Thermodynamics of Enzyme-Catalyzed Reactions


Robert N. Goldberg, Yadu B. Tewari, and T. N. Bhat

Biotechnology Division

National Institute of Standards and Technology

Gaithersburg, MD 20899 U.S.A.

1. Introduction

Thermodynamic data on enzyme-catalyzed reactions play an important role in the prediction of the extent of reaction and the position of equilibrium for any process in which these reactions occur. The importance of understanding the thermodynamics of these biochemical reactions was emphasized by Krebs and Kornberg in their monograph "A Survey of the Energy Transformations in Living Matter".1 Their monograph also contains a useful appendix on Gibbs energy data of biological interest and a table on the thermodynamics of enzyme-catalyzed reactions. However, the amount of data available at that time was extremely limited. Reviews on various aspects of this subject have subsequently appeared.2-10 Each of these reviews, however, has been limited in the extent of coverage given to this area and no comprehensive review exists. Thermodynamic information is also needed in biotechnology when one needs to optimize product yields and to calculate the energy requirements of a given reaction.

It is the aim of this project to provide a compilation of data on the thermodynamics of enzyme-catalyzed reactions. The data presented herein is limited to direct equilibrium and calorimetric measurements performed on these reactions under in vitro conditions. This is the principal thermodynamic information that is needed to determine the position of equilibrium of a given reaction.

The following information is given for each entry in this database: the reference for the data; the reaction studied; the name of the enzyme used and its Enzyme Commission number; the method of measurement; the conditions of measurement (temperature, pH, ionic strength, and the buffer(s) and cofactor(s) used); the data and an evaluation of it; and, sometimes, commentary on the data and on any corrections which have been applied to it. The absence of a piece of information indicates that it was not found in the paper cited.

2. Arrangement of Data

The primary ordering of the data is done in accord with the Enzyme Commission (EC) number assigned to the enzyme which has been used to catalyze the reaction studied. Here, we have followed the classification scheme recommended by the Nomenclature Committee of the International Union of Biochemistry as presented in "Enzyme Nomenclature".11 If more than one enzyme was used to catalyze a reaction, the data is entered in the tables immediately following the enzyme used which occurs first in the numerical order established by the Enzyme Commission. A cross reference is also made in the tables under each of the individual enzymes which have been used in the study cited. When there is more than one reaction catalyzed by a given enzyme, the reactions listed under a given enzyme heading are ordered alphabetically according to the substrates involved in the reaction. Here, the substrates are ordered from left to right in the direction the reactions are written. If necessary, the ordering of the data for any given reaction is determined by the year in which the measurements were performed. If more than one set of measurements appeared in a given year, the entries are ordered by the last name(s) of the author(s).

While the above scheme is a useful way to order the thermodynamic data on enzyme-catalyzed reactions, one possible complication is that a given reaction can sometimes be catalyzed by more than one enzyme. Also, we have generally relied upon the author(s) of a given study for the identification of the enzyme. The user of these tables is therefore advised, when looking for data on a given reaction, to also look in these tables under those enzymes that catalyze reactions that are closely related to the one sought.

3. Evaluation of Data

The subjective evaluation of the data in this database consisted of the assignment of a rating: A (high quality), B(good), C (average), or D (low quality). In making these assignments we considered the various experimental details which were provided in the study. These details include the method of measurement, the number of data points determined, and the extent to which the effects of varying temperature, pH, and ionic strength were investigated. A low rating was generally given when few details of the investigation were reported. For example, in many of the papers cited, the major aim of the study was the isolation and purification of the enzyme of interest. Thus, the equilibrium data were obtained as only a small part of an investigation to characterize many of the properties of that enzyme. A similar rating system was used by Domalski, Evans, and Hearing in their review "Heat Capacities and Entropies of Organic Compounds in the Condensed Phase".12

The complete evaluation of the data contained in these tables would involve the adjustment of the results obtained for each reaction to a common standard state followed by thermochemical cycle calculations to determine how well the data for the various reactions fit together. The adjustment to a standard state generally requires thermodynamic data on hydrogen and metal ion binding to the substrates involved in the reaction as well as information on the activity coefficients of the species in the solution.13 This is beyond the scope of this database. It is our belief that this complete evaluation would have great value in the organization and systematization of these data. However, such evaluations have been performed only for a limited set of biochemical substances and reactions (see refs. 6, 9, and 10). These evaluations, which involved a solution of thermochemical networks, also included thermodynamic data on the condensed phases and, in some cases, electrochemical potentials. Thus, third law entropies and enthalpies of combustion, enthalpies of solution, solubilities, and activity coefficients were also included to make the necessary thermodynamic ties to the condensed phases.

