Creator:Henry Eyring and Burno J. Zwolinski Date Created:October 1947 Place Created: Keywords:Henry Eyring,reaction rate theory Context:article reprinted from the Record of Chemical Progress ************************************************** The Foundations of Reaction Rate Theory and Some Recent Applications Henry Eyring and Bruno J. Zwolinski Henry Eyring was born in Chihuahua, Alexico, in 1901. He obtained degrees in Mining Engineering and Metallurgical Engineering in 1923—24 from the University of Arizona. In 1924-25 he was an instructor in chemistry at the University of Arizona. He then attended the University of California at Berkeley, where he received a Ph.D. in chemistry in 1927. At the University of Wisconsin he was instructor of chemistry from 1927-28 and research associate from 1928-29. In 1929-30 he attended the Kaiser-Wilhelm Institute in Berlin as a National Research Fellow. He then lectured in chemistry at the University of California. In 1931 he joined the Department of Chemistry at Princeton University, where he remained until August, 1946. From 1944 to 1946 he was Director of the Textile Foundation. During World War II he did research work for the Army, Navy and OSRD on the theory of smokes and the theory of high explosives. In September, 1946, he became Dean of the Graduate School and Professor of Chemistry at the University of Utah. Dr. Eyring's fields of specialization include radioactivity, the application of quantum mechanics to chemistry, the theory of reaction rates, and the theory of liquids. He has published about 150 papers and various chapters in chemistry books, and is co-author with Glasstone and Laidler of "The Theory of Rate Processes" (1941) and with Walter and Kimball of "Quantum Chemistry" (1944). Dr. Eyring has held numerous offices in national scientific societies. He is a member of the National Academy of Sciences, the New York Academy of Sciences, the American Philosophical Society, The American Association for the Advancement of Science, the Textile Research Institute, Sigma Xi, and the Utah Academy of Sciences, Arts and Letters. j Reprinted from the July—October 1947 Issue of the Record of Chemical Progress The Foundations of Reaction Rate Theory and Some Recent Applications HENRY EYRING, University of Utah, and BRUNO J. ZWOLINSKI,* Princeton University Though philosophers for centuries were conscious of the importance of the universal variable time in the interpretation of nature, it was barely a century ago that this factor was properly taken into consideration in a series of physical measurements on a reacting system, thus marking the birth of the field of chemicalkinetics. Fortyyearspassed by without any very significant contributions to the development of kinetics. This period came to a close in 1889, when Arrhenius (1) laid the foundation for the collision theory of chemical reactions, and set the scene for the modern development of rate theory. He proposed that an equilibrium existed between normal and activated molecules, and the variation of the specific rate of reaction with temperature could be expressed by the formula k = Ae-E/RT (i) This equation correctly represents the temperature dependence not only of most chemical reactions, but also of certain physical processes. McC. Lewis (2), Hinshelwood (3) and oth ers set the constant A i n Arrhenius' reaction isochore for bimolecu-lar reactions equal to the number of collisions. This served as an excellent guide in correlating and furthering our knowledge of chemical changes. Wherever it was found possible to consider the reacting molecules as hard spherical particles the collision theory admirably interpreted the experimental data. As * Allied Chemical and Dye Corporation Research Fellow. further data were accumulated on reacting systems involving more complex molecules it was found necessary to introduce empirical probability or steric factors into the collision-theory expressions for the specific rate of reaction to obtain agreement with experiment. All efforts expended on improving the collision theory only further emphasized its limitations and the need that existed for a new approach to the problem of reaction rates. A new attack on the problem of the calculation of rates of chemical reactions from first principles utilizing the fundamental properties and molecular constants of the reacting molecules was afforded by the development of quantum mechanics. In 1928, F. London (4) indicated how the methods of quantum mechanics could be employed in calculating the energy of activation of chemical reactions which are termed "adia-batic," in the sense that they do not involve electronic transitions and occur on the same potential energy surface". He developed an equation based on certain approximations showing how the potential energy of a system of three or four atoms with s electrons varied with interatomic distances. This theoretical treatment of the fundamental quantity, the energy of activation, provided new impetus to