NOV 2 1 1362 — -5 e NANMIST PO EPTION DOCUMERT CoLLEGHIDN 3 445k 0023134 5 ORNL-3293 l UC-4 — Chemistry TID-4500 (17th ed., Rev.) e - THERMODYNAMIC PROPERTIES OF MOLTEN-SALT SOLUTIONS Milton Blander OAK RIDGE NATIONAL LABORATORY operated by UNION CARBIDE CORPORATION for the U.S. ATOMIC ENERGY COMMISSION CENTRAL RESEARCH LIBRARY DOCUMENT COLLECTION LIBRARY LOAN COPY DO NOT TRANSFER TO ANOTHER PERSON If you wish someone else to see this document, send in name with document and the library will arrange a loan. L} — LEGAL NOTICE This report was prepared os an occount of Government sponsored work, Meither the United States, nor the Commission, nor ony person acting on behalf of the Commission: A. Maokes any warranty or representation, exprassed or implied, with respect to the accuracy, completeness, or usefulness of the information contained in this repert, or that the use of any information, apparatus, methed, or process disclosed in this report may not infringe privately owned rights; or B. Assumes any liabilities with respeet to the use of, or for damoges resulting from the use of ony information, apparatus, method, or process disclosed in this report, As wused in the above, "‘person acting on behalf of the Commission" In'ctud.ll any employes or contractor of the Commission, or employee of such contracter, to the extent that such emplayee or contractor of the Cummil!ifln, or |mploy-lu of such contractor prepares, di:s-minohl, or provides access to, ony information pursuant to his employment or contract with the Commission, or his employment with such contractar. ORNL-3293 UC-4 — Chemistry TID-4500 (17th ed., Rev.) Contract No. W-7405-eng-26 REACTOR CHEMISTRY DIVISION THERMODYNAMIC PROPERTIES OF MOLTEN-SALT SOLUTIONS* Milton Blander DATE ISSUED 0CT 12 1962 *This paper is to be presented as a chapter in Selected Topics in Molten-Salt Chemistry, Interscience Publishers, New York. OAK RIDGE NATIONAL LABORATORY Qak Ridge, Tennessee - .. ~ e - i UNION CARBIDE CORPORATION OAK RIDGE NATION ’ 3 4456 0023134 5 M for the U.S. ATOMIC ENERGY COMMISSION v CONTENTS INTRODUCT ION eiitiiitieeettireeteeee et esteeetetestessesasesseeseassseessansasssessssssssaasssasasnssessesseensasessasenbesnbennseatsaeseassasnsatnsasesuen 1 ol GENETAL ettt v e st ettt ea et e e Rt e e s e et na e ene e e e n e e nenenee 1 1.2 The Limiting Laws coiieicieiiiccieescceesesesreesiaesssessste e sanesesee e srnenesrnaessressnessaresssasssnnesssnsssasesssnnsssssesessses 1 1.3 The Temkin [deal Solution et n s st s e nn e 3 |.4 Salts Containing lons of Different Charge.......coconveiiieiniiiiiiicc et 7 [.5 Standard States and Units of Concentration ...ceecieciicnirencee et 9 SOLUTIONS WITH COMMON ANIONS OR COMMON CATIONS ..ottt v 11 1.1 Cryoscopic Methods of [nve stigation ...t 11 [1.2 Electromotive Force Measurements ......ccoiiieiiioniiniiiiiiiiiniine sttt enesressiesenne o 12 [1.3 Strongly lonic Salts Containing Monovalent Cations and a Common Anion .......cccocoiiiiiiniiincnnne 16 I1.4 Mixtures Containing Polarizable Cations and a Common Anion.......ccccceeiinieriecriincceneenrcnenee s 29 11.5 Binary Mixtures Containing Polyvalent 1ons. ..t 33 1.6 Discussion of Binary Systems with @ Common Anion ... s 52 [1.7 OBher SYSTEMS .ocviceiiciiiceicireiresie e et esns s ae s eeseasaen e sr e et smeeoeceneesaeeseen e e e bnasaenss s sansss st s srnesassrananns 57 RECIPROCAL SYSTEMS oottt ettt csr s e sseetesses s e essees e sass s se e be s sas e e s ebe s b e se et s b e e sassas s bennasaessbentesansanes 61 LT G@NEIAL ettt ettt ettt et er s e e s tae e mes e e st es e £ et e st e s e e e b e eRe s aene e se e e e e e e R Aba R s e b et e e e s ere s 61 111.2 The Random Mixing Nearest-Neighbor Approximation ..........ccocoiviiiiiiiniiiiinninninnnn e 63 I11.3 Corrections for Nonrandom Mixing: The Symmetric Approximation .......cccoeererieenienreeicncieinnas 67 I11.4 Comparison of the Symmetric Approximation with the Random Mixing Approximation ................ 69 [11.5 The Asymmetric ApProximation ..o eirirniiieiiecieetisecteseesteeea s eressts e s et e e sbessseassassesbesaresenssnnnes 73 [11.6 Conventional Association ConstaNTs ...cciiriiiiiieciee e sas s s saees 75 [11.7 Comparison of Theory with Experiments in Dilute Solutions .o, 77 11,8 Generalized Quasi-Lattice Calculations ... 81 111.9 Association Constants in Dilute Solutions .. 84 MISCELLANEODUS cooeeieteeieeieitecere e e eeeeetese s aeseseessesuessaassaessasasasesssasaessssansessessssasenssesnsissstesnsssssesesssessssssessnsssesnsnsnes 89 REFERENCES ( aT > = r + s . c?nArxs o \\ / —v | N O ~ 0 Q.2 0.4 0.6 0.8 10 M, Fig. 2 Total Molar Heats of Mixing (AHm) in NuN03LiNO3 Mixtures (LiNO3 Is Component 1}. also the greater the difference in size of the two cations). In all the systems an energetic asymmetry in the enthalpies of mixing is present so that for a given pair of nitrates, the value of AH_ is more negative in a mixture dilute in the large-cation nitrate than in a mixture dilute in the small-cation nitrate. The parameter b is a measure of the energetic asymmetry. Assum- ing that the form of Eq.(2)is correct, then the partial molar enthalpies are given by Hy = HS=(a+2b = N2 + (4c = 26)N3 ~ 3eN$, (11.3.3) Hy = H}={a~b=cN}+(2b+ 40N} = 3cNY , (11.3.4) at N, =1, El - H? =a,and N, = 1, Ez - Hg = (a + b), where component 1 has a smaller cation than component 2. Since both the a and the 4 are negative, the partial molar enthalpy of solu- tion can be seen to be asymmetric. Only for systems in which the absolute value of AH s small does it appear that the parameter c is negligible and that the term containing the concen- trations to the fourth power are not necessary to represent the data. 18 UNCLASSIFIED 1500 ORNL-LR-DWG 66345 o =9 @ —4400 "/)’ @ - -, 2 P ~ © 7 // 8 -1300 & T /4"/ = 1200 > X ~ 3 T /;/ < 4100 /y}’ —1000 - 320 - —280 // —240 fi \ o -200 \ -160 / \ -120 Hy, (cal/mole) ™ .\ " Ll L -80 —-40 / ~\ 0 0.2 0.4 0.6 0.8 1.0 N, Fig. 3. Total Molar Heats of Mixing (AHm) in CsNO,.NaNO, Mixtures (NuNO3 Is Component 1). Kleppa, by using the enthalpy of mixing of 50-50 mixtures of the nitrates as a measure of the magnitude of the effect, demonstrated the empirical relation di"‘dz 2 4AH° S Ly ( > = Us? = ~14057 , (11.3.5) dl+d2 where & = (d.' ~ d,)/{dy + d,), and d; is the sum of the radii of the cation and anion indicated, and U is about =140 kcal, The value of U is about the same magnitude as the lattice energy of 19 UNCLASSIFIED "o ORNL-LR—DWG 66346 % —2000 L-LR—DWG £ ~ L o 8 —1920 — ~— L — QN '/. \GP\\ = -1840 o * k> ., ___-_-—--__- X oo’ T 5 1760 Ko e =t~ —-480 —~, 7 < | W n O ~N < —160 / . / \ —80 ° s 0 0 0.2 0.4 0.6 0.8 1.0 N1 Fig. 4. Total Molar Heats of Mixing (AHm) in KN03-LiN03 Mixi'ures(LiNO3 Is Component 1). the alkali nitrates. The results of Kleppa may be rationalized in terms of simple concepts. Since the simplest binary mixtures are those containing monovalent cations and anions, simple solution theories are more likely to apply to these mixtures than to mixtures containing polyvalent ions. Although some of the relations discussed below will be naive, they will serve the main ob- jective of this discussion, which is to relate the solution behavior of molten salts to fundamental physical laws. As discussed in a previous section, a molten salt may be compared to a quasi-lattice. Be- cause of the alternation of charge, the quasi-lattice consists of two sublattices, one of cations, and the other of anions which interlock so that the anions have cations as nearest neighbors and the cations have anions as nearest neighbors. For a mixture of salts with a common anion, the cation sublattice may be considered as being imbedded in a sea of anions. The anions are not excluded from consideration, since the cation environment of a given anion will greatly affect its relative position and energy. Since the solute and solvent in a mixture both have the same anions as nearest neighbors as they do in the pure state, any solution effects are caused by ions further away although these ions further away may, indirectly, affect the nearest-neighbor anions. 20 Table 2. A Summary of the Parameters a, b, and ¢ Derived from the Heat of Mixing Data for Binary Nitrate Systems System T (°C) a (cal/mole) b (cal/mole) ¢ (cal/mole) (Li-Na)NO, 345 - 464 -11.5 ~0 (Li-K)NO, 345 - 1759 - 87 - 463 (Li-Rb)NO, 345 — 2471 - 178 - 945 (Li-Cs)NO, 450 (- 3000) (Na-K)NO, 345-450 ~408.5 -68 ~0 (Na-Rb)NO, 345 ~744.5 - 268 -36 (Na-Cs)NO, 450 -~ 1041 — 435 - 93 (K-RBINO, 345 (- 60) (K-CsINO, 450 ~89.c ~87.5 ~0 (Rb-CsINO, 450 (- 14) (Li-Ag)NO, 350 702 - 108 0 (Na-Ag)NO, 350 677 ~ 156 0 (K-Ag)NO, 350 -303 - 294 0 (Rb-AgINO, 350 ~ 944 - 337 ~297 (Li-TINO, 350 - 901 178 - 294 (Na-TINO, 350 131 241 ~0 (K-TINO, 350 447 - 17 ~Q (Rb-TINO, 350 240 ~15 ~Q %Pparentheses indicate uncertain data, Molten-salts solutions differ from solid-salt solutions in an important respect, |n order to place a large cation in solution in a solid salt having a small cation the structure near the for- eign cation must be distorted. |n a solid, such a distortion is difficult as evidenced by the rigidity of the lattice. Although there is some ability of the ions in a solid to adjust their positions to minimize the energy,*4 the net effect is that the enthalpy of mixing of ionic solids is positive, and there is a strong tendency for icnic solids having a common anion to be mutu- ally insoluble if the cations are very different in size. The structures of molten salts are much less rigid, and the salt can easily accommodate cations of different size. The theory which can most easily be applied to mixtures of molten salts with monovalent ions is the quasi-chemical theory of Guggenheim®? which is based on a quasi-lattice model. Since it may safely be assumed that cations almost exclusively have anions as nearest neighbors in a solu- tion containing only one kind of anion, all the nearest neighbors of the cations will be the same as in the pure salts, and solution effects will be caused by ions further away than nearest nieghbors. The nearest cation neighbors which are next nearest neighbors in the salt quasi-lattice might be considered as a first approximation. 2] If salt 1is AX and salt 2 is BX, then the potential energy of the ion triplet A* X~ A* may be designated by V,,, of B*X~B" by V,, and of A*X-B* by Vu.78 To validly apply the quasi- chemical theory to the model, V,,, V,,, and V,, must be assumed independent of the local environment of the ionic triplets. Although this assumption is not correct, it may serve as an initial working hypothesis. The molar excess free energy of solution and molar heat of mixing of solution as calculated from the quasi-chemical theory will be given by®8 -——AAE NN, AT~ NN —A (11.3.6) = - + .00 |, e AH_ 22 T-N1N2A1—N1N2m + ... ‘ (”.3.7) where A= (L Z°/2) (2V12 -V =Vy= N ZAe /2,1l is Avogadro’s number, and Z” is the number of cation nextenearest neighbors of a cation. Fériand37+38 has discussed the quantity (2V12 -V - V22) = A€’ in terms of the change of the repulsions of next-nearest-neighbor cations. Fdrland represents the configuration of next- nearest-neighbor cations and a nearest-neighbor anion as in Fig. 5 and calculates the Coulombic energy change, Ae_, for mixing the cations in these two arrays of three hard spherical ions. 11N /dy =dy\? Ae_= —e? <_ + —_> (_‘_—i> , (11.3.8) dy d,)\d,+ 4, where e is the electronic charge. The term —ez[(l/dT) + (1/d,)1, for a real ionic salt, can be re- lated to the average lattice energy of the two salts composing the mixture and is analogous to the UNCLASSIFIED ORNL-LR-DWG 40446R oo ‘ J Fig. 5. Configuration of lons for the Calculation of Fgrland on the Change of Repulsions of Next-Nearest Neighbors. 22 empirical parameter U in Eq. (5). A€_ is always negative and tends to be more negative the greater the differences in the cationic radii. Except for small factors the form of Eq. (8) is ob- viously related to the empirical relation (5). Blander !5 has extended Férland's calculations to a hypothetical salt mixture which is ex- tremely dilute in one component and which is represented by an infinite linear array of hard- charged spheres. Although this model is unrealistic for a real three-dimensional salt, it does serve to assess the effect of Coulombic interactions of longer range than the next-nearest neigh- bors. The inset of Fig. 6 is a picture of a portion of the solution of one mole of the solute with an interionic distance 4, in an infinite amount of solute. In Fig. 6 are plotted calcuiated values of —AEC a?]/e2 vs A, where d, = d,(1 + A), and where A€_ is the energy of mixing per molecule of solute. The values of A€_ are always negative and become more negative the greater the value of A and are only about 0.4 times the magnitude of the values calculated from Férland’s simplie model, |f the mutual dissolution of two salts 1 and 2 where salt 1 has the smaller cation is con- sidered, then Blander's calculation indicates that a dilute solution of 2 in 1 will lead to a more negative value of Ae_ than for a dilute solution of 1 in 2, Since Férland’s calculation predicts a symmetry in the energy of mixing, the effect of the long-range interactions is to decrease the total calculated value of | A€_| and to lead to a small asymmetry in the energy of solution. The asymmetry effect means that the parameter A cannot be independent of composition. UNCLASSIFIED ORNL-LR-DWG 52098R 0.07 0.06 ——\ d d(lh’l)r— 008 OOOEOEE )OO — - dum)l-- 0.04 \ \ 0.03 / 0.02 / e \ o \/ -06 -04 02 0 0.2 0.4 06 08 1.0 A -AE e2/d Fig. 6. Values of the Energy of Solution at Infinite Dilution (AE) in Units of(ez/d) Calculated from an Infinite One-Dimensional Solidlike Model. 23 Care must be taken not to ascribe the observed negative deviations from ideal solution be- havior or an asymmetry to only Coulombic effects. For example, polarization may also con- tribute to the energy of solution. The ions in the solid-like linear model for the pure salt have no field on them, but in the mixture represented in the inset of Fig, 6 there is an appreciable field intensity on some of the ions which can polarize the ions. Lumsden’! has calculated the effect of polarization of the anions by the cations in terms of a one-dimensional model essentially including only nearest-neighbor and next-nearest-neighbor ions. He obtained the relation 2 4 2 aF T 1 d, - d Ae, = — x—ae? | — + — | (L—2) , (1.3.9) P2 d, d, d,d, where F is the field intensity on an anion between two cations of different size, and a is the po- larizability of the anion. Polarization of cations, which may not be small, has been neglected. Equation (9) is the same form as (5) and (8), and AEP is negative so that it should be difficult to separate the purely Coulomb interactions from polarization interactions without a valid calcu- lation of the relative magnitude of these two interactions. However, any simple extension such as 91 to three dimensions of a one-dimensional model for either the Coulomb was made by Lumsden or polarization interactions may lead to misleading values for their relative magnitudes. If the solute in Fig. 6 is salt 2 in the solvent 1 where cation 2 is larger than 1, then the field intensity on the anions adjacent to the solute cation is greater than if the solute is salt 1 and the solvent salt 2, For polarizable anions, this would make the energy of mixing more negative and contribute to the asymmetry effect. [f thermal motions are considered, then the tendency of ions to reside longer in regions of high field intensity will also contribute to the asymmetry being in a sense a ' 'positional’’ polarization. |f these simple considerations are valid for a real three- dimensional salt, then at least part of the asymmetry effect is related not only to Coulombic but also to polarization interactions by ions more distant than next-nearest neighbors. In any theory of molten-salt mixtures it appears to be necessary, then, to include long-range interac- tions, except under very special conditions. The comparison of the measurements with the concepts discussed is straightforward. As discussed, the parameter A for a molten salt in (7) is not independent of composition and Kleppa’® has approximated the effective value of A as a linear function of composition, A=a’+ b'N,, (11.3.10) so that for values of =\ small relative to Z'RT, Eq. (7) becomes AH =N Ny(a’+b°N)), (1.3.11) which is the form of the experimental results in the three systems studied by Kleppa which ex- hibit the smallest deviations from ideal solution behavior. 