Adsorption from a one-dimensional lattice gas and the Brunauer Emmett Teller equation

Similar documents
GAS-SURFACE INTERACTIONS

CHEMISTRY 443, Fall, 2014 (14F) Section Number: 10 Examination 2, November 5, 2014

University of Illinois at Chicago Department of Physics SOLUTIONS. Thermodynamics and Statistical Mechanics Qualifying Examination

Simple lattice-gas model for water

Steady One-Dimensional Diffusion of One Species A through a Second Non-Transferring Species B. z y x. Liquid A

ADSORPTION IN MICROPOROUS MATERIALS: ANALYTICAL EQUATIONS FOR TYPE I ISOTHERMS AT HIGH PRESSURE

Flow and Transport. c(s, t)s ds,

Chemical Potential. Combining the First and Second Laws for a closed system, Considering (extensive properties)

Lecture 2: Receptor-ligand binding and cooperativity

Colors of Co(III) solutions. Electronic-Vibrational Coupling. Vibronic Coupling

The Simple Harmonic Oscillator

ikt kt x l ktk xl w1(r) could originate as a configurational entropy term, rather than as potential

the protein may exist in one conformation in the absence of ligand the ligand that binds. On the contrary, the binding of different

Particle Size Determinations: Dynamic Light Scattering: page 161 text

Polynomial Functions of Higher Degree

PREDICTING EQUILIBRIUM MOISTURE CONTENT OF WOOD BY MATHEMATICAL MODELS

SIMPLE MODELS FOR SUPERCRITICAL EXTRACTION OF NATURAL MATTER

3.320 Lecture 18 (4/12/05)

Translation Groups; Introduction to Bands and Reciprocal Space. Timothy Hughbanks, Texas A& M University

Statistical Geometry and Lattices

3.5. Kinetic Approach for Isotherms

Theoretical study of the effect of enzyme-enzyme interactions

Lecture 5. Solid surface: Adsorption and Catalysis

Unusual Entropy of Adsorbed Methane on Zeolite Templated Carbon. Supporting Information. Part 2: Statistical Mechanical Model

Core Connections Algebra 2 Checkpoint Materials

Lattice protein models

Introduction. Model DENSITY PROFILES OF SEMI-DILUTE POLYMER SOLUTIONS NEAR A HARD WALL: MONTE CARLO SIMULATION

Monolayers. Factors affecting the adsorption from solution. Adsorption of amphiphilic molecules on solid support

Phase Transitions. µ a (P c (T ), T ) µ b (P c (T ), T ), (3) µ a (P, T c (P )) µ b (P, T c (P )). (4)

We already came across a form of indistinguishably in the canonical partition function: V N Q =

Lecture 4.2 Finite Difference Approximation

Chapter 9 Prerequisite Skills

MORE CURVE SKETCHING

Surface chemistry. Liquid-gas, solid-gas and solid-liquid surfaces. Levente Novák István Bányai

Group, Rings, and Fields Rahul Pandharipande. I. Sets Let S be a set. The Cartesian product S S is the set of ordered pairs of elements of S,

Physics 53. Thermal Physics 1. Statistics are like a bikini. What they reveal is suggestive; what they conceal is vital.

Thermodynamic expansion of nucleation free-energy barrier and size of critical nucleus near the vapor-liquid coexistence

The Liquid Vapor Interface

CHAPTER 2 THE MICROSCOPIC PERSPECTIVE

The 1+1-dimensional Ising model

Thermodynamics at Small Scales. Signe Kjelstrup, Sondre Kvalvåg Schnell, Jean-Marc Simon, Thijs Vlugt, Dick Bedeaux

MOL 410/510: Introduction to Biological Dynamics Fall 2012 Problem Set #4, Nonlinear Dynamical Systems (due 10/19/2012) 6 MUST DO Questions, 1

Physics 127b: Statistical Mechanics. Lecture 2: Dense Gas and the Liquid State. Mayer Cluster Expansion

18.303: Introduction to Green s functions and operator inverses

Multivariable Calculus

3.1 Graphs of Polynomials

Chapter 5. Chemical potential and Gibbs distribution

2014 Mathematics. Advanced Higher. Finalised Marking Instructions

I.G Approach to Equilibrium and Thermodynamic Potentials

PROBLEMS In each of Problems 1 through 12:

On the Chemical Free Energy of the Electrical Double Layer

Mean-field theory for arrays of Josephson-coupled wires

Introduction. Chapter The Purpose of Statistical Mechanics

Core Connections Algebra 2 Checkpoint Materials

JUST THE MATHS UNIT NUMBER 7.2. DETERMINANTS 2 (Consistency and third order determinants) A.J.Hobson

1. Heterogeneous Systems and Chemical Equilibrium

Chapter 4 (Lecture 6-7) Schrodinger equation for some simple systems Table: List of various one dimensional potentials System Physical correspondence

Fig. 3.1? Hard core potential

A simple technique to estimate partition functions and equilibrium constants from Monte Carlo simulations

Entanglement in Many-Body Fermion Systems

On the local and nonlocal components of solvation thermodynamics and their relation to solvation shell models

Polymer Solution Thermodynamics:

Statistical Physics. Solutions Sheet 11.

Statistical Thermodynamics. Lecture 8: Theory of Chemical Equilibria(I)

CHAPTER 4 BOUNDARY LAYER FLOW APPLICATION TO EXTERNAL FLOW

Physics 127a: Class Notes

Pre-Calculus Summer Packet

arxiv:cond-mat/ v1 17 Aug 1994

The 6-vertex model of hydrogen-bonded crystals with bond defects

At constant temperature, the first term is zero. Using Maxwell relation and for

One Solution Two Solutions Three Solutions Four Solutions. Since both equations equal y we can set them equal Combine like terms Factor Solve for x

Supporting Online Material (1)

Answer Key 1973 BC 1969 BC 24. A 14. A 24. C 25. A 26. C 27. C 28. D 29. C 30. D 31. C 13. C 12. D 12. E 3. A 32. B 27. E 34. C 14. D 25. B 26.

Concepts in Surface Physics

Chapter 6. Nonlinear Equations. 6.1 The Problem of Nonlinear Root-finding. 6.2 Rate of Convergence

Lecture 9 Examples and Problems

Ice Nucleation Inhibition

Intermediate mass strangelets are positively charged. Abstract

Competition between face-centered cubic and icosahedral cluster structures

Effect of Sorption/Curved Interface Thermodynamics on Pressure transient

Chapter 7 PHASE EQUILIBRIUM IN A ONE-COMPONENT SYSTEM

Phase Transition & Approximate Partition Function In Ising Model and Percolation In Two Dimension: Specifically For Square Lattices

Solution of Partial Differential Equations

The distortion observed in the bottom channel of Figure 1 can be predicted from the full transport equation, C t + u C. y D C. z, (1) x D C.

CHAPTER 3 LECTURE NOTES 3.1. The Carnot Cycle Consider the following reversible cyclic process involving one mole of an ideal gas:

2. As gas P increases and/or T is lowered, intermolecular forces become significant, and deviations from ideal gas laws occur (van der Waal equation).

Renormalization Group for the Two-Dimensional Ising Model

Thermal and Statistical Physics Department Exam Last updated November 4, L π

Thermodynamic Properties

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department Statistical Physics I Spring Term 2013

THE DISTRIBUTIVE LAW. Note: To avoid mistakes, include arrows above or below the terms that are being multiplied.

Phase transitions of quadrupolar fluids

Taylor Series and Series Convergence (Online)

arxiv: v1 [physics.comp-ph] 14 Nov 2014

Kolloquium Universität Innsbruck October 13, The renormalization group: from the foundations to modern applications

CLAY MINERALS BULLETIN

Phase Transitions in Condensed Matter Spontaneous Symmetry Breaking and Universality. Hans-Henning Klauss. Institut für Festkörperphysik TU Dresden

Algebraic Expressions and Identities

BET Surface Area Analysis of Nanoparticles *

II. Probability. II.A General Definitions

Transcription:

Proc. Natl. Acad. Sci. USA Vol. 93, pp. 1438 1433, December 1996 Chemistry Adsorption from a one-dimensional lattice gas and the Brunauer Emmett Teller equation (Ising modelreference systemgibbs ecess adsorptiongibbs integral) TERRELL L. HILL* Laboratory of Molecular Biology, National Institute of Diabetes and Digestive and Kidney Diseases, National Institutes of Health, Bethesda, MD 089 Contributed by Terrell L. Hill, October 7, 1996 ABSTRACT An eact treatment of adsorption from a one-dimensional lattice gas is used to eliminate and correct a well-known inconsistency in the Brunauer Emmett Teller (B.E.T.) equation namely, Gibbs ecess adsorption is not taken into account and the Gibbs integral diverges at the transition point. However, neither model should be considered realistic for eperimental adsorption systems. The well-known Brunauer Emmett Teller (B.E.T.) equation (Eq. 36 below) is used primarily to determine the surface area from the physical adsorption of a gas on a solid surface. The adsorbate model on which the equation is based (1) is an assembly of independent linear piles of adsorbed molecules (with no vacancies). The piles are in equilibrium with a dilute three-dimensional gas. Because the gas is dilute, no Gibbs reference-system correction is made. In the actual adsorption of a gas on a solid surface, when the gas pressure is increased (at constant temperature) up to the vapor pressure of the corresponding liquid, the gas begins to condense to liquid and the adsorbate becomes bulk liquid on the solid surface. The Gibbs integral (see Eq. 17 below), up to the vapor pressure of the liquid, is finite and related to appropriate surface tensions. In the B.E.T. model, the dilute gas condenses to liquid at its vapor pressure but the independent linear piles of adsorbate molecules, though they become infinite in size, certainly do not become or resemble a liquid. This fundamental inconsistency in the B.E.T. model is reflected in the well-known divergence of the Gibbs integral in this case, which is unrealistic thermodynamically. The objective of this paper is, in contrast to the B.E.T. model, to treat one-dimensional adsorption in a self-consistent way. This is accomplished by considering adsorption initiated by an attracting site at the end of a one-dimensional lattice gas. The adsorbate is then part of the lattice gas and in equilibrium with it. The B.E.T. equation is recovered as a close approimation in a special case, but the approimation does not hold at higher lattice gas densities: a finite Gibbs integral is always found from the eact treatment. Yagov and Lopatkin (, 3) have previously treated this model by a different method and have obtained, in a different form, the equivalent of Eq. 14 below. However, they did not discuss the main topics here: the B.E.T. equation as a limiting case of the general treatment; the rescue of the Gibbs integral from the B.E.T. divergence by use of the present selfconsistent model; a discussion of the thermodynamic functions of the model; and consideration of a special case with interesting symmetry. Neither the B.E.T. model nor the present model should be thought of as resembling, even slightly, eperimental threedimensional adsorption systems. The model here has interesting properties quite aside from the B.E.T. connection. This is essentially the one-dimensional The publication costs of this article were defrayed in part by page charge payment. This article must therefore be hereby marked advertisement in accordance with 18 U.S.C. 1734 solely to indicate this fact. FIG.1. (a) Reference system: M 7 sites in a circle; three sites are occupied by molecules. Nearest-neighbor occupied sites have an interaction energy w. 0 is the grand partition function. (b) M 7 sites with an adsorbent molecule R at each end. The interaction energy between R and an occupied adjacent site is w.(c)m7sites with two free ends (no adsorbent). (d) M 7 sites with one free end and an adsorbent at the other end. Ising problem with end effects, but it is the end effects that receive the focus of attention. We begin with the general derivation but return to the B.E.T. equation and the other topics mentioned in due course. The Adsorption Model We consider first, as a reference system, the one-dimensional Ising model in Fig. 1a. This is the lattice gas without adsorption and without ends. The number of sites is M (M 7inthe figure), and the number of sites occupied by a molecule is N 0 (N 0 3 in the figure). We are interested here in large systems with M 3. Nearest-neighbor molecules have an interaction energy w (Fig. 1a). A molecule on a site has partition function q(t). The lattice gas is an open system (molecules go onto and off of sites) with thermodynamic variables (chemical potential), M, and T. We shall use the notation, T q qte kt, 0 N 0 M. [1] The variable is an activity, a convenient measure of the chemical potential at constant temperature. The subscript in Eq. 1 refers to the reference system. The nearest-neighbor interaction parameter is y e wkt.if w0 and y 1, we have the simple and familiar relations 0 1, 0. [] 1 0 For attractive interactions, w 0 and y 1. A first-order phase transition does not occur (ref. 4, p. 333) in this onedimensional system, even for very large y (e.g., at very low temperatures). However, by analogy with two- (simple square) Abbreviation: B.E.T., Brunauer Emmett Teller. *Present address: 433 Logan Street, Santa Cruz, CA 9506. 1438