an attack on the rates of chemical changes. The usefulness of this equation in constructing potential energy surfaces for the interpretation of chemical reactions was developed and extended by H. Eyring and M. Polanyi (5). Since 87 88 Record of Chemical Progress the London equation even in its approximate form inyolves quantities which can be completely evaluated only for the simplest system consisting of hydrogen atoms, they considered approximate solutions which led to the "semi-empirical" method for the calculation of the energy of activation. Fundamental contributions came from many sources on the calculation of the absolute reaction rates which are summarized in a recent text (6). A general formulation of the theory of absolute rates of reaction dealing with the calculation of the frequency factor and applicable to any rate process was given by H. Eyring in 1935 (7). A similar treatment was presented by M. Polanyi and M. G. Evans (8) which is sometimes referred to as the transition state method. Both extended an earlier paper of Pelzer and Wigner (9). Briefly, we present the essential points and assumptions of the theory of absolute rates of reaction. A chemical reaction and many physical changes with time are characterized by an initial configuration which by a continuous change of coordinates passes into a final configuration. For each process, there is an intermediate or critical configuration called the "activated complex" or "transition state," situated at the highest point of the most favorable reaction path on the potential energy surface. If a molecule reaches this critical point in a certain region of phase space, there is a high probability that a reaction will occur. The activated configuration is like an ordinary molecule with the usual thermodynamic properties except that it possesses an extra degree of translational freedom along the direction of the reaction coordinate. By assuming an equilib- rium to exist between the initial and the activated state, the specific rate of a reaction can be determined by calculating the concentration of the activated complexes and their rate of passage across the potential barrier or saddle point of the potential energy surface, by using statistical methods. Consider a process in which the rate is determined by the passage over a potential barrier and quantum-mechanical tunneling effects can be disregarded, then v Rate of reaction v = kC (2) where C'* is the concentration of activated complexes per unit volume and the ratio v/S is the mean velocity of crossing in one direction divided by a length S of the activated state. Thus v/S represents the frequency of emptying the length of path 8 of activated complexes. The factor k, the transmission coefficient, is the factor introduced to take care of the possibility that not all the activated complexes reach the final state. For most reactions it is sensibly equal to unity. Treating the activated complexes as normal molecules by replacing the metastable degrees of vibrational freedom by translational motion along the reaction coordinate, we find In addition, the mean velocity v — (kT/2rm^y/\ so that Eq. (2) becomes hT v = kC± J-. (3) Accepting the hypothesis of an equilibrium between the initial re-actants A and B and the activated state, we have for the velocity of a bimolecular reaction July-October, 1947 89 KCaCB^-K*, n and in general we obtain for the specific reaction rate of any order: kT = (4) Applying a thermodynamic formulation, by relating the constant K* to AF*. AH" and AS*, the standard free-energy, heat-content and entropy changes for formation of the activated state, we may express Eq. (4) k' = K^e-a f*/rt n kT = K^-e^ S*>/Re-*.H-*/RT. (5) In terms of partition functions, Eq. (4) reads kT k' = KTFjrBe~Eo/RT' (6) where E0 is the energy of activation at the absolute zero temperature. If we are concerned with non-conservative systems, the Eq. (6) is readily extended to include the effect of an applied external force. Consider an applied force to have a component/ along the reaction path. If this force acts across a symmetrical barrier through a distance X/2 from the normal to the activated state, the free energy of activation Af* in the absence of the external force will be decreased by an amount (X/2). The specific rate of the forward reaction in the presence of the applied force will be = fctfA/2 kT, (7) where ko is the specific rate in the absence of the external force and N is Avogadro's number. Similarly for the reverse process, we have kb' = feoe-A/2 kT, (8) giving for the net rate k' = k,' - k„' = efV2*r _ e-f\/2kT) = 2*o sin h^ff. (9) A wide variety of physical processes including viscosity, plasticity, diffusion, electrochemical phenomena, creep in metals and high polymeric substances are readily interpreted by modifications of Eq. (9) (6). The Non-Equilibrium Theory of Absolute Rates of Reaction In the above formulation of a rate theory which is applicable to any process in which the rearrangement of