24 For values of =X not too small, Eqs. (7) and (10) lead to ~ , (a”+b'N,)? AH = N N,|la"+b'Ny = 2N Ny —— (11.3.12) Z'RT Comparison with Eq. (11.3.2) shows that ¢ can be identified with —2\2/Z'RT. The precision of the measurements is not high enough to detect a change of ¢ with composition. By using an average value of A, )\=a+b/2, as a measure of A, Kleppa showed that a plot of ¢ vs 7\2/RT for the systems (Nq-Rb)Nos, (Na- Cs)NOs, (Li-K)NO,, and {Li-Rb)NO, is consistent with a reasonable range of values of Z* of 10 to 12, This is the number of next-nearest neighbors in an NaCl type lattice and is only a small variance with the number of next-nearest neighbors in some molten alkali halides.®” Equation (11.3.11) corresponds to the random mixing of the cations on the cation sublattice. The presence of the c term, if A varies linearly with composition, implied an appreciable non- random mixing of the cations, and ¢ was termed a short-range order parameter by Kleppa.”® It should be made clear that although the results of Kleppa have been rationalized in terms of the modified quasi-chemical theory, a fundamental premise of the quasi-chemical theory is that A is independent of composition. Consequently, the form of the theoretical equations de- rived, based upon the quasi-chemical theory, although in correspondence with the empirical Eq. (2), requires a sounder theoretical justification. A justification of the form of the empirical Eq. (2) has been made by the methods in the elegant work of Reiss, Katz, and Kleppa.'!® They used a method, which is essentially an adaptation of the theory of conformal solutions,®® in which no model is used. The derivation was made for ions behaving as hard-charged spheres with a sum of radii equal to 4 so that the pair potential ulr)=e0, rd, (11.3.13b) Kr where r is the distance between the two ions in any given pair, and « is a dielectric constant which is assumed constant for a set of salts with a common anion. The potential function can be generalized to the form for a monovalent salt 1 u(r) = i—;f(r/d) , r>d. (11.3.14) This is a less-stringent condition than (13b). Because of the relative rarity of anion-anion con- tacts (except in salts as Lil), or cation-cation contacts, the contribution to the configurational integral of configurations in which ions of the same charge are touching (or almost touching) is very small and is neglected. As a consequence, except in these rare configurations, the total 25 contribution to the potential energy of a given configuration due to cation-cation or anion-anion interactions is independent of small differences in the cation size, and only one parameter of length d for the sum of the radii of a cation-anion pair is necessary for the description of a pure salt. For a mixture of salts, two parameters are necessary, dy and 4,. In the derivation of the theory, a single-component reference salt with the single parameter of length d is transformed into either component 1 or 2 by varying 4. f g; = d/dz. where i = 1 or 2, then the configuration integral for the pure salt i is -BUi Z;=Z(g,) =f--f“g(;|)"3-(d7)2’7, (11.3.15) where U, is the potential energy of the 2% ions (7 cations + 72 anions). Since the cation-anion pair potential is u {r) =if(—r)= g;ulg;7), (11.3.16) then the total potential ~ I >t 8] & EL giuacle;) + 2 2, ucc (1.3.17) where A represents anions, C cations, and the symbols A < A”and C < C” signify that the pair potentials are added in @ manner so that no pair is counted more than once. The molar Helm- holtz free energy,* A_ for pure salt i can be expressed as a series A, dln Z TRT InZ;=InZ(g)=(In 2) _; +{g; = 1) og; g;= 1 (g, =12 /0%2InZ Lle=1D " Y uee, (11.3.18) 2 agf Ei=1 *Only the configurational part of the partition function is treated here. In calculations concerning changes upon mixing, the ‘‘translational’’ part drops out and may be neglected. Although the equations that follow were derived for hard sphere ions which interact with a genera?ized form of the Coulomb potential, the same equations may be derived for more general potential functions. If the core repulsions of a cation-anion pair are of the form /(gir) {(a special case of this form is the hard-sphere repulsions) and if the other interionic in- teractions in the system are such that for any given geometric configuration of the ions, the permutations of the two types of cations over all the cation positions do not lead to a change of the contribution of these other interactions to the total potential of the system, then the equations derived will be the same as the equations to be derived [Egs. (23—27;1 with different values of qSAC in the integrals which are contained in the coefficients. Types of interaction which would be included in this category are not only Coulomb inter- actions but also cation charge-anion multipole interactions and,for the cases in which the two cations have relatively small or equal polarizabilities, all other charge-multipole and multipole-multipole interactions. Salt mixtures which conform to this might be termed conformal ionic mixtures. 26 where Z(g.) has been expanded about g, = 1. Similarly, for a mixture of 1 and 2 the potential energy is given by U,=% = T (630 +3 3, up,’ = u ) + u r) + u 12 = & & E1M“AacC\8 X < 824Ac 82 & & tan (N7} (N %) (N,7) (N7 (N 7) (N ,7) + Z 2z u r+ X 2z u v X T u , (11.3.19) ¢y < & ©1% & < ¢y %% & & & and Z is given by -BU , 7! e 12 - 7l Z.|2 = P —-——-——(d’T) n . 2(81’82) ’ (”.3.20) Al (71)2 7y ln)! and the Helmholtz free energy for one mole is A dlnZ dlnZ -—12 - (In 2) + ( —1)< ) + ( —U( ) - 81:8 =1 gI 82 kT 1482 dg 61.8y=1 g, —— ,lea - 1? Y., (11.3.23) ™ dldz dy—d, 2 E AG” = NN, 0(T,p) <——-—> + ., (11.3.24) " dld2 and dy —d, 2 AH =N N, Q(T,p) (———) +..., (11.3.25) 7 d1d2 where [, 8, and Q are functions characteristic of a single *"test’’ salt. The influence of the factor (a’l 152’2)2 on the thermodynamic excess functions is much weaker than the influence of the factor (d, ~ d2)2. As a consequence, the form of Eqs. (23), (24), and (25) is similar to that implied by Eqs. (8) and (9) and is consistent with the empirical relation (5). The higher order terms in the theory of Reiss, Katz, and Kleppa were complicated. The higher order terms are simplified by the choice of particular relative values of the perturbation param- eters g, and g, so that (g, — 1) = (g, ~ 1).'7 This condition implies that for each particular mixture a ‘‘test’’ salt is chosen. The calculations lead to the result AAE SN N, POT 4 N NNy = N)O8% + [N NR + N NN, — N1+ L, (11.3.26) where P 1(32 zn) kT 2 \z?2 1z 0 (_71'2F 72BD B® ) —=4 +— |, kT 6z 2z% 3z3 R and § are complicated functions, dl_'d2 0= |————1, dl+d2 B=-B7’a, D =B%% e+ (7 - Nl , 28 B37 F=- 2 f fquc bac*Pacre” (d7)27 (71)? +3(= -1 fquAC ¢AC'¢A’c”_.3-BU (d,r)i’r'z- + (7= N)(7 - 2)f...fquc bprcrbprrcrreBY (an?T P, O, R, and S are characteristic properties of the test salt. In a similar fashion, the heat of mixing may be shown to be ’ ’ 1ol AH =N N PB4+ N NNy = NJQ B3+ [N NR 4 N NNy = N)2SI84 + ., (11.3.27) where the prime signifies the proper temperature derivative of the primed quantity. The form of Eq. (27) is seen to be consistent with (2) if a=P 82 -08% + RS* + 5%, b =203, and ¢ =-45’8%. This constitutes proof that the form of Eq. (2) is consistent with rigorous theory. The methods of Reiss, Katz, and Kleppa may thus be used to support in a rigorous manner the form of the empirical Eqs. (2) and (5), as well as the approximate form of Eqs. (8) and (?) which had been derived on an intuitive basis. Powers, Katz, and Kleppa’41%6 have measured volume changes of mixing of several com- positions of each of the binary alkali-nitrate mixtures (No»K)NO3, (No-Li)NO3, (Nc-Rb)NOa, and (Na-Cs)NO,. The average values of the quantity AVE/N]N2 are listed below: Mixture Temperature (°C) AvE/N 1N, (cm3/mo|e) (Na-K)NO3 350 0.26 +£0.08 425 0.28 +0.08 (Na—Li)NO3 310 0.26 £0.02 (Na-RB)NO, 340 0.82 £0.10 (NCI—CS)NO3 425 1.37 £0.12 All of these volume changes are positive and obey the approximate equation AVE =N N, VB4, where V”'= 22,000 cm?/mole. These positive deviations from the additivity of the molar vol- umes, significantly, are found in mixtures in which the heats of mixing are negative. No satis- factory theory has been proposed for this. The only data on activities in mixtures of alkali halides with a common anion has been ob- tained from cryoscopy. Unfortunately such data is not isothermal and uncertainties in the phase diagram and in the heats of fusion as well as the necessity for precise measurement of liquidus temperatures to obtain reasonable values of the excess free energies reduce the value of this source of information. The component LiF in mixtures of LiF-KF, LiF-RbF, and LiF-CsF3.3¢ exhibit negative deviations from ideal solution behavior, which are more negative (the activity coefficients are smaller) the larger the difference between the sizes of the two cations. The 29 same is true for the component LiCl in mixtures with KCI, RbC!, and CsCIl.''? This appears to be in accord with the ideas presented in section ({1.3). However, the work of Cantor?! on cryoscopy of NaF in mixtures with KF, RbF, and CsF have indicated that there is a small positive excess free energy which becomes more positive in the order KF < RbF < CsF. Since purely Coulomb or polarization interactions would be expected to always lead to negative de- viations from ideal solution behavior, it is clear then that even in mixtures of the highly ionic alkali halides other types of interactions are important. In the next section we will show that these interactions may be, at least in part, related to the dispersion interactions of the solute cations. Some discussion of this for alkali halides has been made.?! 1.4 Mixtures Containing Polarizable Cations and a Common Anion In order to separate the various physical interactions which are significant in determining the solution behavior of molten salts, it is advantageous to compare two different mixtures of salts in which the major difference in the solution properties can be related to the differences in the prop- erties of one ion. As an example, mixtures of alkali nitrates with silver or thallous nitrates would be suitable for such a comparison with mixtures containing only alkali nitrates, since the differ- ence in the properties of Ag* and T1* jons from those of Na* and Rb* is largely related to the rel- atively high polarizabilities of Ag* and T1*,*+105 Kleppa has measured the heats of mixing of AgNO, and TINO, with all of the alkali nitrates except CsNO3.77'79 By fitting his data to equation (11.3.2), where N, is the mole fraction of either AgNO, or TINO,, Kleppa obtained the values of the parameters a, b, and c which are listed in Table 2. The observed deviations from ideal solution behavior differ from those of the corre- sponding mixtures of alkali nitrates with NaNO, or RbNO,. In addition to the interactions present in mixtures of alkali nitrates, an additional interaction needs to be postulated to rationalize the ob- served results. This difference has been shown to be in reasonable agreement with a calculation of the London dispersion energy of interaction of next-nearest neighbors. '® The predominant term of the London dispersion interaction energy between two ions is the dipole-dipole term, k! rr —kl 6 Uty ==, CL/d°, (11.4.7) where 5:”is a constant probably in the range of 1 to 2 and depends on the structure of the melt, d for a pure salt is the cation-anion distance with the cation-cation distance assumed propor- tional to 4, and 4 for a mixture is an average cation-anion distance. The paramater Cii is given by101 3a ol k1 kRCL chlae ————, (11.4.2) 2 I +1, *1t should be noted that although the Pauling radius of Ag+ ion is 1.26 A, the interionic distances in AgCl and AgBr and the relative molar volumes of liquid AgNO, and NaNO, are more consistent with a radius of about 0.95 A which is close to that of Na+. 30 where & and [ are the two cations, ais the polarizability of an ion, and I has been estimated !0 for the alkali cations, TI* and Ag*. Values of c and I are listed in Table 3. A calculation of the contributions of this interaction, AUfP, was approximated from the dispersion energy change represented crudely by the process AXA + BXB - 2AXB, A AULRE = 2048 —uhh - BB, (1.4.3) where the solutes are AX and BX, where Sg’ 2 1.8, and where 2d, o =d, s +dgg.* Equation (11.4.3) is an approximation to the contribution to AH?H'S/NTN2 so that the relation for molten nitrates (11.3.5) is modified to become XZAARDS = Us? + AUAB (11.4.4) The value of U = =140 for alkali nitrates includes a small positive contribution from van der Waals' interactions so that a correction is needed which will make AUff less positive.'® A cruder but simpler approximation to AULB may be made in a manner similar to an approximation useful in 68 2 e x (forr - fuBE), (11.4.5) where the values of C__ in Table 3 in conjunction with a value of S;’: 1.8 may be used with Eq. nonelectrolyte solution theory. (5) and ionic radii for roughly estimating AU4AB. From Table 3 it can be seen that C,, will be quite large for Cs*, Rb* and K* ions and the positive term, AUfP, in (4) may be large enough to cancel the negative values of U8? for mixtures of, for example, NaF with KF, RbF or CsF. The calculations of Lumsden?®’ are in accord with this and this may be used to rationalize the * A better approximation for dpg is [(a’i + d; )/2] ]/2, which differs little from (dA + a’B)/2 when d, is not very different from dg. The factor for Sg' contains a small correction for interactions of longer range than next-nearest neighhbors, Table 3. Polarizability and Potential Parameter Used for Estimating Cation-Cation van der Waals’ Interaction lon a x 1324 Ix 1012 (em™) (ergs/molecule) Lit 0.030 90.9 Na® 0.182 56.8 k* 0.844 38.2 Rb* 1.42 33.0 cst 2,45 39.0 Agt 1.72 30.0 + Tl 3.50 30.0 31 ! mentioned at the end of the previous section. However all of these methods results of Cantor® are approximate and are useful largely for semiquantitative estimates of solution behavior. Laity®3 has shown that A 4.