It is highly desirable that equilibrium be approached from both directions in any equilibrium investigation. The failure of some investigators to do this is a serious oversight in the experimental work and is a primary source of systematic error in much of the "equilibrium" data reported in the literature. Another source of systematic error in many of these studies is sample impurity. As an example, Miller and Smith-Magowan10 have pointed out that samples of nicotinamide-adenine dinucleotide (NAD) and nicotinamide-adenine dinucleotide phosphate (NADP) can also contain the respective alpha-isomers. Since, most enzymes are specific for the beta-isomer, the presence of an unrecognized amount of the alpha-isomer can be a source of systematic error in equilibrium or calorimetric measurements. Occasionally, we have performed calculations on the data given in the papers or extracted values from figures in the papers. For example, we have sometimes been able to calculate an equilibrium constant where none has previously been calculated, e.g. if the investigator(s) have given results in terms of per cent conversion for a reaction of the type A = B.

4. Thermodynamic Conventions

There are two fundamentally different types of equilibrium constants given in these tables. This is illustrated by the following example for the hydrolysis of adenosine 5'-triphosphate (ATP) to adenosine 5'-diphosphate (ADP) and orthophosphate:

ATP(aq) + H2O(l) = ADP(aq) + orthophosphate(aq). (1)

The apparent equilibrium constant for the overall biochemical reaction (1) is

Kc' = c(ADP)c(orthophosphate)/{c(ATP)co}. (2)

The ATP, ADP, and orthophosphate each exist in several different ionized and metal bound forms: e.g. ATP4-, HATP3-, H2ATP2-, MgATP2-, MgHATP-, etc. Thus, ATP has often been denoted in the literature as SigmaATP or as (ATP)tot. When it is clear that one is dealing with total amounts of substances, it is not necessary to use either the Sigma or "tot" and these designations can be dispensed with. This has been done throughout the tables of data. In the above equation, co is 1 mol dm-3; it is included to make the apparent equilibrium constant dimensionless. The subscript c has been attached to the apparent equilibrium constant to indicate that it is based on the concentration (molarity) scale.

The apparent equilibrium constant can be used to calculate the standard transformed Gibbs energy of reaction DeltarG'o at specified conditions of temperature T, pressure p, ionic strength (Im when the units are mol kg-1 and Ic when the units are mol dm-3), pH, and pMg:

DeltarG'o = -RT lnK'. (3)

The gas constant R is equal to 8.31451 J K-1 mol-1. Either DeltarG'oor the apparent equilibrium constant K' can be used to determine the position of equilibrium of biochemical reactions.14,15

It is also possible to choose a reference reaction involving selected solute species:

ATP4-(aq) + H2O(l) = ADP3-(aq) + HPO42-(aq) + H+(aq). (4)

The equilibrium constant for this reference reaction is

Kc(ref) = c(ADP3-)c(HPO42-)c(H+)/{c(ATP4-)(co)2}. (5)

Equations which relate these two different types of equilibrium constants have been derived.16,17 To calculate the equilibrium constant Kc(ref) from the apparent equilibrium constant K, or vice versa, one needs the equilibrium constants for the binding of hydrogen and metal ions to ATP4-, ADP3-, and HPO42-.

The standard or thermodynamic equilibrium constant for reaction (4) is written in terms of activities a:

K(ref) = a(ADP3-)a(HPO42-)a(H+)/{a(ATP4-)a(H2O)}. (6)

The determination of K(ref) requires either a knowledge of both Kc(ref) at a single ionic strength and of the appropriate ratio of activity coefficients or an extrapolation of values of Kc(ref), determined at several ionic strengths, to I = 0. The distinction between the equilibrium constant Kc(ref), as given in Eq. (5), and the apparent equilibrium constant K, as given in Eq. (2), will be maintained throughout this database. Where possible, we have also included calculated values of standard equilibrium constants which were given in the various papers examined.

To avoid confusion between these different types of equilibrium constants and to avoid ambiguity about whether specific species or sums of species are intended, ammonia, for example rather than NH3 or NH4+, will be written when we mean total ammonia. The chemical formulas will be used when we mean one of these specific species. Other substances such as carbon dioxide (CO2, HCO3-, and CO32-), and orthophosphate (H2PO4-, HPO42-, and PO43-) will be treated in the same manner. Exceptions will be made for water which will always be written as H2O and for gaseous hydrogen and oxygen which will be written as H2(g) and O2(g), respectively.