matter involves surmounting a potential barrier, the tunneling effect was disregarded. Though for the majority of chemical reactions which do not involve a transfer of electrons this effect is negligible, we have to take the barrier penetration into consideration when dealing with the decomposition of N20 or the inversion of ammonia (10). If the potential barrier is nearly flat, the velocity of reaction is corrected for penetration or tunneling effects by multiplying the specific rate constant by the factor due to Wigner (11), i1 ~2iCfi)2)' Where is the imaginary value of the stretching vibration along the coordinate of decomposition. The leakage effect is generally small and can usually be neglected without serious error. The other point which requires further consideration is the appearance of the probability factor g—Eo/RT jn t}je rate expressions, where E0 is the difference in the residual or zero-point energies of the 90 Record of Chemical Progress initial and the activated state. It arises from our basic assumption that an equilibrium exists between the normal and the activated states which is not disturbed to any extent during the course of the rate process. The validity of this assumption will now be examined as tantamount to a complete formulation of the theory of absolute rates of reaction. The applicability of equilibrium theory to chemical reactions was first suggested by Arrhenius. This hypothesis was discussed by Mar-celin (12), who expressed the opinion that for measurable reactions occurring under normal conditions of temperature and pressure the assumption is correct. Further discussions of this matter were given by Wynne-Jones and Eyring (13) and also by Guggenheim and Weiss (14). The equilibrium postulate underlies every proposed theory of reaction rates. The success of the crude collision theory and more so of the theory of absolute rates, which closely interprets many diversified physical phenomena, is strong evidence for the correctness of the equilibrium hypothesis. All this, however, is a posteriori evidence that an adequate supply of energetic or activated. molecules is maintained during all stages of the reaction, and more direct quantitative reasoning is desired. Our present knowledge of chemical dynamics is not sufficiently advanced to investigate in detail the individual collision processes which give rise to an infinite variety of energetic molecules and to determine what fraction of these fortuitously propitious collisions determines the concentration of the activated state. An ingenious approach to this complex problem was made by H. A. Kramers (15). To elucidate the applicability of the absolute rate theory for calculating the velocity of chemical reactions he considered the effect of Brownian motion on the probability of escape of a particle (caught in a potential well) over a potential barrier. Kramers' results indicated that the theory of absolute rates of reaction gives results correct within 10% over a wide range of viscosity values. In his method of investigation the classical mechanical diffusion theory was employed, so it is of interest to take into account the quantized nalure of molecular levels. This has been done in a recent investigation (16) of the non-equilibrium theory of absolute rates of reaction, wherein results were obtained essentially in agreement with Kramers' calculations. Consider reactants passing by a series of propitious molecular collisions from a set of energy levels to a subsequent set of levels corresponding to the final states of the products. It will be assumed that values for the specific rates of transition kv from level i to level j are known, which in principle at least are calculable from the quantum-mechanical theory of collisions (17). Restricting ourselves to reactions in which we can neglect the concentration changes in all species except A and designating the number in the ith level by Au we have the following set of n rate equations M - knAi) (10) jVi corresponding to the n possible energy levels of the reactants and products taken as one. Degenerate levels, like other levels, each carry a separate subscript. In each equation of the set (10), the summation July-October, 1947 91 extends over all n values except i. The solution is readily obtained for the set of linear differential equations with constant coefficients. Let us try the solutions At = Bie"' where Bt is a constant, b is the characteristic parameter, and t is the time. Substituting in Eq. (10), we obtain the simultaneous set of homogeneous algebraio equations £ [knBi -(*,-, + b)Bt] - 0. (11) M' Solving for the Bt's, we have specifically for the constant Bu 0 .................k„ B 1 = 0 ki„ki, + 6) 'Vj '_ ........k„> Mj klnkln -£(*„< + b) Mj (12) Aside from the trivial solution Bi = Bi = ... = B„ = 0, a solution for non-zero values of the Bt's exists only if the characteristic determi-ant, that is, the denominator of (12), is set equal to zero. With this equal to zero, the n values of the