¢¢ is negligible in the cell AgNO,4(N ) AgNO (N ) g Ag NaNO, NaNO, and the emf of this cell is given by AE =RT In {a /aj) . The measurements are consistent with the expression uf =840 N2, (11.4.6) where 1is AgNO, and 2 is NaNO,. The results did not exhibit the asymmetry in the heats of mixing found by Kleppa for the same system. Although the total excess entropy is small rela- tive to the total entropy of mixing, it is negative and is not small relative to AH_ or AGE; TASE = AH - AG" = (=156 N; = 163)N N, , m so that although Eq. (6) has the form for regular solutions the excess entropy does not appear to be negligibly small. There have been many studies of mixtures of silver halides and alkali halides using the formation cell AgX WX X, (p = 1 atm) graphite, y where M is an alkali metal ion and X is a halide. The emf of this cell can be related to the activity of AgX by p]—-,ucl)z—F(E—-Eo)=RT|n ay, (11.4.7) where 1 is AgX. The most extensive work on these systems has been the work of Hildebrand and Salstrom who studied mixtures of AgBr with LiBr, NaBr, KBr, and RbBr. 86,116,117 | Fig. 7 are plotted values of #I]E for AgBr (component 1) vs Ng. Within the experimental precision, ,ufl‘: is independent of temperature and can be represented by the equation 2 ,ufl? = AN . (11.4.8) Values of A are given in the table below and may be rationalized in terms of Eq. (4) using the data in Table 3. Many studies of mixtures of AgCl with alkali chlorides have been made. Unfor- tunately, there are significant differences between different measurements on the same systems. The most reliable and consistent studies appear to be those of Salstrom '8 and of Panish %4 on the (Li-Ag)C! and the {(Na-Ag)Cl systems. Although there is scatter in the high-temperature data 32 Volume Change Mixture A (cal/mole) of 50-50 Mixture (cm3/mo|e) {Li-Ag)Br 1880 -0.13 (Na-Ag)Br 1050 +0.17 (K-Ag)Br ~ 1480 +0.27 (Rb-Ag)Br - 2580 +0.42 (Li-Ag)ClI 2100 {Na-Ag)ClI ~ 800 UNCLASSIFIED ORNL-LR-DWG 66348 1600 /7 1200 /e a @ _ L~ LU-— /A/'a i o ® 0 fi ' o \. \\Kar -400 b~ - o ® N \.. \ . RN -800 RbBr N\ -~ o \ -41200 0.2 0.4 0.6 0.8 1.0 2 N2 Fig. 7. Values of the Excess Chemical Potential of AgBr (Component 1) in Mixtures with Alkali Bromides. 33 of Panish, the results of both Salstrom and of Panish on the AgCl-LiCl system lead to positive deviations from ideality, which follow, approximately, Eq. (8) with A = 2100 (cal/mole) from 500 d104 to 900°C. Small positive deviations from ideality have been foun in the AgCl-NaCl system. There is too much scatter in the results to be able to represent ,u']‘: precisely, but crudely ,u’]‘? = 800 Ng. The work of Stern is consistent with these results. 24 In all of these chloride systems there is considerable scatter and uncertainty, and it cannot be clearly shown that the data can be best expressed by an expression as Eq. (8) and that X is truly independent of temperature. Measurements of AgCI-KCI mixtures by Stern are doubtful. 2% The measurements of Mur- gulescu and Sternberg indicate that for AgCl-KCI mixtures 102 ui = =1555 N2, and that the excess entropy of mixing was nearly zero. However, the values of E observed by Murgulescu and Sternberg differed from those given by Salstrom, Panish, and Stern by about 9 mv at 500°C. An interesting comparison with solid solutions is exhibited by Panish.'%4 Although the molten salt system AgCl-NaCl exhibited only small positive deviations from ideality, the measured deviations from ideality in the solid solutions were more positive. This illustrates the fact that, aside from other effects, the accommodation of ions of different sizes in a given material leads to a greater positive {or less negative) free-energy change in a crystal than in a liquid. 124 4hat the volume change upon mixing of It was pointed out by Hildebrand and Salstrom 50-50 mixtures of the four systems containing AgBr, which are listed on page 32, could not be related to a weakening or strengthening of the interactions of the ions or with the devia- tions from ideality. As with the results on alkali nitrates for AH_, the values of A;f:; vary in a direction opposite in sign to that of AVE with variations in the cation. I1.5 Binary Mixtures Containing Polyvalent lons Although there has been much experimental work on mixtures containing polyvalent ions, very little theoretical discussion based on fundamental physical principles has been published. This section will be devoted to the presentation of thermodynamic data to give the reader an idea of the magnitudes involved and, where enough data exist, to pointing out the correlation of proper- ties of mixtures with the physical properties of the ions. Where it is considered necessary, a discussion of the principles of measurements will be included. In the next section, a discussion of these data and a critique of the description of these data in terms of '‘complexes’ will be made. Kleppa and Hersh?7 measured the molar heats of mixing of Ca(NO,), with LiNO,, NaNO,, KNO,, and RbNO, at 350°C. By using a heat of fusion of Ca(NO,), of 5.7 kcal/mole obtained by 34 * 0 -Hl extrapolation of their measurements, the limiting heat of solution of liquid Ca(NO,), (E] in Table 4) at 350°C obeyed the empirical relation HY = H =0.3 = 225((r,,/2) = (r,/ N 2/1d, + 4,12, which relates the radii of the divalent and monovalent cations (r++ and r+) to the observed heats of mixing. The heats of solution decrease with increasing radius of the alkali cation. It should be noted that the heat of solution in LiNO, is positive. No simple representation of the concentration dependence of the molar heat of mixing was made. It was noted, however, that the slope of plots of AHm/N][Ca(Noa)z is component 1] vs N, for mixtures with KNO, and RbNO3 had maxima at Ny =0.25-0.33 (or at equivalent fractions N{ = 0.4-0.5). The results in these two systems prob- ably can not be represented by an equation with as few terms as (11.3.2). Table 4. Extrapolated Values of the Limiting Heats of Solution of Ca(N03)2 — 0 Solvent HI ~ HY {kcal/mole) LiNO3 +0.25 N(:lNO3 -0.9 KN03 ~3.0 The most extensive comparative studies of binary mixtures containing polyvalent ions have 30,3134 who measured the freezing point lowering been the cryoscopic measurements of Cantor, of NaF by polyvalent salts. NaF can be considered as a prototype of an ionic salt. In Fig. 8 are plotted the liquidus temperatures of NaF {component 1) in mixtures with the alkaline earth fluorides. The upper line is the calculated liquidus temperature for an ideal solution with the data contained in Table 1. For an ideal solution at the liquidus _ nideal ay =N, , and the activity coefficient in a real solution is given by ideal N ¥y =— Ny at the liquidus, where Ni]d“' and N, are the compositions of NaF in the ideal and real solu- tions respectively at the same temperature. A freezing point fower than the ideal value means that y, <1 so that the solutions all exhibit negative deviations from ideality. The smaller the radius of the alkaline earth the greater the deviations from ideality. The Ca?* ion has about the same radius as Na*, but the NaF-CaF, mixture exhibits negative deviations from ideality. This illustrates the effect of charge. Deviations from ideality in the 35 UNCLASSIFIED ORNL-LR-DWG 67567 1000 980 960 940 920 900 TEMPERATURE (°C) 880 860 \ CaFp 3 Mng 820 O 5 10 15 20 25 30 SOLUTE (mole %) Fig. 8. Liquidus Temperatures of NaF in Mixtures with Alkaline Earth Fluorides. NaF-BaF, system are small. Since the Ba2* ion is larger than the Na* ion, the large size of the divalent ion appears to, at least partially, compensate for the greater charge. The excess free energies of NaF, #1]5’ at the liquidus temperatures in mixtures with the alkaline earth flu- orides are plotted in Fig. 9 vs Ng (where 2 is the solute). (Note that these values of pfi? are not isothermal.) For comparison with monovalent cation salts, data with LiF and KF as solutes are also plotted. The Li* ion is about the same size as the Mg2* ion and both are smaller than the Na* ion. If the LiF and the MgF, mixtures are compared with NaF, the deviations from ideality in both appear to be negative, being much more negative in NaF-MgF, mixtures. On 36 UNCLASSIFIED ORNL-LR-DWG 67571 50 KF 0 g —— -50 \\ \.\ i, \>\ CQFé ///// -350 -400 \ MgF, -450 \ ) BEFz -500 O 0.02 0.04 0.06 0.08 2 (1 _NNaF) Fig. 9. Excess Chemical Potential of NaF in Mixtures with the Alkaline Earth Fluorides LiF and KF as Calculated from Liquidus Temperatures. the other hand, the K* ion is about the same size as the Ba2* ion, both being larger than the Na* ion. The deviations from ideality of the solvent NaF in mixtures with KF and BaF, are both small. A further illustration of the influence of charge is shown in Fig. 10, which gives ,ulls for NaF in mixtures with Can, YF3, and Thf:4 in which salts the interionic distances are about the same. Al| of these illustrations show that the deviations from ideal solution behavior are related by a function which appears to be monotonic in the charge of the solute cation, Z, and in 'I/dz, where d, is the sum of the cation and anion radii of the solute. However, other effects such as van der Waals' interactions, ligand-field effects on transition metal ions, etc., will be superimposed on the effects of charge and radius of the ions. Figure 11 gives a parallel plot of yll‘: at 20 mole % of solute and the lattice energies of the solid solutes MnF,, FeF,, CoF,, NiF,, and ZnF,. The measured cation-anion distance in solid MgF, is about the same or smaller UNCLASSIFIED ORNL-LR-DWG 67569 0 m — ] -100 N T -200 \ N N N\ ¥ C0F2 -300 \\\\ 3 \ 2 -400 O Y Fy —————- w o Iy £ 3 -500 \ ~600 \ —700 ) % ThF, -800 0 0.010 0.020 0.030 2 (1- NNGF) Fig. 10. Excess Chemical Potentials of NaF in Mixtures with CaF,, YF3, and ThF, as Calculated from Liquidus Temperatures, than those of these transition metal ions. The greater negative deviations from ideality found for mixtures with the transition metal fluorides are therefore not related solely to the radii of the ions. Since the pattern of the lattice energies with a maximum at NiF , or CoF, is explained by ligand- field theory for octahedral or to tetrahedral symmetry respectively, 103 then the pattern of ,ufi; and the differences from the NaF-MgF, system suggest that the change of the ligand-field effect upon dissolution is related to the deviations from ideality of NaF. Having monovalent ions as next- nearest neighbors in the mixture, as compared to divalent ions as next-nearest neighbors in the pure transition metal fluorides, probably leads to a greater ligand field and a great ligand-field stabilization of the solute component in the mixture than in the pure salt. Whatever the specific structure of the melt and of the ligands about the transition metal ion, it is apparent that the effect of the ligand-field stabilization on the solvent is in the same order as might be expected from ligand-field theory for the solute. 38 UNCLASSIFIED ORNL-LR-DWG 67576 740 - T20 \ o o 0 5 -1 a / P léJ 700 - w / // w // Q —~ C 2 < - - 680 _T P -~ - ~ - -~ - -~ 660 &< 500 —~ p— o L \\' w450 -7 > - - -~ C -~ n - B - - w - Y 400 7 > Q P = -~ - Q _- - ~ ~ < 350 = o & W =2 // & -7 L 300 Man Fer CoF2 NiF2 Cqu ZnF2 SOLUTE Fig. 11. Parallel Plot of the Lattice Energies of Some Transition Metal Fluorides and the Excess Chemical Potential of NaF in Mixtures with Transi- tion Metal Fluorides. Cantor34 has also made cryoscopic measurements on NaF with ZrF4, HfF4, ThF4, and UF, as solutes. In Fig. 12 are plotted values of yfi? Vs Ng at the liquidus for these four mixtures. The deviations from ideality are all more negative than those for the alkaline earth fluorides which fure ther illustrates the effect of charge. The effect of radius appears to be reversed for these tetra- valent salts, since Zr** which has the smallest radius also has the smallest negative deviation from ideality. The cause of this is not clear, although steric hindrance related to anion-anion contacts in the coordination shell adjacent to the tetravalent ion has been suggested as a limit- ing factor.®4 Thus any tendency by a tetravalent ion to have a high coordination number might 39 UNCLASSIFIED ORNL-LR-DWG 67568 -100 \\ —200 \ -300 k {cal) ) -400 \ \ i A\ -500 \ M HfFg — ZrFa -600 —— - \\ NoF -700 e S -800 0 0.010 0.020 0.030 2 (1 _NNOF) Fig. 12. Excess Chemical Potentials of NaF in Mixtures with ZrF,, HfF4, UF4, and ThF4 as Calculated from Liquidus Temperatures, be sterically limited for ions as small as Zr** and this might limit the magnitude of the devia- tions from ideal solution behavior. Another comparative study including polyvalent ions covering a relatively broad range of con- centrations was the emf measurements by Yang and Hudson ' 33 by use of cells of the type MCI 7 C]2 . (LiCl-KCI eutectic) For M = Pb2* Cd?*, Zn?* Mg?*, Be2*, activities were calculated from the relation 0 0y _ 1 In the five systems measured, the deviations from ideality were always negative (y] ~ 3000 o o W_ i | 2000 ~ o / () ~ / N /fl © S / (J\"b O’b / 1000 [ & L © & __Noc,‘\ ’\/ poCl2 \ 0 ~ ,/%°(J = S PbCl,-BaCl, PbCl, - LiC| -1000 : 0 02 04 0 02 04 0 02 04 2 2 2 Ny N Ny Fig. 13. Excess Chemical Potentials of CdCl,, PdCl,, and ZnCl, in Mix- tures with Other Chlorides as Obtained from EMF Measurements. The values of E? at 600°C used by Lantratov and Alabyshev were 1.4987 v for ZnCl,, 1.3382 for CdCIz, and 1.2215 for PbC|2, which are —4.3, =3.2, and 11.0 mv different from the values cal- culated by using the data in Table 6. Because of these differences in EY, the values in ,u’.lg must be considered uncertain by about 200 cal. Discrepancies in values of E? reported by different workers are common and are probably related to the solubility of metals in their own pure salts. In Table 7 are given values of y, at 600°C and values of the total volume change per mole of mixture, AV, at N, = 0.5. With an increase in the size of the alkali cation in mixtures with alkali chlorides values of v, decrease and values of AV increase. These relative variations in the deviations from ideal solution behavior and volume changes are in the same direction as was observed in alkali nitrates and in mixtures of AgBr with alkali bromides. As in the meas- urements of Yang and Hudson, mixtures of alkali halides with ZnCl,, CdCl,, and PbCl, exhibit less negative deviations from ideal solution behavior in that order E E E (Yznc1, <¥cact, <¥puci, M Fzact, Po., “mgcCl, where pure solid MgO and pure MgCl, are taken as standard states having activities equal to unity, K= where pgl and pg are the equilibrium partial pressures of Cl, and O, measured at equilibrium 2 2 with pure liquid MgCl, and pure solid MgO, and p, and p are the partial pressures at equi- 2 2 librium with a mixture. Reznikov approached the equilibrium from two sides and his results do not differ greatly from, but are probably more reliable than those of Treadwell and Cohen. His values of the activities of MgCl, are listed below: MgCl,-kcl MgCl,-NaCl T (°C) Mole % MgCl, {50 mole % MgC|2) 100 75 50 33.3 750 1.0 0.47 0.10 0.010 0.15 850 1.0 0.50 0.11 0.011 0.19 950 1.0 0.54 0.12 0.013 The activity coefficients in the mixture with NaCl are higher than in the mixture with KCl as expected. Other measurements using heterogeneous equilibria of the melt with a gas phase include the work of Blood'? and co-workers on the standard free energy of formation of NiF, in NaF-ZrF and LiF-BeF, mixtures using the equilibrium Ni + 2HF = NiF, +H, . The equilibrium quotient K, given by 2 PHF was constant in dilute solutions indicating that NiF, obeyed Henry's law. (HF and H, at the temperatures and pressures involved are essentially ideal gases.) The standard free energy . - . - » . . - * (chemical potential) of formation of NiF, in its standard state in solution (1y;g ) could be 2 calculated from the equation * 0 #Nin :ZGHF - RT In Ky - 47 Vapor pressure measurements afford a method of measuring activities in molten-salt mix- tures. Unfortunately, the large number of complex compounds found in the vapor often make it difficult to analyze vapor-pressure data. As there is no general discussion of this in standard texts, some of the principles involved in deriving activities from vapor-pressure measurements will be discussed.