Chemistry: Hill Proc. Natl. Acad. Sci. USA 93 (1996) 1439 and three- (simple cubic) dimensional lattice gases, phase transitions occur at 0 1 when T T c, we can take 0 1 as locating a pseudo-transition point. The onedimensional 0 (, y) is given below in Eq. 15; 0 1 when 1y. Incidentally, there are somewhat more elaborate onedimensional lattice gas systems that do ehibit a first-order phase transition (5, 6). In Fig. 1b, the lattice gas has a permanent adsorbent molecule R at each end. As indicated in the figure, the interaction energy between the adsorbent molecule and a first-row lattice gas molecule is w. Any perturbation of q for a first-row molecule can be included in w. Usually w 0 and z e wkt 1. If z 1 and especially if z 1, etra lattice-gas molecules will tend to accumulate on the sites near R (the effect would be spatially propagated by y 1, and especially by y 1). This etra accumulation, or surface ecess (Gibbs), represents the amount of adsorption. It is possible for the surface ecess to be negative. Let be the grand partition function for the system in Fig. 1b, with variables, M, T (M is very large), and let 0 be the grand partition function for the reference system (Fig. 1a) with the same, M, T. Then (ref. 4, p. 39) N N 0 ln 0 T is the etra accumulation of molecules at two ends and N e N N 0 ln 0 1 T [3] [4] is the desired Gibbs surface ecess (amount of adsorption) at one end. An alternative to the use of ( 0 ) 1/ in Eq. 4 is to substitute (see Fig. 1) 1 1/ in place of 1/, 1 (z 1). Both methods yield the same result, though the second has less complicated algebra. For the second method, we need to find 0 and 1.Weuse a standard matri procedure for both. For 0 (ref. 4, p. 35), 0 A 1, A, 3 A M, 1 1,..., M1 TrA M 1 M M, [5] A 1 1 and 1 and are the two eigenvalues of A, [6] y 1, 1 y 1 y 4 1, [7] with 1. Because M 3, 0 1M. For 1 (ref. 4, p. 339), 1 r 1 A 1, A M1, M c M 1,..., M1 r 1 A M1 1, M c M, [8] 1,..., M1 r 1, z, c 1 1. [9] After a similarity transformation (ref. 4, p. 339), we find (using M 3 ) 1 1 M1 1 1 z z 1 1 [10] 1 1 1 z z 1 0 1 1 [11] 0 1 1 [1] 1 0 1 0 0 1 Finally, from Eq. 4, N e y u 1 1 1 1 z z 1 1 1 1. [13] zy u 1 z z 1 y u u 1 u y y, 1 1 y 4 1. [14] Also, for the reference system, because 0 1M, 0 N 0 M ln 1 T y u 1 1 y.[15] There are two Gibbs integrals of particular interest for this model: 0 0 N e d N e 0 dln 1 0 ln 1 0 ln 0 d ln 1 1y 0 ln 0 z y 1 [16] y 1 z y 1 1 y 1 1. [17] The lower limit does not contribute. Both integrals are finite. The second is significant because 1y locates the pseudotransition point, as mentioned above. Ecess Thermodynamic Functions Because we have been discussing a single one-dimensional lattice gas chain of sites, the thermodynamics of small systems (7) could be used here. However, since the appropriate adsorption thermodynamic formulation is already available (ref. 8, pp. 5 53), we use this alternative approach. We consider a surface with B adsorbent molecules, from each of which a lattice gas chain of M sites etends vertically. M and B are both large. Molecules at chemical potential go on and off of the sites. The chains are independent of each other. The total number of lattice sites is MB (a volume ) and the mean total number of occupied sites is N a. The basic thermodynamic relations for this system are (ref. 8, pp. 5 53) de a TdS a dmb db dn a [18] E a TS a MB B N a. [19] The adsorbent molecules merely present an eternal field; they are not included in the thermodynamic functions. The reference system consists of the same MB sites but with no adsorbent molecules present. The lattice gas chemical potential is again and mean properties of the reference