characteristic parameter b may be found by substituting values for the transition constants ky and solving the nth order determinant. By substituting the determined value of the kth root b/c into Eq. (11) in the usual way, one is led to the solutions Bjk = GkCj/c, where for each j and k one obtains a numerical value for Cjk, and Gk is the same arbitrary constant for all j's. Summing all the particular integrals, the general solution of (10) is Ai = £ Bikeh> = £ GkCilceh' (13) In (13) the values of Clk and hk are known, and by putting t = 0 and the j4«'s equal to their initial concentrations, the arbitrary constants are readily calculated. Each At thus becomes a completely determined function of time. The applicability of the general procedure outlined above will be shown in a specific case. The following simple model was chosen in which the initial state was considered to consist of levels 1 and 2 and the final state of levels 3 and 4, with 4 designated as the level of lowest energy. This case corresponds to n = 4 in the general expressions given above. The rate equations are = — jk,jAi + ki,A, + al Ml d4r = knA, - + al M 2 knA3 + k,i Ak k32A3 + k,2A, dA, dt dAt dt = knA\ + kraAl — y^kijAz + knAi M3 = k]tAi + kuAi + kuAl — 4 jA,. M4 (14) Proceeding according to the general method outlined for the case of n levels, the particular solutions are assumed to take the form Ai = Biebt (i=l,..., 4). (15) To make possible a solution of the characteristic 4th order determinant for the parameter b and, subsequently, to obtain values for the 92 constants Bi, ..., Bit certain assumptions have to be made with reference to the n(n—1) = 12 reaction rate constants and proper values chosen to represent their magnitudes. If we assume that a molecule in the activated state has the same probability for decomposition along the reaction coordinate to any level of the final state, then for the forward process kvi = ku and = fc>4. Similarly for the reverse process, ku = ki2 and k32 = k31. If in addition we limit ourselves to reacting systems of small heats of reaction, so that the energies of the respective molecular levels in the initial and final state are approximately the same, and combine the resulting relations between the ky's with those based on the first assumption, the following set of relations is found to exist between the twelve specific reaction rate constants of the system: klZ = kit = kil = kit kn = k32 = k-u = ku ki2 = k< >, ku r= ku The second assumption implies further that the transmission coefficients of the specific rate constants with reference to a similar pair of transitions, as, for example, or k32 or ku and are approximately equal. On the basis of the above assumptions for the fe/s, the characteristic 4th order determinant is readily diagonalized to yield expressions for the parameter b in terms of the four determining rate constants &12, &13, k2i and k23. Choosing the following plausible values for the ki/s, k12 = 0.01 = 0.1 ka = 1 ku = 0.001 the quantities bk and Cjk have been calculated and are given in Table 1. With the particular integrals of Eq. Record of Chemical Progress i (15) fully determined, it is only Table 1 k Cih C24 Cn Cik 0 1 0.01 0.01 1 -1.306 -1 1.437 -1.437 1 -1.111 1 -1 -1 1 -6.737 -6.959 6.959 x 10-3 -1 x 10-' x 10- "U necessary to specify the supplementary conditions when t = 0, and evaluate the arbitrary constants Gi, ..., Gi to permit us to obtain the explicit expressions as to how the population in the four levels varies with time of reaction. We will then be able to compare the'rates of reaction under equilibrium and non-equilibrium conditions. Introducing the arbitrary constants, the general solution of our simple four-level system is Ai = Gie6'' - G2eM + Gte<"' -G^'' A2 = Gi X 10 "V1' + 1.437G2e62' - G3e63' - 6.959G4 X 10 "V"' A3 = Gi X 10 -yi< - 1.437G2e^", - G3eb>' + 6.595G4 X I0~3eb>' A, = Gie6l< + G2eM + G3eb* +Gie'"' (16) where: 61 = 0, b2= -1.306, b3 = -1.111, 64= —6.737 X 10-3. By specifying the supplementary initial conditions (i. e., when t= 0) for the cases of equilibrium and non-equi-librium, the arbitrary constants of the respective general solutions are calculable. Under equilibrium conditions, the equilibrium concentrations of the various species At defined as Hi, • • •, Tii are "> — 4 VJ-W e-*i/kT j= 1 where D is the total concentration of the molecules A in the system and July-October, 1947 93 the statistical weights wt for all levels are taken to be equal. From the principle of detailed balance, we may write that ktj = ktie~''i/kT where ejt = e(, so that Eq. (17) is equally well written: m = -JL- (,•_ l, ...,4). (18) v kii This represents the supplementary conditions for the equilibrium case. For the non-equilibrium case, it is assumed that the concentration in the lowest level of the initial state (level 1) is at its equilibrium value and the concentrations of the species A in all the remaining levels of the initial and final states are zero; thus the supplementary conditions