®? The chemical potential of a component in a mixture is related to the fugacity, /, of the component 'u"|=RT ]n f-l ’ (”.5.]) and for the pure liquid #(1) =RT In f?, /i o _p?_—_RTln —=RT In a, . (11.5.2) i The fugacity is defined in such a way that /,/p, + 1 as P+ 0, where p, is the partial pressure of the component in the vapor and P is the total pressure. In investigations of salt vapors it is generally assumed, and will be assumed here, that, except for the formation of associated species or compounds in the vapor, the vapor behaves ideally so that the fugacity of a species in the vapor is equal to the partial pressure of the species.* if only a monomer is present, 0 Py py=p;=RTIn—=RTIna,. (11.5.3) 0 Py If a vapor with a monomer vapor molecule represented by M, at total pressure, P, associates into several species (M])2 (M])3 (M])i M —_ —_— —— ] < < < I 2 3 i where (M!)2 is a dimer, (M])3 a trimer, etc., then the total pressure P is (if there is only one com- ponent in the vapor) P=( )+ @)+ (pdy+...=2(p,);, (11.5.4) where (p,), is the partial pressure of the associated species (M,).. When the vapor is at equilib- rium with a mixture (or pure substance) ity(mixture) = p,(vapor) . * Although this assumption may be valid, it has never been investigated. It is probable that at pres- sures approaching one atmosphere in alkali halides some of the interactions of the dipoles in alkali halide vapors are large enough to have an appreciable effect on the fugacity of the vapor even when the molecules are too far apart to be defined as an essociated species. 48 One total mole of component 1 in the vapor would have the chemical potential {per mole) (6y) (p4) (p4) (29); (), L L g b =Yt (11.5.5) ,u](vapor) = (p)] +7 (u - 3 (,u)3 + = where (1), the chemical potential of the associated species, is given by {); = RT In {p,); if the non ideality of the vapor is due to the association only, and (#,)./7P is the number of moles of species i in a portion of gas containing a total of one mole of M,. Because of the equilibrium, (Ju)] =— =, {(11.5.6) Combining Eqgs. (5) and (6) we get ,u](vopor) ={p); =—, (11.5.7) so that Py, RT (p), =——In P, iKY, I =RT|I‘161 py =g =RT In - (11.5.8) In order to measure the activity of the component 1, one need only know the partial pressures at a given temperature of one species containing 1 only which is in equilibrium with a mixture which is in equilibrium with a mixture or with the pure liquid component. At low pressures Eq. (8) is valid for component 1 independently of all other species in the vapor. The heat of vaporization AH]. of species j is given by d(‘u]./T) Rd In b v @ M = The variation in total pressure with temperature for any mixture or any number of species is P _ ] P ==L M= - XA, 11.5.10 d(1/T) k P d(1/T) Zp i, X AH,, (11.5.10) Rd In P p]-d]n b; where X], is the mole fraction of species j in the vapor, and j can be any species. Vapor pressures of mixtures have been measured by several methods. The Rodebush and boil- ing point methods#:21.33 120 make a measurement of the total pressure P. In the transport methods the vapor at equilibrium above a liquid is swept away with a known volume of inert gas and ana- lyzed. If only one component of the liquid is vaporized, then the apparent vapor “‘pressure,’” PT' is =PIT=p,+20p)), +3(p )y + ... =Zilpy),, (11.5.11a) 49 where PT* is an apparent ‘‘pressure’’ calculated assuming that the only species is the monomer. For more than one component Tr P,"= 2, (11.5.115) where P:' is the apparent transport pressure for the nth component and v is the number of mole- cules of 7 in species ;. The vapor composition at equilibrium with a given liquid to obtain association constants may be analyzed by using more than one experimental method under a variety of conditions and partial pressures. A complete analysis of the vapor in equilibrium with a mixture requires knowledge of the association constants for all of the species in the vapor. This analysis may be extended to the case where more than one component is vaporized. The precision of partial-pressure meas- vrements decreases very markedly the greater the number of species in the vapor. The vapor pressure of ZrF in equilibrium with mixtures of ZrF , with LiF'21 NaF 120 gnd with RbF33:121 have been studied by the transport method '29:121 and by the Rodebush tech- nique.3® Values of ,u';‘: for ZrF, at 912°C are given in Fig. 15 with the value of png {calculated 4 from the equation given by Cantor for the vapor pressure) at the measured melting point of 912°C; log pngd (mm) = 12.542 — 11,360/T (X) . (11.5.12) At high ZrF , concentrations the major species in the vapor is ZrF ,. Deviations from ideality are large, and are larger the larger the alkali cation. Some of the values of the vapor pressure used in these calculations were not directly measured but were extrapolated from other temperatures. Although Sense and co-workers report vapor pressures of the alkali fluorides in these mixtures, they do not in any case properly correct for the presence of associated species in the vapors. Cantor and co-workers33 have reported that in RbF-ZrF , mixtures the excess entropies are positive and that the excess enthalpies as obtained from temperature coefficients of vapor pres- sure data exhibit both positive and negative values. Although these conclusions are more reli- able than those obtained from emf data, the temperature coefficients are subject to large errors, The most thorough study of the vapor pressures of a molten-salt mixture is that of Beusman,® who partially studied LiCl-FeCl, mixtures and studied more completely KCI-FeCl, mixtures at temperatures from about 850 to 1000°C by using the Rodebush technique for measuring the total pressure and the transport method for measuring the vapor composition. In the vapor above mixtures of KCI and FeCl, the presence of the species FeCl,, Fe,Cl,, KFeC|3, KCI, and K2C|2 was consistent with his measurements. Calling these species 1, 2, 3, 4, and 5 he could solve for the number of moles of each of these species in a unit volume of vapor and, hence, for the partial pressures at equilibrium with the melt from measurements of the total pressure of salt (P == _RT/V), and by a chemical analysis of the chemical compounds swept 50 UNCLASSIFIED ORNL-LR-DWG 67572 16,000 ROF -ZrF, 14,000 4 912°¢ / 12,000 / — 10,000 @ o £ > NaF - ZrFy O : ] L V. 1 | // o 4 A /. S 2000 M//)/ N Fig. 15. Excess Chemical Potentials of ZrF, in Mixtures with LiF, NaF, and RbF as Obtained from Vapor-Pressure Measurements. out by a unit volume of gas from above the pure components and from above the mixtures. He solved the simultaneous equations n,=ny+tnyg+ngtn, +n,, ne =ng+n,+ 2, 2 "y ”1=KFeC|2”2 5/ nU 2 nyg=Kecr s <_P . 51 where _ is the total number of moles of all species in a unit volume, 7 _and n. are the number of Fe and K ions in a unit volume, and K and K are the dissociation constants in pressure FeC|2 KClI units for the dimer of the subscript component and were evaluated from data obtained with the pure materials KCl and FeCl,. The presence of a trimer in LiCl vapors made this procedure very im- precise since the calculations, which essentially involve subtracting large numbers, are much more sensitive to errors in the measurements when more species are involved. In Fig. 16 are plotted Beusman's values of uf for the two components FeCl, and KCI at 900°C derived from the values of the partial pressures. It is apparent that values of uF are all negative. Calculation of #II:;CI from ,uléeaz by integrating the Gibbs-Duhem relation E E N,duy + Nyduy =0 leads to about the same values as were measured. The deviations from ideality of KCl are somewhat greater than those for FeCl,, and the appar- ent values of both the excess entropies of mixing and partial molar heat of solution are positive. Barton and Bloom? have measured the vapor pressures of PbCl,-KCl, CdCl,-KCl, and CdCl - 113 NaCl mixtures at 900°C by using boiling point and transport methods. At concentrations of UNCLASSIFIED ORNL-LR-DWG 67574 10,000 ,'LFeCIa 8000 » 900°C Hxel 3 6000 /o 4 £ @ g / S //// ° w 4000 L/ 1 : 2000 // 0 0 0.1 0.2 0.3 0.4 0.5 e Fig. 16. Excess Chemical Potentials of KCl and FeCl, in KCl-FeCl, Mixtures as Obtained from Vapor-Pressure Measurements. 52 less than 60 mole % alkali halide they could neglect the volatilization of the alkali halides and vapor compounds containing alkali halides. Their results are in fair agreement with the emf meas- urements of Lantratov and Alabyshev®® on the PbCl,-KCl| system. They found that the apparent deviations from ideal behavior in the CdCl,-KCl| system were smaller than in the CdCl,-NaCl sys- tem. The measurements in the systems containing CdCl, are open to question. 1.6 Discussion of Binary Systems with a Common Anion The results in the previous section exhibit certain very general features for mixtures of salts of a monovalent alkali cation with salts of a polyvalent cation. The most obvious feature is the variation of the thermodynamic properties with cation radius and charge. The deviations from ideal solution behavior of both the alkali ion salt and the salts with polyvalent cations usually become more negative (or less positive) with an increase in the radius of the alkali cation and with an in- crease of the charge or decrease of the radius of the polyvalent cation. This type of behavior has often been ascribed to *‘complex ion formation'’ or to ‘‘complexing.’'’*3+62:85:97 Thig terminology has been used so freely and in so many different senses that some of the “‘explanations’’ of solu- tion behavior in terms of "‘complexes’’ are merely redundancies of the observed facts and add noth- ing to the understanding of solution behavior in terms of physical concepts. As a consequence, in this section, a discussion and critique will be given of this concept, Among the most reasonable and careful considerations of the concept of complex ions are those of Flood and Urnes>> and Grjotheim.®? Flood and Urnes, for example, discuss the liquidus curves of RbCl, KCI, and NaCl in mixtures with MgCl,. They reason that a mixture of an alkali halide with an alkali salt of a large divalent anion will exhibit only small deviations from ideality. Evi- dence for this comes from the apparently negligible deviations from ideality found in the liquidus curves of Na,50, in mixtures with NaCl and with NaBr.?2 (Note the work of Cantor on the parallel effect of cation radius.) Flood and Urnes propose that the component M,MgCi , containing the MgCl42' grouping would exhibit small deviations from ideality based on the Temkin definition. Thus at low concentrations of MgCl,, n n — 2n ci™ MCI MgCI2 Ao = = (11.6.1) MC . e n + 7 7 -~ n cI- Mc142‘ MCI ™ TMgCl, The procedure of Flood and Urnes is essentially a redefinition of components. They show that the liquidus temperatures’® (and activities) for KCI and RbCl are in reasonable agreement with Eq. {1). The liquidus temperatures (activities) of NaCl in NaCl-MgCl, mixtures exhibit positive devia- tions from the calculations based on Eq. (1). This was ascribed to a partial dissociation of the MgC|42" ion. Thus, by the redefinition of components, and by the careful choice of systems, rea- sonable correlations with the data were obtained. Although such a procedure has the advantage of being simple, there are many criticisms which can be made, The major criticism, perhaps, is that this method can be applied to very few systems 53 and does not lead to quantitative predictions which can be made a priori. For example, the lig- uidus temperatures of NaF in NaF-BeF , mixtures are too low to be described by any redefinition of components which is consistent with possible structural concepts. Since the Be?” ions are so small, a coordination of Be2" cannot be expected larger than four and yet a mixture of, for ex- ample, NaF and Na_BeF, would have to be described as exhibiting negative deviations from ideal solution behavior. On the other hand, no reasonable choice of a ‘‘complex ion'' grouping or com- plex component can be invoked to explain the small deviations from ideality of NaF in NaF-BeF, mixtures, Further, although the thermodynamic data may be described by choosing a particular complex component, this does not necessarily imply the existence of the ions of this component in the melt. Except for the very stable (relative to the separate ions) complex ions as NO,~, PO43", and 5042', a simple comprehensive description of the solution behavior of mixtures with a common anion cannot generally be made with only one *“complex ion'' and little can be learned about solu- tion behavior a priori from such an approach. The absence of a simple explanation of the solution behavior of molten-salt mixtures is evident in NOF-ZI‘F4 mixtures shown in from the analogy between a in HCI-H O mixtures and a H,O ZrF 2 Fig. 17. In water the O-H interaction is very strong so that it is only slightly ionized. At low UNCLASSIFIED ORNL-LR-DWG 66350 1.0 //, 4 7 0.8 7/ 0.6 Y 7/ 02 // 0.4 / / / // / / / / // NaF - ZrF, A // / 0 0 0.2 04 0.6 08 10 Na Fig. 17. Plot of the Activity of H,0 in H,0-HCl and H,0-HBr Mixtures at 25°C and the Activity of ZrF, in ZrF -LiF, ZrF -NaF, and ZrF ;-RbF Mixtures at 912°C. 54 , , : + - . concentrations of HCI the solution may be understood in terms of solvated H and Cl~ ions inter- acting in a dielectric medium, In mixtures with HBr the activity of water is even lower than with o : , + HCI. The limiting law at concentrations of HCI greater than the concentration of H™ from the self ionization of water is 0 [aY] ,quo—,quo = RT In (]—QNHCI), (11.6.2) since HC| behaves as two particles. lfy,, is defined by 2 0 - =RT In N , (11.6.3) FH,0 THH 0 H,0YH,0 then at low concentrations 1 - 2NHC! - <. (11.6.4) YH.0 Interionic interactions of H® and Cl™ will cause ¥y _o to differ at higher concentrations from the value given by (11.6.4). Beyond the range of validity of the Debye-Hiickel theory, this is unpre- dictable although there is a persistence of the negative deviations from ‘‘ideal’’ solution behavior. In NaF-ZrF , mixtures, the solvent ZrF , may be considered to be more highly ionized than water. Consequently, the self ionization of ZrF , and the one particle limiting law will probably hold to higher concentrations than in water so that the deviations from ideality in dilute solutions based on an equation such as (I1.6.3) will be smaller in the Nc:F-ZrF4 mixture than in HCI-HQO mixtures. Apparently the smaller (less negative) deviations from ideal solution behavior in the NaF-ZrF , mixture as compared to HCI-H,0 mixtures persist at high concentrations. Just as with water, the farger the size of the *‘foreign” ion (CI~ and Br~ in H,0 and Li*, Na*, and Rb*in ZrF4) the greater the deviations from ideality. This does not explain the observed solution be- havior in NaF-ZrF , mixtures but merely suggests that any fundamental explanation in concen- trated solutions is at least as difficult as in concentrated solutions in water where it is clear that the H* and CI~ ions are solvated but no valid quantitative predictions can be made in terms of structural concepts. Because of these apparent inadequacies of the concept of ‘‘complex ions'’ in describing solution behavior in mixtures containing one type of anion it is in order to discuss and attempt to classify