3. Before measurements, the samples were outgassed at
KBi = CR2ðaqueous ion exchangeÞ : K0.21 H0.20 MnO2 · 0.37H2 OðsÞ
room temperature under turbo-molecular pump vacuum at room temper-
+ 0.11 H+ðaqÞ + 0.12 O2ðgÞ ature for 24 h to remove most preadsorbed water without losing phase
integrity or surface area. Specific surface area was calculated using the five-
= K0.10 H0.07 MnO2 · 0.245H2 OðsÞ + 0.245 H2 OðliqÞ + 0.11 K+ðaqÞ . point Brunauer–Emmett–Teller (BET) method (41).
For proper interpretation of thermochemical data, one must know the
Phase transformation of potassium birnessite to cryptomelane complete composition of each sample, including the cation content, the Mn
is slightly more favorable with the solid oxides reacting without AOS (these two together then give the oxygen content), and the water
aqueous ion exchange, although ion-exchange reaction is also quite content. A combination of techniques was used to obtain these parameters.
favorable. Two factors may contribute to the energetics of this Metal cation concentration of each sample was measured using inductively
particular reaction. The first is that a main driving force for this coupled mass spectrometry with a quadrupole mass spectrometer (7500a;
transformation may be the push toward higher oxidation state (a Agilent Technologies). Cation contents were determined for 20- to 30-mg
thermodynamically favorable direction) despite the transformation samples digested in 3 mL of 6 M HCl and then diluted with 18 mega-ohm
to a higher surface energy phase (less thermodynamically favorable water and HNO3 to yield dissolved samples in a 3% (wt/vol) HNO3 solution.
direction). An analogous argument was made in terms of the in- The HCl and HNO3 were trace-metal grade reagents.
fluence of hydration/dehydration on room temperature phase trans- Iodometric titration (42) was used to measure Mn AOS and speciation, by
formation (28) with thermodynamic competition between hydration the reaction of iodide with dissolved manganese oxide samples at room
(favoring the lower SE phase) and oxidation (pushing toward the temperature. Specific details of the titration of Mn(III and IV) oxidation state
formation of the higher SE but more oxidized phase because oxi- and calculations for complex Mn oxides are described elsewhere (6). Total
dation is strongly 10 exothermic). The second factor harkens back to oxidized equivalents, [Ox], of manganese may be expressed as the Mn AOS
observations made in a review by McKenzie (11) that significant acid obtained from the titration: [Ox] = 2[Mn(IV)] + [Mn(III)]. Subsequently, the
leaching of potassium from between the layers of birnessite precedes O/Mn ratio, as in MnOx, is calculated as: x = 1 + 1/2[(total oxidized equiva-
phase transformation of birnessite to cryptomelane This finding was lents)/(total moles of Mn)], where “total moles of Mn” for a manganese
incidentally observed later by Villalobos et al. (17) while mimicking oxide phase assemblage is taken to be “1.” The Murray titration is used to
birnessite syntheses in natural aqueous environments. obtain “total oxidized equivalents.” In terms of x, 2 SDs of the mean can be
In conclusion, the thermodynamic data obtained in this work offer very low (±0.002) with strict adherence to proper analytical methodology,
explanations for many of the trends of phase formation and trans- although limitations in analytical glassware and propagation of errors in-
formation among birnessite and cryptomelane phases in nature and creases the reported error to x ± 0.01.
in the laboratory. Specifically, this study shows strong competition The samples were heated to 973 K at 10 °C/min under a flow of argon
between surface energy and oxidation effects in the transformation (75 mL/min) in platinum crucibles using a Setaram Labsys Evo thermal analysis
of K-birnessite to cryptomelane. Thus, cryptomelane formation is [thermogravimetric analysis (TGA)/differential scanning calorimetry] instrument
favored by more oxidizing conditions. In contrast, compositional and associated Calisto software. TGA data were corrected for buoyancy by run-
ning an empty crucible. H2O content was determined from the weight difference
differences in birnessites are driven by concentration effects in
(obtained on a microbalance) before and after annealing the samples overnight
aqueous solution coupled with small differences in surface ener-
at 975 K. Part of the measured total weight loss was due to loss of oxygen as Mn4+
gies. Therefore, birnessites can be highly variable in composition,
in the original sample was reduced on heating, and part was due to dehydration.
reflecting changes in aqueous solution chemistry with little ther- The former was calculated from the difference in manganese oxidation state
modynamic constraint from the energetics of the solid phases. All between the original and annealed sample measured by the Murray titration (42).
of the phases studied have substantially lower surface energies This value was subtracted from the total weight loss to give the weight loss due to
than anhydrous manganese oxides, making the layer and tunnel water. The details of this method are described elsewhere (6).