for the chosen non-equilibrium case when < = 0 are that Ai = D/2.02 .and Ai=A3=Ai=0. The value for Ai was obtained by substituting the chosen values for the rate constants ktj into Eq. (18), the expression for the equilibrium concentration. The respective arbitrary constants for the two cases considered have been calculated and are summarized in Table 2. Substituting these , Table 2 Gk Equilibrium Non-Equilibrium G, D/2.02 0.2451D G, 0 -1.193 X 10-3£> G, 0 0.2451 X 10-2D Gt 0 -0.2463D values into the general expression given in Eq. (16), the general solution for the case of equilibrium is simply A' = 2MX10"D A. = rt 4 2.02 ' (19) where the concentrations in the various levels are constant. Proceeding similarly for the case of non-equilibrium, the general solution is, Ai = Z)[0.2451e!,»' + 1.193 X 10-V'< + 2.451 X 10-V'1 + 0.2463eM] Ai = DX 10-2[0.2451eM - 0.1714e<"< - 0.2451ei"' + 0.1714e6il] A3 = D X 10_2[0.2451e6lt + 0.1714ei»' - 0.245le1'' - O.mie6'1] At = D[0.2451e6'1 - 1.193 X 10"V"' + 2.451 X 10-V'1 - 0.2463e'<<] (20) where: 6i = 0, b2 = -1.306, b3 = -1.111, 64=-6.737 X 10-3. To test the Soundness of the equilibrium postulate of the activated complex theory of rate processes, it is required to show to what extent the equilibrium between initial levels is disturbed, as molecules from the activated state rearrange or decompose into the products of the final state. This is best demonstrated by formulating an expression for the ratio of the actual rate to the equilibrium rate for the process, and calculating the variation of this ratio r with the amount of the substance that has reacted. The effect of transitions between levels of the initial state is taken into consideration in our expressions for the concentrations Ai and A2 as given by Eq. (20). The actual velocity of the forward reaction is then Va = (ki, + ku)Ai + (fea + ku)Ai. (21) To calculate the rate of reaction by the theory of absolute rates of reaction, the basic assumption is made that the population of the levels in the initial state for the reacting species A is determined by the 94 Record of Chemical Progress Maxwell-Boltzmann distribution function. Referring to this velocity as the "equilibrium" rate ve for the forward process, we have e-ti/kT v, = (Rl3 + ku) (Ai + A2) 2--1- ^ e—ei/kT 8 = 1 p — ti/kT (k2, +kidiAi + At) ^-, (22) Y^e-'i/kT i = 1 where (Ai+yl2) is the total concentration of the substance /I in the initial state. Hence, from Eqs.. (21) and (22) the ratio F of the actual rate to the equilibrium rate is Maxwell-Boltzmann probability factors, we arrive at the expression r = iVi + N&m/hT ew/kT I e-ea/kT^ 1 -)- eea/kT 7>[iVi + N2e= In p<)rn'/«= (VlMf)'") -(v'i1" <''>'"'- <44) In these equations Z is taken as the number of atoms in the chain and 5 the number of atoms in the segment. Reasonable values for pt and 5 give the values for a found by Flory (25) and Eyring and Powell (26). From the above we see that probably the equilibrium theory of reaction rates is seldom in appreciable error. A theory of the ignition temperature of combustibles has been given, and finally we have given a framework for discussing the flow of large molecules which is capable of extension and wide application. REFERENCES (1) S. Arrhenius, Z. physik. Chem., U, 226 (1889). (2) W. C. McC. Lewis, J. Chem. Soc., 113, 471 (1918). (3) C. N. Hinshelwood, "Kinetics of Chemical Change," Oxford University Press, 1940. 102 Record of Chemical Progress (4) F. London, Z. Electrochem., 35, 552 (1929). (5) H. Eyring and M. Polanyi, Z. physik. Chem., B, 12, 279 (1931). (6) S. Glasstone, K. Laidler and H. Eyring, "The Theory of Rate Processes," McGraw-Hill, New York, 19U1, p. 11. (7) H. Eyring, J. Chem. Phys., 3, 107 (1935). (8) M. Polanyi and M. G. Evans, Trans. Faraday Soc., 31, 875 (1935). (9) H. Pelzer and E. Wigner, Z. physik. Chem., B, 15, 445 (1932). (10) A. Stearn and H. Eyring, J. Chem. Phys., 5, 113 (1937). (11) E. Wigner, Z. physik. Chem., B. 19, 203 (1932). (12) A. Marcelin, Ann. Physique, 3, 158 (1915). (13) W. F. K. Wynne-Jones and H. Eyring, J. Chem. Phys., 3, 493 (1935). (14) E. A. Guggenheim and J. Weiss, Trans. Faraday Soc., 34, 57 (1938). (15) H. A. Kramers, Physica, 7,284-304 (1940). (16) Presented by the present authors at the 111th national meeting of the American Chemical Society, April, 1947. (17) N. F. Mott and H. S. W. Massey, "Theory of Atomic Collisions," Oxford University Press, 1933. (18) T. E. Leontis and F. N. Rhines, Metals Technology, 13, Technical Publication No. 2003 (1946). (19) Die Korrosion Metallische Werk-stoffe, S. Hirzel, Leipzig, 1936, p. 120. (20) C. Brown and H. Uhlig, J. Am. Chem. Soc., 69, 463 (1947). (21) Definition introduced by Dr. P. H. Emmett. (22) P. S. Flory, J. Chem. Phys., 9, 660 (1941). M. L. Huggins, Ibid., 9, 440 (1941). (23) H. Eyring, Ibid., U, 283 (1936). (24) J. Kendall, Meddel. Vetenskap-sakad., Nobelinst., 2, 25 (1913). (25) See Reference in note, p. 100. (26) R. E. Powell and H. Eyring, J. Am. Chem. Soc., 65, 648 (1943).