some of the effects and interactions which have been included in the terms ““‘complex ion'’ or '‘complexing’’ in the hope that such a procedure would be more instructive and useful in future attempts at deriving quantitative theories. Most definitions of ‘‘complex ions'’ or of “‘complexing'’ fall into two categories. In the first category a '‘complex ion’’ is usually conceived as a microscopic structural entity. A complex ion can be most clearly defined as a grouping of at least one central cation and near- neighbor anions having a particular configuration. |f each grouping is isolated from others and shares no anions, then the grouping is a finite complex. NO, ™, P043_, and 5042_ ions are finite complexes, If the groupings are all interconnected by shared anions, then infinite complexes are 55 present. By these definitions all pure salts are infinite three-dimensional complexes and very di- lute solutions of one salt in another always contain finite complex ions although the configura- tions of all the finite complexes are not necessarily uniform. X-ray, uitra-viclet, infra-red, Raman spectra, and other methods of investigating structure are means for investigating these complex ions. “‘complex ions’’ or ““complexing’’ are used to describe a tend- In the second category the terms ency to stabilization. This is the least satisfactory use of this terminoclogy, since so many differ- ent interactions and concepts are included in this usage that less information is conveyed than by the use of the word stabilization. For pure materials as for example AgCl, NiF, or HgCl,, specific interactions (van der Waals’ ligand field and covalent binding) give rise to more negative values of the energy of formation from the isofated ions than might be expected for alkali or alkaline earth halides where Coulomb inter- actions are relatively more important. |n solutions the tendency to '‘complexing’’ or toward the stabilization of a component in solution is characterized by negative values of uE. Some of the solution effects which influence the values of p& are: (a) Coulomb effect. The discussion in section Il shows that Coulomb interactions in mixtures of salts containing monovalent ions lead to negative values of u. This effect appears to be pres- ent in mixtures containing polyvalent cations, Long-range interactions are very significant in this effect and as a consequence a quantitative description of this effect in terms of finite complexes can only be fortuitously correct. (b) Polarization effect. The field intensity at an ion position will, in general, not be zero be- cause of ionic motions and because of the difterent sizes and charges of cations. For example, an anion having two cations the same size but of different charge as near neighbors will tend to have a Coulomb field intensity on it. As a consequence, the electrons on the anion and the thermal mo- tions of the anion will be "polarized’’ so that the negative charges reside a greater fraction of the time near the cation with the higher charge. In a pure molten salt this effect will be expected to be smaller than in a mixture, and the net contribution will lead to a relative stabilization of the mixture (negative contribution to the deviations from ideal solution behavior). (c) van der Waals' interactions. As in mixtures containing monovalent cations these interac- tions usually will lead to a positive contribution to the deviations from ideal solution behavior for systems containing polarizable cations. To illustrate with a clear-cut example, the systems NaCl- PbCl, and AgCI-PbCl, might be compared. In the former the measured deviations from ideal solu- tion behavior of PbCl, are negative (#Ebcl < 0), and from the Gibbs-Duhem equation it can be shown that P’EoCl is also negative, Measurzements in the latter system] 16 indicate that p'fgCl (and “EbCI ) is essentially zero at all concentrations. The major differences between these two sys- tems are probably related to the high polarizability of the Ag’ ion as compared to Na* and hence to the contribution to van der Waals’ interactions. Quantitative estimates of the magnitude of this effect in such systems are tenuous at present. 56 (d) Ligand field effects.'®3 These interactions will tend to stabilize pure salts of transition metal ions and particular configurations of near-neighbor anions will tend to be more probable. Such stabilization, regardless of the specific symmetry of the near-neighbor anions, will tend to be monotonic with the strength of the negative ligand field. For given anions as near neighbors to a particular transition metal ion in a mixture, the negative ligand field will be attenuated by more distant cations with the attenuation tending to be smaller the smaller the charge and the larger the radius of these other cations. The dissolution of a transition metal salt, NiF, for example, in an alkali fluoride would lead to a replacement of next-nearest neighbor Ni2* by monovalent alkali cat- ions. This will lead to a stabilization of NiF, (,ufiiF < 0), which would be more pronounced the larger the alkali cation. The influence of ligand-field interactions will be limited by steric require- ments and in mixtures with alkali metal salts will probably lead to negative contributions to the deviations from ideal solution behavior of both components, (e) Packing and steric effect. To satisfy the tendency toward local electroneutrality it is probable that small highly charged cations will tend to have a larger number of anions as near neighbors than cations of low charge. Any energy changes (stabilization) related to this effect will be sterically limited in accordance with the values of the anion-cation radius ratios. All of the factors mentioned are included in the concept of ‘‘complex ion"' of or “‘complexing’’ when it is applied to stabilization, Some of these effects may be concomitant with a foreshorten- ing of cation-anion distances (e.g., coulomb, polarization, and/or ligand field) or with a tendency toward specific configurations of anions about cations (e.g., ligand field and/or packing). In all cases, these factors influence the free energy differences between pure salts and salts in solution, It may be preferable to refer to the observation of negative values of uF as a stabilization, since such a stabilization is not necessarily related to the observation of a ‘‘complex ion'’ as a struc- tural entity. By this usage, no unwarranted implications about the structure of the melt need to be made, The existence of solid or gaseous compounds which are made from the two salts in a solution cannot be used as evidence that particular ““complex ions'’ are formed in solution. Although many of the factors and interactions which lead to relatively greater stability of gaseous and solid com- pounds may also give rise to negative deviations from ideal solution behavior, many of the factors influencing the structure of solids or gases have no counterpart in liquids. For example, in solids steric repulsions of the ions are more important than in liquids and have a strong influence on structure; and in gases the entropies of association are generally negative and give rise to a strong influence in favor of forming the simplest compounds. Kinetic definitions of ‘‘complex ions’’ in terms of the lifetime of a grouping or of the relative mobility of ions2? cannot be clearly related to equilibrium thermodynamic properties or to ‘‘complex ions’’ as a structural entity unless these lifetimes are very long. . . , : E : , Since there is no adequate theory for most binary mixtures, u~, for any component in a given mixture containing polyvalent cations, must be estimated empirically by comparison with known systems containing mixtures of the same charge type and the same anion. Keeping in mind the 57 types of interactions which influence the values of ,uE, reasonable estimates may be made by anal- ogy with known systems or by interpolation., The development of a theory, as, for example, by the extension of the perturbation theory of conformal ionic mixtures to mixtures containing cations of different valence, would be an aid in such estimations and might be used to confirm empirical re- lations such as was proposed by Kleppa. I1.7 Other Systems Measurements of the activities of lead halides in the mixtures F’bC|2-ZnC|2 (ref 131) and PbBr-ZnBr,, (ref 117} indicated small negative deviations from ideal solution behavior in the for- mer and small positive deviations from ideality in the latter. (Calculations of the activities of ZnCl, in the first system by use of the Gibbs-Duhem relation were in reasonable agreement with ]26) I activities calculated from measurements of the partial pressures of Zn(.:l2 in this system, n these two systems there is no difference between the mole fraction of a component (NPbX ) and the product of the ion fractions (NPbNi =Np, =Np,x_) and there is little ambiguity in defining 2 activity coefficients, On the other hand, in a system as F’bCIz-F’bBr2 there is some c:mbigui‘ty,28 : 2 y2 N2 since Np /Ny =Ny = prxz. cients depends on the type of compounds, |f the lead halides were very stable molecular com- In such systems, consequently, the definition of activity coeffi- pounds and did not react with each other (were not molten salts), then the activity coefficient PbX, ™ NPbx2}’Pbx2° lecular, and where one might consider the exchange PbCl, + PbBr, == 2PbCIBr), in order to be would be defined by « For ionizing salts {or where the compounds are mo- consistent with the limiting laws, the activity coefficient is better defined by a, , = NowNX¥Yppx_ * On this latter (and more realistic) basis, the activity coefficients, Yppge,. N F’bC|2-F’|:>Br2 mixtures, are larger than unity. V17 The choice is not always clear-cut as many com- pounds cannot be strictly classified as either molecular or ionic salts, Very few other measurements on binary systems that have a common cation have been made. Precise measurements by Toguri, Flood, and Fgrland®® on the exchange equilibria Cl, + 2MBr = 2MCl + Br, (1.7.A) in LiCl-LiBr, NaCl-NaBr, and KCI-KBr mixtures were used to investigate the activity coefficients of the alkali halides in these mixtures. The equilibrium constant for (A) is 2 2 2 NMCE}’MCllblzs.r2 YMcl K = _ K’ , (11.7.1) M N2 2 M 2 MBrYMBr cl, YMBr where K;A is the measured equilibrium quotient, Taking the logarithm of K|, and using as a first approximation for the activity coefficients in any one binary system 2 a2 RT Iny,c =ANyg, 9nd RT Iny,g, =AN{cy, 58 then RT In K’ = RT In K + 2ANZ - N2 ). (1.7.2) Plots of RT In K* vs (N2 —~ N2 ) led to the values of A in Table 10, which indicate small posi- MBr MCl tive deviations from ideality. For these relatively large and polarizable anions, packing or van der Waals' effects have been proposed as possible contributing factors. To contrast this, an analysis prop p g 4 with the data in Table 1 indicates small nega- of the liquidus temperatures of LiF-LiCl mixtures tive deviations from ideal solution behavior for both components. Since the F~ ion is smaller and less polarizable than Br~™, it would seem that at least one of these two properties of the ions is significant, Table 10. Values of A (cal) from Equilibrium Measurements in Binary Systems with a Common Cation (M) Li Na K MBr-MClI 150 350 530 M2Cr 207.-M 2Cr04 0 ~300 ~ 500 Similar measurements of the equilibrium M,Cr,0,=M,Cr0, + %Cr,0(solid) + %02 (11.7.B) . . . 4+ o+ + in molten mixtures of chromates and dichromates have been made?? for M = Li ,Na, K, or TI". The equilibrium constant, if CrO42_ = X%~ and Cr2072' = Y2-, is given by 3/4 Nx(?’oz) Y, X Y, X Ky = ———— - K}, ; (11.7.3) N Y Y Y M,Y M,Y by using the approximation that RT In YMox = ANAZA yand RT Iny, = ’\N:\ x the values of A M, X M2Y 10 and are seen to be small, When M* was an alkali ion the stability of M,Cr,0O, relative to could be obtained from the slope of a plot of In K!:A vs (N ). These are given in Table M,CrO, increased as the size of the M?* cation increased and consequently the equilibrium con- stant K, (and the equilibrium quotient K‘;) for reaction (B) decreased with an increase of the size of M*. This is also true for the equilibrium in reaction (A). These facts are useful for anticipat- ing some of the properties of molten reciprocal salt systems discussed in section lll. For example, consider the equilibrium (B) in a mixture of N02Cr207 and N02Cr04. The equilibrium constant is given by ~RTInK., =F®° + 1 Fd - F° (11.7.4) Na N02Cr04 2 Cr2C):3 Nu2Cr207 39 and is, of course, dependent only on the properties of the pure reactants. If the composition of the mixture is altered so that the Na* ion is gradually repiaced by K * ion, the equilibrium (B) will gradually go more to the left and K will decrease. When very little Na®ion is left and the melt is essentially a mixture of K,Cr,0, and K,CrO,, the value of K will be equal to the value of Kl‘( in the mixture containing only the K cation. For this case one obtains from Eq. (3) In K =InKg=InK ~In (yNuzx/yNa2Y) =In K, —In (szx/szY), (1.7.5) where the activity coefficients are all in a solution containing mainly K* ions and very little Na* ion. Introducing Eq. (4), one obtains In (yNazx/yNazY) = In (K /K) + In (szx/)’sz) Ayo - + In (szx/szY) , (11.7.6) where A#O is the free energy change for the reaction of the liquids in (C) N<:2Cr207 + K2Cr04\‘: K2Cr207 + Na,CrO, (11.7.C) and the last term in Eq. (6), In (yK x/yK Y), can be seen to be small in this case from the data in Table 10. The value of Au® is negative and the ratio of the activity coefficients of the components Na,CrO, and Na,Cr, O, is much greater than unity, and in simple cases such as this, Na,CrO, ex- hibits positive deviations from ideal solution behavior and Na,Cr,0, exhibits negative deviations from ideal solution behavior. Thus N02Cr04, which is a member of the stable pair in reaction (C), exhibits positive deviations from ideality and Na,Cr,O,, which is not a member of the stable pair, exhibits negative deviations from ideal solution behavior. This tendency is present in all recip- rocal systems. Flood and Maun4? have measured In K”as a function of the ion fraction of Na* in mixtures of Na*, K, Cr042', and Cr2072' ions. A plot of In K’ vs Ny, given in Fig. 18 can be seen to be nearly linear in the cation fraction. The data fitted the equation In KNa,K=NNaIn Kiat Ny In KK+bNNaNK’ (11.7.7) where b is —0.2 at 662°C. Similar measurements in the TI7, K+, Cr042", Cr2072' system are plotted in Fig. 18. The quantity b is discussed by Flood and Maun, is related to the proper- ties of binary mixtures made up from the four ions in the system,??and is probably small when all the binary systems have small deviations from ideal solution behavior. These properties of recip- rocal systems have been used in an ingenious derivation of a zeroth order theory of these sys- tems.®3+%4 A more complete description of reciprocal systems is given in section |ll. Since linear relations are often useful from a practical point of view, two linear relations which apply to ternary systems having a common anion will be stated.>® These apply to ternary systems in which the solution properties of two of the components (components 1 and 2) do not differ greatly, mixtures of these two components do not exhibit large deviations from ideal solution behavior, 60 UNCLASSIFIED ORNL-LR-DWG 67570 0.5 0 In the previous section Ay and Ap have been discussed. The term Ap% (or AH%) is related to a variety of types of interactions. When values of A,qu cannot be obtained from tables, it is sometimes useful to be cognizant of one of the major influences on Apg, that of coulomb interac- tions. For the alkali halides, for example, the largest contribution to Ap% is the Madelung term . 