phases thermodynamically favored at high surface areas. Thus, the The custom built Calvet twin calorimeter and techniques for measurement of
initial formation of fine grained manganese oxide phases with formation enthalpy by solution calorimetry in a molten salt solvent at high
layer and tunnel structures by relatively rapid precipitation from temperature have been described previously (24, 25). Before calorimetric mea-
aqueous solution near room temperature in soils, sediments, or surements, samples were equilibrated with the controlled temperature and hu-
under laboratory conditions is thermodynamically driven. midity conditions of the calorimetry suite. Sample pellets (∼5 mg) were dropped
into molten sodium molybdate (3Na2O–4MoO3) (20 g) at 975 K, with oxygen
Methods flushing through the calorimeter at 30 mL/min and also bubbling through the
The goal of synthesis was to make a series of birnessites and cryptomelane of solvent at 5 mL/min. Flushing and bubbling maintains oxidizing conditions,
constant cation composition but varying particle size for thermochemical removes evolving water vapor, and aids dissolution. The methodology has been
measurements. McKenzie (11) and Bricker (36) reviewed several methods for described in detail previously (18–25, 43, 44). To obtain the mean enthalpy of
producing various Mn oxides. The present work used a recent slight modi- drop solution (ΔHds), experiments were replicated for a total of eight per sample.
fication of these methods (32), but overall, the synthetic strategy was based The difference between each replicate experiment value is within the 2-σ error.
on reduction of aqueous salts of permanganate by chloride at room tem-
perature at low pH followed by warm (32 ± 5 °C), wet aging. At high enough ACKNOWLEDGMENTS. We thank the US Department of Energy, Office of
salt concentration, with other variables held reasonably constant, we were Basic Energy Sciences, for financial support of this research (Grant DE-FG02-
able to produce samples of almost identical counter-cation concentration, 97ER14749).
CHEMISTRY
layer manganese dioxide related phases. Chem Mater 10(2):474–479.
14(7):1744–1753.
13. Birkner N, Navrotsky A (2012) Thermodynamics of manganese oxides: effects of
34. Lopano CL, Heaney PJ, Post JE, Hanson J, Komarneni S (2007) Time-resolved structural
particle size and hydration on oxidation-reduction equilibria among hausmannite,
bixbyite, and pyrolusite. Am Mineral 97(8-9):1291–1298. analysis of K-and Ba-exchange reactions with synthetic Na-birnessite using synchro-
14. Golden DC, Dixon JB, Chen CC (1986) Ion exchange, thermal transformations, and tron X-ray diffraction. Am Mineral 92:380–387.
oxidizing properties of birnessite. Clays Clay Miner 34(5):511–520. 35. Post JE (1999) Manganese oxide minerals: Crystal structures and economic and envi-
15. Ma LZ, Hartmann T, Cheney M, Birkner NR (2008) Characterization of an inorganic ronmental significance. Proc Natl Acad Sci USA 96(7):3447–3454.
cryptomelane nanomaterial synthesized by a novel process using transmission elec- 36. Bricker O (1965) Some stability relations in the system Mn-O2-H2O at 25 ° and one
tron microscopy and X-ray diffraction. Microsc Microanal 14:1–7. atmosphere total pressure. Am Mineral 50:1296–1355.
16. Drits VA, Lanson B, Gaillot AC (2007) Birnessite polytype systematics and identifica- 37. Tissue BM (2013) The concept of activity. Basics of Analytical Chemistry and Chemical
tion by powder X-ray diffraction. Am Mineral 92(5-6):771–788. Equilibria (John Wiley and Sons, Hoboken, NJ), pp 199–211.
17. Villalobos M, Toner B, Bargar J, Sposito G (2003) Characterization of the manganese 38. Stumm W, Morgan JJ (1996) Aquatic Chemistry: Chemical Equilibria and Rates in
oxide produced by Pseudomonas putida strain MnB1. Geochim Cosmochim Acta Natural Waters (John Wiley and Sons, New York), 3rd Ed, pp 103–106.
67(14):2649–2662. 39. Kielland J (1937) Individual activity coefficients of ions in aqueous solutions. J Am
18. Navrotsky A (2014) Progress and new directions in calorimetry: A 2014 perspective. Chem Soc 59(9):1675–1678.
J Am Ceram Soc 97(11):3349–3359. 40. Davies CW (1962) Ion Association (Butterworths, London), pp 37–53.
19. Navrotsky A, Ma C, Lilova K, Birkner N (2010) Nanophase transition metal oxides show 41. Brunauer S, Emmett PH, Teller E (1938) Adsorption of gases in multimolecular layers.
large thermodynamically driven shifts in oxidation-reduction equilibria. Science J Am Chem Soc 60:309–319.
330(6001):199–201. 42. Murray JW, Balistrieri LS, Paul B (1984) The oxidation state of manganese in ma-
20. Mazeina L, Deore S, Navrotsky A (2006) Energetics of bulk and nano-akaganeite, rine sediments and ferromanganese nodules. Geochim Cosmochim Acta 48(6):
β-FeOOH: Enthalpy of formation, surface enthalpy, and enthalpy of water adsorption. 1237–1247.
Chem Mater 18(7):1830–1838. 43. Navrotsky A (1997) Thermochemistry of new, technologically important inorganic
21. Castro R, Ushakov S, Gengembre L, Gouvêa D, Navrotsky A (2006) Surface energy and materials. Mater Res Bull 22(5):35–41.
thermodynamic stability of γ-alumina: Effect of dopants and water. Chem Mater 44. Navrotsky A (1997) Progress and new directions in high temperature calorimetry re-
18(7):1867–1872. visited. Phys Chem Miner 24(3):222–241.
Birkner and Navrotsky PNAS | Published online January 27, 2017 | E1053