14330 Chemistry: Hill Proc. Natl. Acad. Sci. USA 93 (1996) system etend right up to the surface (no end effects; compare Fig. 1a). For the reference system, then, de g TdS g dmb dn g [0] E g TS g MB N g. [1] On subtracting Eqs. 0 and 1 from Eqs. 18 and 19 (e.g., E E a E g ), we obtain relations among the Gibbs ecess functions: de TdS db dn e B [] E TS B N e B. [3] The physical significance of is a spreading pressure, tending to increase B. From Eqs. and 3 we deduce N e d SBdT d. [4] At constant T, d ktdln. Then (see Eq. 16) kt0 d ktn e dln T constant [5] N e dln ktln 1 0. [6] Thus Eq. 13 gives an eplicit epression for () and, of course, Eq. 14 provides N e(). To obtain SB and EB, we use Eq. 4: S B T TT T. [7] Eq. 5 gives () T and (T) follows easily from Eq. 1. A fairly long calculation (taking w and w as constants), using Eqs. 13 and 6, yields (T). The final result is SB EBT T N e T, [8] E B N e kt dln q dt 1 yw 1 1 zw yw 1 1 yw 1 1 1 z z 1 1 y1 yw. [9] The terms in w and w arise from nearest-neighbor or adsorbent interactions. The w term has a simple interpretation because the reference system has no w interactions. Thus, the coefficient of w gives 1, the probability that the first-row site in the adsorption system is occupied: 1 z 1 1. [30] 1 z z 1 The coefficient of w gives the mean number of ecess w interactions in the adsorption system. The general i can be found as follows. Replace q for the row i site by q. The matri procedure can be generalized to give 1 (). Then i ln 1 () at 1. The factor counts row i occupancy. One finds (i ) i 1 i1 i1 z i 1 i i1 1 i1 1 i. 1 1 1 z z 1 [31] If i 3, i 3 0.Ify1, i (1 ) for i. Eq. 31 is also correct for i 1. The following sum is easy to perform: N e i 0 i1 1 z 1 z 1 1 1 z z 1. [3] This is a third equivalent epression for N e. Relation to the B.E.T. Equation To simulate the B.E.T. model we need very large y (low temperature) to eliminate almost all vacancies in the adsorption lattice gas pile of molecules. Also, the pseudo-transition point is 1y, ory 1. Thus, y here corresponds to in the B.E.T. equation ( 1 is the transition point). Further, here at very small, N e [(z 1)y]y. Because c is the corresponding B.E.T. epression, c corresponds to (z 1)y here. Ordinarily c is of order 10 or 100; hence z is even larger than y. Thus, c 7 zy. In summary, B.E.T. 7 y, c 7 zy, c 7 z. [33] We now consider Eq. 14 for N e with y and z both very large. Because y O(1), 1. Hence, we use the following epansions: 1 y 1 y, 11 1y y 1 y,1zz 11 z. 1 y [34] If we write the four terms in []ineq.14 as A BC D, we then find A D y y, C 1 y 1 y, B z 1 y z. [35] Thus, only B contributes significantly. Then N e B z 1 y1 y z. [36] The same result follows in similar fashion from ref. or ref. 3. In view of the correspondences in 33, this is the familiar B.E.T. equation. Table 1 refers to the case y 10 3, z 10 5, B.E.T. c zy 00. Fig. shows, for this case, N e from Eqs. 14 and 36, as functions of y. Also included in the figure is 0 (y), Eq. 15, for y 10 3. It is seen from the table and the figure that the B.E.T. approimation to N e is adequate in this case up to about y 0.75. As mentioned at the outset, the B.E.T. Table 1. The B.E.T. approimation with y 10 3, z 10 5 y N e (eact) B (eact) Eq. 36 (B.E.T.) 0.1 1.063 1.063 1.063 0. 1.5 1.5 1.5 0.4 1.650 1.65 1.654 0.6.468.477.49 0.8 4.733 4.84 4.994 0.9 8.05 8.641 9.994 1.0 8.153 16.309