2( 11 1 ) - Ae + - - , dAX dBY dAY dBX where dil' =TT where 7. is a cation radius and r is an anion radius. |t can be shown that if T Tyr OF T, >7p and r, < ry then the Madelung term is positive. This tendency leads to the general reciprocal Coulomb effect which is valid for all the alkali halides. This ef- fect is such that in a reciprocal system with two cations and two anions the two stable components (stable pair) as evidenced by A,ui are the small cation-small anion component and the large cation- large anion component. These two components** would tend to exhibit positive deviations from ideal solution behavior and the other two negative deviations. From a consideration of the Made- lung term one would expect positive deviations from ideal solution behavior for the stable pair to increase in the order [NaF-KCI] < [LiF-KCI] < [LiF-CsCH < [LiF-CsBr]. The last two systems ex- hibit such large deviations from ideal solution behavior as to have liquid-liquid miscibility gaps which have been observed.® The reciprocal Coulomb effect probably applies for salts of different valence containing nonpolarizable ions and is in such a direction that in a given system the salts *The criteria which are discussed and used by Bergman and associates are the values of AG(A'u) of the solids at room temperature which in view of their crude correlations are equivalent to Ap_o or AHO. **All four components are not indepsendent of each other and only three of the four are true components in the Gibbs sense, 63 with the smallest or most highly charged cations and smallest or most highly charged anions will tend to be a member of the stable pair. Obviously, the Coulomb effect is not the only important one and many deviations from the gen- eralization are to be found, especially for systems containing polarizable ions. For example the reaction AgNO ; + NaCl == AgCl + NaNO, has a large negative value of Au® or AH? (about —15 to =17 keal/mole) which is considerably more than the Coulomb effect and which is probably the result of the large stabilization of AgCl by van der Waals' (London dispersion) interactions '®! of Ag” and CI~, By contrast with the binary systems discussed in section |l, the interactions are, in general, much larger in reciprocal systems as they are mostly between nearest-neighbor cations and anions rather than next-nearest neighbors and consequently one would expect to find many reciprocal sys- tems with very large deviations from ideal behavior. In the following chapter some of the theories will be discussed which have been advanced for these systems beginning with the simplest approx- imation and continuing with approximations of increasing complexity. {I.2 The Random Mixing Nearest-Neighbor Approximation®* This derivation is based on the Temkin quasi-lattice model. For the simplest member of this class of systems, that containing the two cations A* and B* and the two anions X~ and Y™, the model is an assembly of charges in vacuo and consists of two interlocking sublattices, one a lat- tice of the cations A* and B* and the other of the anions X~ and Y™, The nearest neighbors of the cations are anions and of the anions are cations. The total entropy of mixing is AST/R =-2n InN, - En]. In N, and for any component is Eij - S:.)]. =~R In NN, where 7 and j are cations and anions respectively, All of the ions have the same coordination number Z. The model is restricted so that all of the ions of the same charge are the same size. This restriction eliminates any difference in the long-range Coulombic interactions be- tween either A” or B* ions or X or Y ions and their respective environments, and limits the model to short range extra-Coulombic effects which are assumed to be nearest-neighbor interac- tions.* The form of the equations derived will probably apply even to systems with different-size ions. In Fig. 19 is a two-dimensional representation of the quasi-lattice. If the pair interaction energy of AT.Y = is €, of B*.X~, €,, of AtX-, €5, and of B*.Y-, €4 then A€=€4+€3-€2-El-—--fi-—, (111.2.7) *|f random mixing is assumed, or for a dilute solution only one pair need be the same size to eliminate differences in the Coulombic interactions. 64 UNGLASSIFIED ORNL-LR-DWG 31157A gt ry— Bt r- gt r- gt r- y= |4t @ B* + y- gt @ Bt gt y- Bt y- Bt y- Bt r- . . ) [ *A* — gt ry- Bt y- Bt y- gt r- y~ |4t @ gt + y~ 8t @ g* Bt r=— BY r- gt y- Bt r- Fig. 19. Two Dimensional Quasi-Lattice Representation of the Process A++ X~ &= AX in the Solvent BY, where A¢€ is the energy change for the interchange of the circled X~ and Y~ ions and is the energy of formation of the ion pair A*-X=. If there is random mixing of the cations and of the anions on their respective sublattices, then the fraction of positions adjacent to any given cation occupied by a given anion will be equal to the ion fraction of that anion. The assumption is made that the relative energy of each A*-X~ pair is Ae. This is equivalent to the assumption of the noninter- ference of pair bonds or to the assumption of the additivity of bond energies. Since the total num- ber of positions adjacent to any ion is equal to n.7 or E].Z, then the total energy or enthalpy of the solution is E :ZAZNY(E + K) +ZB ZNX(€2 + K) +5A2Nx(€3 + K) +EBZNY(64 +K)=H ; ] (111.2.2) T f where ZK is the value of the energy of interactions of the A* or the B* ions with ions beyond the nearest-neighbor anions. The partial molar enthalpy or energy of solution is 0 _ & 0 / HI.]. - Hz.]. = Ez.]. - Ez.]. =2 (1=~} - N].) ZAE (111.2.3) where the — sign is pertinent when ij is AX or BY, and the + sign is pertinent if it is AY or BX. Remembering that u.. = H.. - TS, then 17 ty 1y i =gy =t (1= N)1=N) ZAE + RT In NN, (1.2.4) and RT In Yii = i(]—Nl.)(]—N].) ZAE . (111.2.5) 65 The derivation of Eq. (5) is implied by the work of Flood, Fgrland, and Grjotheim who have, how- ever, emphasized a somewhat more general relation. Equation (5) is strictly valid only for cases in which AE is small relative to RT so that one might reasonably be close to random mixing of the ions. The form of Eq. (5) is probably valid in some cases where there is only a small deviation from random mixing and is instructive and important for the qualitative understanding of solution behavior, Flood, Fgrland, and Griotheim propose a method for making a crude estimate for ZAE from the heat contents of the pure components. Figure 20 is a two-dimensional quasi-lattice repre- sentation of the metathetical reaction (l11.1.A) for which the heat change is AH® per mole.* Since all ions of the same charge are the same size, only extra-Coulombic nearest-neighbor interactions are changed in this reaction, Since the number of nearest neighbors for each of the salts is 7l per mole of salt, then for the reaction AH® = ZAE, if each pair interaction energy were the same. In real systems the pair interaction energies are probably a function of the number and kinds of anions which are nearest neighbors to a given cation so that AE will not be truly constant and will only be roughly approximated by (AH%/7). A relation analogous to (5), but somewhat more general, has been derived by Flood, Férland, and Grjotheim>* where Ayo is the change of chemical potentials for the metathetical reaction (I11.1,A). RT Iny, = +(1=N)(1-N) Aul, o ~ *For many reciprocal salt pairs probably ASO = ( so that AHO = A‘u . ORNL-LR-DWG 31158A 0 UNCLASSIFIED Aty a4t y- Bt x- Bt x~ y™ a4t r- a4t + x~ g% x gt Aty oAt yo gt x- gt x~ r *’j \ At x- AY x° gt vy~ Bt r- x~ At x4t + x~ BY v~ BT At xT At XxT gt vy~ B8t y- Fig. 20. Two Dimensional Quasi-Lattice Representation of the Metatheti- cal Reaction AY(liq) + BX(lig) &= AX(lig) + BY(liq). (111.2.6) 66 If the deviations from ideality are large enough, then the solution will tend to separate into two liquid layers. Since the theory is symmetrical in composition, the upper consolute temperature, T, below which temperature two liquid phases will form, will be at a composition such that N, = Ny=Ng=N,= ]/2 It may be calculated from Eq. (5) or (6) by setting the derivative da, /dN , equal to zero in mixtures of AY and BX, where Na=Ny=N,y=Ng=N, and Npy is the mole fraction of AY in a mixture made up from the salts AY and BX. The expression for the upper con- solute temperature derived from Eq. (5) is ZAE ., AHO - = (1.2.7) € 4R 4R and from Eq. (6) A 0 _2r (111.2.8) € 4R To illustrate Eqs. (5), (6), (7), and (8) let us consider the dissolution of a mole of liquid AgCl in NaNO,, where the ions Ag+, No+, NOa—, and Cl~ correspond to AT, B X", and Y~ respectively. From published data®?:78:114 51 the pure salts, Au® = +17 keal and AH® = +15 keal at 455°C. It can be seen from Egs. (I11.1.5) and (l11,1.6) that the components AgCl and NaNO ., bers of the stable pair, should exhibit positive deviations and AgNO ; and NaCl should exhibit which are mem- negative deviations from the Temkin ideal-solution behavior. The results are similar for the sys- tem Ag ' I(+, NO3-, Cl™. In both these systems the calculated upper consolute temperature is well above the melting point of all the possible components that can make up the system, and two im- miscible layers are present in this system. However the measured upper consolute temperature is much lower than that calculated from Eqs. (7) and (8). Similarly in the system Li*, K, CI~, F~, where the stable pair is LiF-KCI, the values of A,uo and A% at 1000°K are about +17 keal;3159.78 yet two liquid layers have not been detected in the quasi-binary system LiF-KCl,>! although the calculated consolute temperature is very much higher than the measured liquidus temperatures. Clearly Au® and Al are not the sole measure of the deviations from ideality in reciprocal molten- salt systems, In mixtures for a given class of salts, such as alkali halides, they probably serve as a guide to the relative deviations from ideality. For example, the positive deviations from ide- ality in LiF-KCI quasi-binary mixtures are greater than for the NaF-KCI mixtures. The values of 51 An analysis of the quasi-binary Ap® for these two systems are +17 and +8 kcal respectively. liquidus temperatures for LiF-KCl and NaF-KC! in which the stated components exhibit positive deviation from ideal behavior and of the liquidus temperatures for LiCI-KF and NaCl|-KF mixtures in which the stated components exhibit negative deviations from ideal behavior has shown that Eq. (5) or (6) only describes the solid-liquid equilibria in a semiquantitative manner.3' The short- comings of these two equations stem from a variety of possible reasons. Fgrland®® has discussed the influence of those interactions which reciprocal systems have in common with binary systems containing either two cations and one anion or two anions and one cation. As discussed in section Il, these interactions are of longer range than nearest-neighbor interactions. Fgrland has discussed 67 this possibility for the hypothetical case in which this effect can be described in terms of the equa- tions of regular solutions. From the derived relations it can be shown that if the binary systems ex- hibit negative deviations from ideality, then the correction terms to Egs. (5) and (6) are in a direc- tion which makes the activity coefficients smaller and which lowers the calculated upper consolute temperature. Although this correction is in the right direction, it is not large enough to lead to a good correspondence of calculations with experiment. As discussed in the following sections it will be shown that two other important effects which have been experimentally demonstrated are present. One effect is related to the nonrandom mixing of the ions which, except for extremely small deviations from ideality, leads to magnitudes and a concentration dependence of the devia- tions from ideality which are very different from Eqgs. (5) and (6). The second effect is the non- additivity of pair bond interactions. I11.3 Corrections for Nonrandom Mixing: The Symmetric Approximation For the case in which AE is not very small relative to RT, corrections for nonrandom mixing of the ions must be included. Flood, Fefland, and Grjotheim have given a preliminary discussion of nonrandom mixing.>* Explicit calculations based on the nearest-neighbor quasi-lattice model 14,18 1 1d Blander and Braunstein. ' 2 have been made by Blander, In the following sections approximations based on the quasi-lattice model will be used to cal- culate the effect of nonrandom mixing (or associations) on the calculated thermodynamic proper- ties of the model system., These calculations will also be related to conventional association con- stants for associations of the A" and X ions to form ““complex ions”’ + - Hm~- mAY & aX :_AAm Xn {m=n) and will be used to illustrate some of the properties of these constants. It should be noted that some of the relations derived may also be derived without the use of a quasi-lattice model.'? The model is useful in defining the parameter Z and in the statistical counting in the theoretical calcu- lations. In dilute solutions of A* and X ions in BY most of the associated species {or ‘‘complex ions'’) Aanflm"") are isolated from one another by solvent B* and Y~ ions and are easily defin- able. This is in sharp contrast with solutions having only one kind of anion where complex com- pounds are not easily defined since all cations will have the same anions as near neighbors re- gardless of the properties of the solution. 12 63 The symmetric approximation'“ is essentially the quasi-chemical theory of Guggenheim.®® |n this approximation as in the others in section ||| only nearest-neighbor interactions are taken into account, The assumption is made that the interaction of any given adjacent pair of ions is the same independent of the local environment. A given A* ion may interact as many as Z X™ ions and a given X~ ion may interact with as many as Z A* jons with the relative energy of each interaction being Ae. The total number of the pairs A*-X", B+-X-, A*Y", and B*-Y is Z(n, + nB)n. If v’ 68 is defined as the fraction of positions adjacent to the A* ions that are occupied by X~ ions, then the number of pairs of each kind and the total energy of such pairs are given below: Type of Pair Totol Number Total Energy Aty Zn (1 =Y M =R ~5’ 0 B x~ Zlne~n, ¥ WM =5] 0 Atx” Zn, YN =5’ Zn, Y'AE By~ Zng—nc+n, ¥ M =R, -5/ 0 For simplicity the relative energies of the pairs other than A*-X~ are arbitrarily set at zero. This makes no difference in the final results. R’ and R; are the number of positions adjacent to all the A" and B” ions respectively; S. is the number of positions adjacent to the A% ions occupied by X~ ions; and Sl: is the number of positions adjacent to the B* jons occupied by X ions. The number of ways of distributing these pairs, w_, is (R; + R))! @ = TR AT (11.3.1) (R, — SNSI(R, = S)LS] 6 As in the quasi-chemical approximation,®3 when w, is summed over all possible values of Y7 the value for the total number of configurations is incorrect. A normalizing factor can be calculated to correct this so that the combinatory formula is o sTURE = sDIUST IR - ST, + 11 [y + m )AL 3 T SIUR =SOSR, = SN T (g WL (2, )Y (2 )1 (13- ” a a a b where the superscript dagger (T) on a symbol signifies the value of that quantity for a random dis- tribution of ions so that Y1 = Ny- The most probable distribution is obtained by maximizing QS' under the condition of constant total energy and constant number of ions involved and is given by Y NX -~ NA Y _ , 111.3.3 1-Y ]-—NA-NX+NAY B ( ) where 8 = exp (~AE/RT), and where the absence of a prime (*) on Y (or QS) signifies the value of that quantity in the most probable distribution. The total energy is —AE.=Zn, YAE ==AH_., (111.3.4) and the total entropy of mixing is ASp=kInQ_ . (111.3.5) 69 The total Helmholtz free energy can be calculated from Egs. (4) and (5). The following equation is obtained for the partial molar free energy by differentiating the Helmholtz free energy ST Rt inn, Ny () (111.3.6) -— - n ! oty Fay ~Hay ATY \7C Ny where ,ui'l; is the chemical potential of AY in its standard state* and zZ . 1-Y YAY {or yAY) = (W) . (111.3.7) Because of the symmetry of the problem, Eq. (7) is valid for all of the components by merely rede- fining Y and AE. In this approximation (as well as the random mixing approximation) the assumption of the non- interference or additivity of pair interactions has been made so that the energy of attachment of an A* or an X~ ion to any X~ or A* ion respectively is always A€ independent of the number of other ions attached to the A* or X~ ions taking part in the attachment. Thus the energy change for the process X Hm+1wn) Am Xn+(m-—n) . A+.\.fiAm+] ) and for AL X T X T = X ) are the same and are independent of the values of m or n. As will be discussed later this places restrictions on the relative values of the successive association constants, The A* and X~ ions associate if A6 <0 and Y > N, and they will be solvated by the B*and Y™ ions if Ae> 0 and Y ! It has been assumed that Auo = ZAE, and a reasonable value of Z = 4 has been used in the calculations, A calculation of Yay from the two approximations is also given in Table 11 along with values measured at the liquidus temperature at 50 mole %. The symmetric approximation (again for Z = 4) Table 11. Calculated and Measured Parameters at 50 Mole % AY-BX LiF-kCI®! LiF-NaC1%4 NaF-KCI®! Ai® (£ ZAE) (keal/mole) 17.1 9.1 8.0 Random mixing approximation (TC (°K) 2150 1140 1310 Yay 7.8 3.2 2.7 Symmetric approximation (Z = 4) T_(°K) 1560 830 730 Yay 4.7 2.8 2.4 Measured temperature (liquidus) (°K) 1045 973 1010 Y oy (from measurements)® 3.2 2.6 1.8 ayAY at liquidus temperature where AY is the alkali fluoride. 71 leads to values of the activity coefficients of LiF and NaF, which are much closer to those de- rived from the measurements than those calculated from the random mixing approximation. The dif- ference between the experimental results and the calculations from the symmetric approximations is small enough so that the correction for long-range interactions proposed by Fgrland®® and men- tioned in section [ll.2 may be large enough to account for the differences. To illustrate this for a particularly favorable case, in Table 12 are given vaiues of YLiF [yL iF(meas)] in LiF-KCI mixtures calculated from the liquidus temperatures®'+%4 using the heats of fusion in Table 1. Also given are values of ¥L;r calculated from the random mixing and the symmetric approximations [yLiF(symm)]' In the last column is given yLiF(symm) 2 2 yLiF(meus) where, in this case, N¢=Ng =Ngey The form of this quantity (A log yLiF) is consistent with the form of the relation given by Férland®® for the correction factor, A log y , y, which is to be added to log Yay in order to account for the influence of interactions of longer range than nearest neighbors when these interactions obey the equations for regular solutions 2 2 RTAlogy, = Nedx,y * NyXa,g * NgNyIN (A, =20 + Ny (A, =201, (111.4.2) The terms A, and A, are related to the deviations from ideality in AX-BX and AY-BY systems re- spectively and Ay y accounts for the same type of long-range interactions as Ay and Ay but in mix- tures containing both X™ and Y™ ions. Similarly A, and Ay are related to the deviations from ide- ality in AX-AY and BX-BY systems respectively and in mixtures containing A" and B* AA,B refers to the same type of interactions as Ay and A, The magnitude of )\A,B and AX,Y are probably Table 12. Activity Coefficients of LiF in the LiF-KC| Quasi-Binary YLiF yLiF(symm) Liquidus Temperature N, .o From Symmetric Random ‘°9m Nyct Measurements Approximation Mixing LiF 1078 0.90 1.10 1.13 1.08 1068 0.80 1.35 1.45 1.38 0.72 1056 0.65 1.98 2.40 2.71 0.68 1053 0.60 2.30 2.93 3.70 0.67 1045 0.50 3.24 4.72 7.84 0.66 1040 0.46 3.77 5.88 11.17 0.67 1028 0.35 6.28 12.02 34.36 0.67 1020 0.30 8.33 17.97 62.44 0.68 1005 0.25 11.44 29.05 123.6 0.71 72 closely related (perhaps a weighted average) to A, and A, and A, and Ay respectively. These parameters are discussed in section II. In the system discussed here (LiF-KCI) the last term in Eq. (2) is probably small. From the last column of Table 12 a value of (’\XY + )\AB) of - 3200 cal/mole is calculated if the last term is neglected. This is reasonable for the interactions of the ions involved (see section II}. This unexpectedly good agreement is probably fortuitous in view of the approximate nature of the equations for regular solutions as applied to molten salts and the agreement may not be as good in other systems. However, further detailed investigations of such systems, especially in reciprocal alkali halide mixtures, would be interesting for comparison with these considerations where the symmetric approximation is used for nearest-neighbor interactions and Eq. {I11.4.2) is used as a correction factor. This correction factor when included in the calcu- lations of the consolute temperatures, T will lead to much more realistic values than are calcu- lated from Eq. (1). It should be borne in mind that neither the symmetric approximation nor any other approxima- tion which contains the implicit or explicit assumption of the additivity of pair interactions can be generally valid for all molten salts and that neither can give better than semiquantitative results, This will be discussed in a later section. If Eq. (111.3.3) is solved for (1 = Y) in terms of (3~ 1), N,, and N, then AI 1-v) -b+ \/b2—4ac - - 2a and for small values of ac/bz, c ac 2a%c? (1=Y)=-— {14+ —+ + o], (111.4.3) b b2 b* where a = N, (B~ 1), b=1 + (NgNy =N, N B =1, e ==N and taking the logarithm of y, . one obtains v Substituting Eq. (3} in (111.3.7) ac 202(.“2 Iny,y=Z1In <]+b—2+ v +...]=ZInb. (111.4.4) The meaning of the symmetric approximation is made ciear by Eq. (4). Since N, and N appear in exactly the same way in b and in the product ac, the interchange of particular numerical values of N, and N will lead to the same value of y, . If N, and N, are variables, then the function YAy is symmetric about the line N, = Ny By expanding the logarithms in Eq. (4), one obtains Z In YAy =—ZNBNX(‘8_ ])+§-[(NBNX)2+2NANBNXNY](B__'|)2 Z[ ) 6N, N )2 3 (B -1)°3 111.4.5 “g(NB“fl +6Ng NN NG + 3NN (NN THB=-1) + o (1IL4.5) 73 The remaining terms are sums of products of (8 -~ 1)? and (NB NX)!’_”(NA NY)", where p > 3 and p>nZ0. If (B~1)is small, then only the first term is important and ~ AE (B~1)=(e~QB/RT _n T _Z 4 .., RT so that Eq. (5) reduces to Eq. (|11.2.5), which was calculated from the random mixing approxima- tion, For small enough values of N, N, or Ng Ny the higher terms in Eq. (5) are small relative to the first, so that Iny,y = —NgNyZ(B=1), (111.4.6) which has the same form of the concentration dependence as Eq. (111.2.5) but does not contain the implication that there is random mixing of the ions, I11.5 The Asymmetric Approximation One of the weaknesses of the symmetric approximation is the assumption of the additivity of pair interactions which means that in dilute solutions, for example, the energy for forming the pair AX from AT and X~ in the solvent BY is the same as forming AX2- from AX and X~ and A2X+ from AX and A*. Measured association constants in dilute aqueous solutions indicate that this is not valid, especially if the central cation is polyvalent.® Thus, any generalization of the theory which includes a description of polyvalent cations and other special interactions must include a correc- tion for the fact observed in aqueous and molten-sait solutions and discussed in a later section that species such as, for example, Cd2C|3+ are not stable in dilute solutions whereas CdCl, is stable in solution. In the theory which follows only monovalent ions are considered for simplicity. However, most of the relations derived for the association constants in dilute solutions apply to systems containing polyvalent ions as well. The approximation given in this section is the asymmetric approximation which accounts for species as AX (CdCl,, CdC|3_, AgCl,~, AgC|32') and neglects ionic groupings as A X (Cd2C|3+, A92C|+). The applicability of this approximation to real systems will depend on the specific nature of the system. The purpose of the approximation is to derive relations which relate the influence of asymmetry of the ionic interactions to the thermodynamic properties of the solu- tion, In the asymmetric approximation'® the anion portion of the lattice is divided into two regions, a and b. Region a contains all anion positions adjacent to one Ation and (Z = 1) B* ions, and re- gion b contains all other anion positions. In a solution dilute enough in A* to neglect positions adjacent to two A" ions, the number of positions in region a is ZnAYL = L, and in region & is n(nx +n,)~L,=Lg. The X~ ions in region a are more stable by the energy Ae. If A€ is nega- tive, the concentration of X~ ions in region a will be greater than in region & or, in other words, there will be an association of AT and X ions. |f X’ is the concentration in ion fraction units of 74 -, . , . . .. : + . , the X ions in region a, then it is also the fraction of positions adjacent to A" ions occupied by X~ ions. The A and X ions associate when X” > N, and are solvated when X“ that the thermodynamic association constants may be evaluated from the derivatives of In y;Y or In y;X by the relations which have been derived under the reasonable assumption that in very dilute solutions all species obey Henry's law. dlny, 9 Iny. AY B X = | -k, (11.6.1) R B X AY R Av=0 R av=0 Rpx=¢ Rgx=0 2 2 * “lny,y 0% Inygy , (8% - K2_2K, K, (111.6.2) IR2 dR. . OR B X Bx 7% Ay Rpvy=p Ravy=0 R 76 2 * 2 * a ]n }/AY 8 In }/BX 2 2T = K2-2K K, (111.6.3) 2 AY/ . IRy AY=0 Ray=0 Rpx=0 Rpgx=0 where Ry = flz‘j/nBY and where n is the number of moles of the solute component ij (AY or BX in this example). The association constants are in mole ratio or mole fraction units which are the most rational units in molten-salt solutions. These relations are not unique for calculating the as- sociation constants and many other derivatives of functions of the activity coefficients may be used. |t should be noted that there is a single limit of the derivatives of the single-valued func- tions In y:j at infinite dilution of all solutes. Therefore these equations define true thermodynamic association constants under conditions where the calculation procedure includes solutions dilute enough so that all species may be reasonably expected to obey Henry’s law. By using Egs. (1), (2), and (3), expressions for association constants have been calculated from the asymmetric and symmetric approximations [Eqs. (111.5.2), (i11.5.6), (111.3.3), and ([11.3.7)] and are given below: Association Constant Asymmetric Approximation Symmetric Approximation K, Z(B-1) Z(B-1) zZ-1 zZ -1 E (e S Z -2 zZ-2 Ky ( ; %B—l) ( ; %B~U K. ()= (222 - n - —(B-1 12 2 2 This table makes the differences between the symmetric or quasi-chemical and the asymmetric ap- proximations clear. In both approximations K, 2K, 3K nK 3 n Z-1 Z-2 Z-n+1" A [[7aN 1 z (111.6.4) 1 — n Z which are the statistical ratios of Adams and Bjerrum. 149 Thus these approximations are shown to be equivalent to the Adams-Bjerrum ratios® in dilute solutions. In the symmetric approximation K, =K, but in the asymmetric approximation the effective value of K., is - ]/2, which is essen- tially equivalent to zero. Although a negative value of an association constant is meaningless thermodynamically, it can be understood in terms of the model. |f all the ions are randomly mixed and the solution is ideal, all the K's are zero. Since in the asymmetric approximation the condition has been introduced that no more than one A* ion be a nearest neighbor to any one X~ ion, then : : , - + oy . there is less than a random number of A" ions in positions near an A*-Xion pair. Thus the effec- tive value of K, must be less than zero. This will occur if the A¥ ions repel each other. 77 It is clear that these two approximations can correspond exactly to real systems only for special cases or in very dilute solutions, where only the first association to form AX is important. However approximate these models are, they are still useful for semiquantitative descriptions of solution behavior. Moreover, as will be shown in the next section, both models lead to a predic- tion of the temperature coefficient of the first association constant, K., for the association of mon- atomic ions which is correct within the experimental precision of measurements which have been made. 1.7 Comparison of Theory with Experiments in Dilute Solutions Measurements of y:\gNO , the activity coefficients of AgNO, in the mixture Ag *' B*, CI7, and NO, ™ dilute in Ag+ and Cl~, have been made using the concentration cell AgNO, ag|2MNOsil Be A, (111.7.A) BNO, B0, where B is an alkali metal ion {or a mixture of alkali metal ions). In Fig. 21 are plotted measured values of —log yA aNO, vs Ry at 385°C at two values of RAgNO3‘ The solvent BNO, in this case is a 50-50 mole % mixture of NaNO ,-KNO,. The activity coefficients decrease with increas- ing concentration of KCl, the decrease being sma[ler the larger the initial concentration of AgNO,. Obviously the concentration dependence of —log VA NO, is very badly approximated by (l11.2.5) or (111.2.6) and the magnitudes of —log VA aNO, would requn'e very improbable values of ZAE or Ay (about —300 kcal/mole). This large discrepancy is undoubtedly related to nonrandom mixing of the ions. A comparison of these measurements with calculations based on the asymmetric and symmet- ric approximations is made in Fig. 21 and shows that the measured concentration dependence of -log y;\gNO corresponds only roughly to these approximations. At low RAgN03 both approxima- tions are essentially the same and at the higher chloride concentrations indicate a lower activity coefficient than is measured. This, probably, stems from the fact that in this system ~1 K2<< 5 )(B—l). N The activity coefficients at the higher concentrations of AgNO, lie between the two approximations Z-1 indicating that - 4 < K, << >(B — 1). The same is true if the solvent is pure NaNO, or KNO, with measurements in KNO, being closer to the asymmetric and in NaNO, to the symmetric approximation, These comparisons indicate that these two approximations, although much more re- alistic than the random mixing approximation, can be, at best, semiquantitative. One reason for *|n these dilute solutions RAgNO3: NAg and RKCI = NCl' 78 UNCLASSIFIED ORNL-LR-DW7S 57004 0.3 ’ 7 =385° Ragno, = 0.2 X107 3. SYMMETRIC OR / ASYMMETRIC APPROXIMATION / 0.2 7 AN > _ -3 2 / Fagno,= 22 *10 > . o 7 /"_ o o ' K 0A //.‘/< /// ' - / ’ SYMME TRIC APPROXIMATION S .