Chemistry: Hill Proc. Natl. Acad. Sci. USA 93 (1996) 14331 If y is increased (beyond 10 3 ) along with z in the same ratio (c zy), the 0 curve in Fig. becomes steeper at 0 1 (the slope is y 1/ 4) so that the B.E.T. approimation would be adequate for larger y, and the N e 3 0 approach would occur at smaller y. The Gibbs integrals using the eact N e, Eqs. 16 and 17, are finite. The B.E.T. Gibbs integral to y 1 diverges because of the factor (1 y) 1 in Eq. 36. For large y and z, the integral in Eq. 17 is approimately ln[z(y) 1/ ]. In an eperimental physical adsorption system at T T c, N e 3 as p 3 p 0 (vapor pressure), but the Gibbs integral to p 0 is finite. The interaction (w, w) part of EB in Eq. 9 becomes, in the B.E.T. special case, using Eqs. 34, E ww zw zyw B 1 y z 1 y1 y z. [37] The coefficient of w is the epected B.E.T. 1 and the coefficient of w can be shown to be the B.E.T. mean number of first-neighbor pairs in a pile. Other Special Cases FIG.. The solid curve labeled N e is the case y 10 3, z 10 5, plotted against y. The dashed curve labeled B.E.T. is an approimation. The solid curve labeled 0 (different scale) is 0 (y) for the reference system with y 10 3. equation does not include a Gibbs reference system correction. The reference system here, for self-consistency, is the onedimensional Ising model (Fig. 1a). The B.E.T. omission of a Gibbs correction is appropriate in the present model so long as 0 (y) (Fig. ) is negligible. Above y 0.75, 0 is not negligible: Eqs. 14 and 36 differ significantly. Note that the eact N e 3 0as 0 31 (at y 1). This is to be epected because the introduction of adsorbent can hardly add molecules to an already virtually saturated reference system; hence, N e 3 0. (However, if z is small, the adsorbent can still subtract molecules when 0 3 1.) The shape of the N e(y) curve in Fig. resembles qualitatively eperimental supercritical Gibbs ecess adsorption isotherms (9). This is not surprising because any finite y 1inthe one-dimensional lattice gas model corresponds to T T c (T c 0 K): all N e curves are supercritical. The interested reader will find it easy to eamine special cases of Eq. 14 either analytically or by computer-generated curves. I mention only a few eamples below. Also, Yagov and Lopatkin () show some curves in the form N e( 0 ), using 0 () to eliminate. From Eq. 16 we note that the complete Gibbs integral is zero if z y 1/. Also, Eq. 14 yields N e 0ifweputzy 1/ and 1y. Fig. 3 includes N e() for the particular case z 10, y 100. This curve, and others with z y 1/, suggest the symmetry property N e() N e(1y )orn e(y) N e(1y). It is easy to deduce the zero Gibbs integral from the symmetry property. Further, the symmetry property has been confirmed using Eq. 3. Also, from Eq. 31, i () i (1y ) 1 for all i; i (1y) 1. If y 1, the sites are independent of each other. Eq. 14 reduces in this case to N e z 1 z 1 1 0. [38] The only contribution is from the first row. When z 1, the adsorbent is, in effect, removed. When z 0, the first-row site is empty ( 1 0); the remainder of the chain (rows, 3,... ) behaves like a full z 1 chain. Thus, N e z 0 1 0 N e z 1 0 N e z 1. [39] This can be confirmed from Eq. 14. Fig. 3 includes N e() for z 1, y 100. When y 0 (nearest-neighbor sites cannot be occupied), one finds from Eqs. 14 and 15 that, as 3, 0 3 1 and N e 3 1 4. When y and z are both large (as in B.E.T.), N e y 1/ 4at 1y. Some epansions at small : N e z 1 4 3y z zy z 0 y3 1 z z zy z [40] y z zy. FIG. 3. The two upper curves are N e() for z 10, y 100 and z 1, y 100. The lowest curve is 0 () (different scale) for y 100. I am indebted to Dr. Gregory Aranovich for reviving (50 years later) my interest in adsorption problems and for bringing refs., 3, and 9 to my attention.

1433 Chemistry: Hill Proc. Natl. Acad. Sci. USA 93 (1996) 1. Hill, T. L. (1987) Linear Aggregation Theory in Cell Biology (Springer, New York), p. 34.. Yagov, V. V. & Lopatkin, A. A. (1990) Russ. J. Phys. Chem. Transl. of Zh. Fiz. Khim. 64, 564 566. 3. Yagov, V. V. & Lopatkin, A. A. (1993) Russ. J. Phys. Chem. Transl. of Zh. Fiz. Khim. 67, 1399 140. 4. Hill, T. L. (1985) Cooperativity Theory in Biochemistry (Springer, New York). 5. Hill, T. L. (1967) Proc. Natl. Acad. Sci. USA 57, 7 3. 6. Hill, T. L. & White, G. M. (1967) Proc. Natl. Acad. Sci. USA 58, 454 461. 7. Hill, T. L. (1994) Thermodynamics of Small Systems (Dover, New York). 8. Hill, T. L. (195) Adv. Catal. 4, 11 58. 9. Findenegg, G. H. (1983) Fundamentals of Adsorption (Engineering Foundation, New York), pp. 07 18.