///\\ r | i ////.,// ASYMMETRIC APPROXIMATION p § 7 4 0 0 10 2.0 3.0 Rey X 103 Fig. 21. Comparison of the Concentration Dependence of Measured Values of —log ngNO in NaNO4.KNO, (50-50 Meole %) Mixtures with Theoretical Calculations Based on the Symmetric and Asymmetric Approximations. this is the nonadditivity of pair bond interaction energies in dilute solutions. This means that the relative values of successive association constants do not, in general, correspond to the values given on page 76. In the next section a generalization of theory will be made which will include the possibility of the nonadditivity of pair bond interactions. The theoretical evaluation of K, however, is meaningful for certain systems and in solutions dilute enough in A" and X~ so that the most important species is AX, the temperature dependence of the activity coefficients (and of K]) is predicted by these two approximations. Measurements of —log V;«gNOB in dilute solutions of Ag* and CI™ inthe three solvents Nc1N03,70 KNO3,23'94 and 50-50 mole % N0N03-KNO369 mixtures were compared to theory, By comparing the approximation - . . * which was closest in concentration dependence to the measured vaive of ~log Yagno, ot low con- 3 centrations of Ag' and Cl~ values of K, could be evaluated and are given in Table 13. This pro- cedure for evaluating K| has been shown to be equivalent to more conventional extrapolation pro- 79 Table 13. Values of AE.l Obtained from the Comparison of Theory with Experimental Dato —AE.I {kcal) T (OK) K] =Z(B_'|)(a) Asymmetric Approximation, Ag+, K+, Cl-, N03- 623 6.12 5.85 5.62 553 643 6.17 5.89 5.66 498 658 6.21 5.93 5.69 460 675 6.17 5.87 5.64 396 696 6.18 5.88 5.63 348 709 6.17 5.86 5.62 315 Symmetric Approximation, Ag+, No+, cl, NO3- 604 5.10 4.83 4.62 277 637 5.12 4.84 4,62 226 658 5.17 4.88 4.65 205 675 5.10 4.81 4.57 176 696 5.13 4.83 4.59 160 711 5.12 4.81 4.56 146 773 5.14 4.82 4.55 110 Asymmetric Approximation, Ag+, (Nu+, K+), Cl-, N03- 506 5.6 5.4 5.2 1050 551 5.57 5.33 5.13 644 658 5.67 5.38 5.15 302 752 5.72 5.40 5.13 180 801 5.6, 5.24 5.04 133 (“)K] in mole fraction units. cedures?® if used correctly. To evaluate the parameter AE(AE ) contained in the theoretical ex- pression for K, a value of Z must be assumed where K,=Z(B~-1)=Zlexp (-AE,/RT) - 1], (11.7.1) In molten salts a range of values of Z which covers all reasonable possibilities is 4 to 6. In Table 13 are given values of AE, calculated for values of Z = 4, 5, and 6. |n any one system and for any one value of Z the values of AE, thus calculated, within the estimated experimental error, do not vary with temperature. This means that Eq. (1) correctly predicts the temperature coefficient of K in these systems. In the NaNO ,-KNO, system this prediction is correct over a range of 295°C and 80 for a variation of K, by a factor of about 8. At low enough concentrations of Ag*and CI” so that the only important species is AgCl the variation of —log y‘AgN03 with temperature, within the ex- perimental precision, is also correctly predicted. This is illustrated in Fig. 22 which gives a plot of —log y:\gNoa in a dilute solution of Ag” and Cl ™ in NaNO, at several temperatures. The dashed lines were calculated from the symmetric approximation using the parameters given in Table 13, Using these essentially constant values of AE, leads to an excellent correspondence of the calcu- lated and measured valuves of —log y:\gNO at low concentrations of Ag* and CI~, 3 UNCLASSIFIED ORNL-LR-DWG 44359 0.4 7=331°C 4 364°C R, =0.30x10"° AgNO3z ™ 7 ------ THEORETICAL CALCULATIONS j 0.3 ‘ 402°c O 2 202 ™ 2 - g ‘ { 500°C 7, 04 0 2 3 4 (x10™3) o 1 RNaCI Fig. 22. Comparison of the Temperature Dependence of —log y;\gNO in 3 Nr.:NO3 with Theoretical Calculations, 81 The theory, in essence, leads to a prediction of the ‘‘configurational’’ contribution to the en- tropy of association so that from measurements at one temperature one may also calculate the heat of association, AH ]': dinK, -AH! dlnZ(B-1) = = ] . (11.7.2) ) R d( % d( ‘/T Since AE, is independent of temperature, AH? = AE, (B_B_.]_) (111.7.3) where it is to be remembered that AE, can be calculated from measurements at a single tempera- ture. Equation (3) for AH ] may be confirmed {within the experimental precision and within the range of values of AE, for the three values of Z) from the values of K, given in Table 13. Other reported values of AH T which differ from Eq. (3)38 were calculated from too few points and over too short a range of temperatures to be significant. I11.8 Generalized Quasi-Lattice Calculations'? The comparison of both the symmetric and asymmetric approximations with experiments make it evident that less stringent restrictions on the relative energies of association are necessary for a comprehensive theory. In this section a generalized calculation based on the quasi-lattice model will be discussed. The purpose is to calculate more general expressions for some of the higher association constants. For simplicity, the assumption is made, as in the asymmetric model, that the solution in the sol- vent BY is so dilute in A jons that one can neglect all groupings of A" and X ions containing more than one A jon. From a calculation of the partition function for the assembly of AY, B X, and Y ions calculations were made of the Helmholtz free energy, the chemical potential for the component AY, and, hence, the activity coefficients of AY, y:\Y, in terms of the ion fractions of the ions, Z and 3., where 3. = exp (~AA /RT) and AA, is the “*specific bond strength’’ or the ‘‘spe- cific Helmholtz free energy change’’ for the association ([FaN AX{Z=D | X —a AX{1 ), 15iS7. (I1.8.A) i In this approximation AA, # Ad, £ AA, # AA;in general. |t must be kept in mind that the symbol AXS.]"i) represents an A" ion having i X ions and (Z — i) Y~ ions as nearest neighbors. Thus (AAl./n) is the free energy change for exchanging one X~ ion in the body of solution with a par- ticular Y ion adjacent to the A” ion in the grouping AXSET.). The term AA . is related to partition functions for the individual ions involved in the association (A) (which is really an exchange of ions) so that AA;=~RT In| —r], (111.8.1) 82 where 774/ and Wq;; represent the product of the partition functions of the individual ions, &, in- volved in the association process (A) evaluated before and after the association process respec- tively. If the partition functions are separable so that - =E,/RT Thi = 9ri® ’ where g, represents a partition function for the internal degrees of freedom of the ion of type %, then AA; = AE, - TAS, = (RE/ = XE/.) = RT In , {111.8.2) z 1 i z ki N T4 and the “*specific'’ entropy term, AS., contains only contributions from the internal degrees of free- dom and excludes statistical or combinatory factors for the groupings of ions. For negligible ~o changes in the internal degrees of freedom of the ions involved in the association process AS. =0 and (dAA./dT) = 0. This is the case for the values of AA | in the systems cited in Table 13, The statistical mechanical calculation'* leads to the equations for some of the successive as- sociation constants {in mole fraction units) (111.8.3) Ky=2Z(B;~1, (a) K1K2=-Z—(%:-]—)(Blfiz—2fil+1), (b) K1K12=Z_(Z;!;])‘(181512-251+]), (c) K1K2K3=Eg_—;.i(z-—;a(818283—38182+3,8]—1), (d) Z(Z - 1NZ-2)(Z - 3) 4! Ky KoKgKy = (B1B2B3By =481 By B3+ 68,8, =48, +1). (e Equation (111.8.3a) is the same as the expression for K, given in the table on page 76 if AA | = AE |. The terms in Z are spatial and statistical factors and the terms in B; are related to the bond energies, For the case in which AAy=AA,=AA;=AA and B, =3, = B3 = B, the statistical ratios of Adams and Bjerrum apply. Some of the relations derived from (111.8.3) exhibit surprising properties. For example by di- viding (111.8.3b) by (111.8.3a) one obtains the expression B,-B K, =[(z~1)/2] {(32-1“(%) . (111.8.4) It can be seen from (111.8.4) that K, depends not only on Z and B, butalsoon B,. If B, is small, this dependence may be relatively significant. If, for example (2 - ]/B]) > B, > 1, then there ex- 83 ists a tendency for the association* of AX with X to form AX2_ and yet the values of K, may ap- pear to be negative.** This unusual and apparently contradictory result arises because of the re- quirement that the conventional association constants, K., be almost zero in an almost ideal solu- tion. The standard states for some of the associated species under this requirement which is inherent in the commonly accepted methods of describing associations in solution cannot be under- stood in g simple way and |lead to unusual properties for weak associations. An analogous situa- tion occurs when gas virial coefficients are interpreted in terms of clusters,” /74 The assumptions made in the calculation of Eq. (3) are that the ith X ion attaching itself to an A% ion can do so in (Z — i + 1) equivalent positions. Different relations would be obtained under different assumptions. If, for example, only a linear AXz- ion triplet can form, the second X~ ion has a nonzero value of A4, in only one of the (Z — 1) sites near an AX ion pair which is not already occupied by an X™ ion, For linear AX2— then Ky Ky = [z(z = 0/20{18, B/(Z = W] + 1 = [2/(Z - VIR, } (111.8.5) and BQ"1 B]"'] i 1 Ky=%|By—-Z + The stepwise association constant for formation of linear AX2" is Ké and would be smaller than K, for a nonlinear grouping even with the same values of Z, 8,, and B, Thus the comparison of the Eqs. (3b) and (5) demonstrates in this simple case the general principle that the greater the tendency toward ''directionality’’ in a ''bond'’ the lower will be the association constant, if all other factors are equal, Equations (5) and (3a) lead to conclusions differing from those of Bierrum7 on the ratios of successive association constants for linear AX,". For values of 8, = 8, >> 1 for example, KI/Klz 2 27, where Z is a maximum coordination number, In Bjerrum's derivation this number is a characteristic coordination number N, For a common case in which 8, << 8,, N is two and much smaller than Z. The error in the calculation of Bjerrum arises from the fact that when the total pos- sible number of X ligands is restricted to N in his derivation, the total number of positions adja- cent to a spherical A¥ ion which are available to the first ligand is simultaneously limited to N al- though the first ligand is actually able to attach itself in any one of Z positions. The equations discussed in this section can be derived for nearest-neighbor interactions inde- pendently of the lattice model. The coordinator number Z in such a derivation would be the ratio of the volume of the first coordination shells adjacent to a mole of A* ions to the volume of a mole of solvent anions. Such a derivation would apply to polyvalent cations, *f AA:' is negative and Bi > 1, there will be a tendency toward the association of AXS._z_-l'i) and X~ to form AXf.]-'i). **Negative values of K, are meaningless thermodynamically, and apparently negative values usually mean a repulsion of the ions involved rather than the assumed association, 84 II1.9 Association Constants in Dilute Solutions In this section a compilation is given of association constants (in mole fraction units) which have been measured in reciprocal molten-salt systems, Measurements of associations invelving the Ag* ion have been largely made with cells of the type (I11.7.A) and the most reliable measurements for associations involving T1¥, Pb2*, and Cd?* with halides from cells of the type A(NO ) AgX(solid) n Aq |BX AgX(solid)| | (111.9.A) g g . .7, BX BNO, BNO, using silver-solid-silver halide (AgX) electrodes where At s T|+, F’b2+, or Cd2*. The emf of cells (I11.7.A) and (111.9.A) may be related to the activity and activity coefficients (y*) of AgNO, or BX respectively. To avoid confusion, it should be emphasized that these activity coefficients are defined so as to encompass all solution effects including ionic associations (‘‘complex ion'’ formation). At concentrations where Henry's law is obeyed by all species (probably true at ¢on- centrations below 0.5 mole %) it represents only those deviations from ideal solution behavior which are caused by association in solution, This usage is simpler than the usage most often em- ployed in aqueous solutions where deviations from ideal solution behavior are subdivided into “*ac- tivity coefficient’’ effects (related to the ionic strength) and an effect due to associations. Ther- modynamic association constants may be computed from these measured activity coefficients by an extrapolation method.?% Some of the association constants cited here have been recalculated from the data in the literature.2* In cases where errors in calculating association constants may be sig- nificantly larger than the errors stated by the original workers and not enocugh data were available to correct the calculations, the association constants are given in parentheses or omitted, From the tabulated association constants (Tables 14 and 15), values of A4, were calculated from Egs. (111.8.3) for Z = 6 and are given in Table 16 for monatomic ions. For other values of Z, AA. would be somewhat different {for Z = 4 the AAz. would be more negative by about 0.4 to 0.6 kcal) but the differences between the different values would be about the same. The differences in AAin Table 16 are related to the association constants (for K;>> 1) by AAI.'—- AAiH Z_RT In Kz.'/KI.". In every case where measurements were made at more than one temperature for associations in- volving monatomic ions only, values of AA, for a given association in a given solvent and for Z = 4, 5, or 6 were independent of temperature within the experimental uncerfuinties.* Thus it appears that, for monatomic ions, the temperature variations of K, and K, as well as of K, may be pre- dicted from Eqgs. (111.8.3) by using constant values of AA, and it appears that the entropy of as- sociation is largely the *‘configurational’’ entropy calculated from the quasi-lattice model. For *There did appear to be trends in the variation of AAI. with temperature in some cases. The total varia- . . . . + tions were smaller than the experimental errors in all coses except for AAI for the formation of CdBr in 50- 50 mole % NaNOa-KNoa, where the variation of AA'I was slightly larger than the estimated experimental errors, 85 Table 14. Campilation of Associatian Constants from EMF Measurements (see also Table 13) Associating T (°K) Solvent lons K] K2 Kl2 References 675 NaNO, Agt 4B 633 246 280 95 711 500 180 200 733 430 151 167 773 325 103 120 606 NaNO ,-KNO , (53-47 mole %) Agt+cl” 381 145 38 647 302 97 649 Agt+Be” 1,008 (360) 38 687 781 (199) 528 Pb2t s Br” 199 39 576 153 579 67 529 cd?t i ee” 1,170 550 39 547 1,030 510 571 810 513 NaNO ,-KNO 4 (50-50 mole %) TiteBe” 31 15 27 519 Agt+CN™ 230,000 140,000 80,000 10 559 220,000 105,000 60,000 93 599 190,000 50,000 36,000 513 cd?t 1B 1,520 680 ~0 25 573 990 450 ~Q 513 cd?t 4 ” 5,330 2,200 ~Q 25 563 3,130 1,300 ~Q 513 Pb2* 4 Br” 250 125 ~Q 27 573 170 85 ~Q 92 623 KNO, Agt+cI” 553 215 <40 94 658 460 169 20 709 315 17 <40 676 Agt 4B 932 370 293 2 711 768 285 230 725 728 273 208 747 617 228 174 773 540 195 145 675 Agt+1” 5,420 2,700 3,555 2 636 Ag+4-5042_ 11.6 132 681 12.1 706 12.7 722 13.3 513 LiNO 4-KNO 5 (80-20 mole %) cd?t 4B 4,300 1,700 26 513 (6535 mole %) 3,600 1,600 444 (50-50 mole %) 7,500 3,300 513 3,000 1,300 513 (40-60 mole %) 2,500 1,100 513 (26-74 mole %) 2,300 1,000 553 (40-60 mole %) T1 4 Be” 56 30 27 86 Table 15. Association Constants from Other Measurements Solvent T (°K) Species K] Method 129 + + Nc:xNO3 580 CdCl 190 =50 Cryoscopy PbCIt 60 +20 Cryoscopy LiN03-KN0335 (50-50 mole %) 453 cdcit (900)% Polarography Pocl? 270 +80 Polarography NaNo3-|rB rx>rY + rArY + r,>r r,