A similar situation exists for enthalpies of reaction. Thus, the standard transformed enthalpy change for a biochemical reaction (e.g. reaction (1) above) at specified conditions of temperature, pressure, pH, pMg, and ionic strength will be designated as DeltarH'o and the enthalpy change for a reference reaction (e.g. reaction (4) above) as DeltarHo(ref). Enthalpy changes for a reference reaction which have been adjusted to a standard state will be designated as DeltarHo(ref) at I = 0. The standard transformed enthalpy of reaction DeltarH'o can be calculated with the van't Hoff equation from apparent equilibrium constants which have been determined as a function of temperature. Unless indicated otherwise, we have assumed that DeltarH'o is independent of temperature (i.e. DeltarCp'o = 0) when performing this calculation. An enthalpy of reaction which has been determined calorimetrically will be designated as DeltarH(cal). To obtain either DeltarH'o or DeltarHo(ref) from calorimetric data requires that corrections be made for heat effects due to possible changes in pH, pMg, ionic strength, and for the protonation of the buffer. In the case where the solutions are well buffered and the changes in pH, pMg, and ionic strength are small, the major correction is for the enthalpy of protonation of the buffer by any protons released (or absorbed) as a consequence of the reaction. Thus,

DeltarH'o = DeltarH(cal) - DeltarN(H+)DeltarHo(buff) (7)

where DeltarN(H+) is the change in the binding of H+(aq) accompanying the biochemical reaction and DeltarHo(buff) is the enthalpy of ionization of the buffer. Since this correction can be substantial, it will always be noted in these tables whether or not calorimetric results have been corrected for the enthalpy of protonation of the buffer. Unless indicated otherwise, we have relied upon the correction(s) made by the actual investigators. Enthalpies of reaction calculated from equilibrium constants measured at several temperatures require no such correction.

In this database, all units of energy are given in terms of the joule with conversions from results in calories based upon 1 cal = 4.184 J. The equilibrium and calorimetric results summarized in this database were all done at atmospheric pressure ( 0.101325 MPa). The standard pressure for thermodynamic results is 0.1 MPa.

Occasionally, some investigators have calculated apparent equilibrium constants where the concentration of water was included. Thus, for the above example involving ATP hydrolysis, the apparent equilibrium constant might be written as

Kc' = c(ADP)c(orthophosphate)/{c(ATP)c(H2O)}. (8)

Where this has been done, we have adjusted the reported values of K to the usual convention represented by Eq. (2). At low solute concentrations, the concentration of water c(H2O) is 55.35 mol dm-3 at 298.15 K and 55.14 mol dm-3 at 310.15 K.

Equilibrium constants should be expressed as dimensionless quantities. However, the numerical value obtained for the equilibrium constant of an unsymmetrical reaction will depend upon the measure of composition and standard concentration selected for the reactants and products. Thus, for the chemical reaction

A(aq) = B(aq) + C(aq), (9)

Kc = c(B)c(C)/{c(A)co}, Km = m(B)m(C)/{m(A)mo}, and Kx = x(B)x(C)/x(A). Here, m and x are, respectively, molality and mole fraction and mo = 1 mol kg-1. The equilibrium constant expressed in terms of mole fractions is automatically dimensionless. In this database, the various types of equilibrium constants for unsymmetrical reactions are denoted by their appropriate subscripts; no subscript is needed for a symmetrical reaction. When necessary, results have been adjusted to units of mol dm-3 to obtain the numerical value(s) of Kc given in these tables.

The density of the solution enters into the conversion of equilibrium constants which have been determined on a concentration basis, namely Kc to Km, or Kc to Kx. This causes a slight complication in the calculation of the enthalpy of reaction from the derivative of Kc with respect to temperature since the derivative of the density of the solution with respect to temperature enters into this calculation. Fortunately, this is a small effect that can be neglected except for the most precise and accurate data. Also, this is not a problem for symmetrical reactions where the equilibrium constants are the same for any standard state concentration. Enthalpies of reaction calculated from the temperature derivatives of equilibrium constants which are on a molality or mole fraction basis are identical to calorimetrically determined enthalpies of reaction which have been corrected for the enthalpy of buffer protonation.

Alcohol dehydrogenase (EC 1.1.1.1) catalyzes the NAD/NADH coupled conversion of ethanol to acetaldehyde:

NAD(aq) + ethanol(aq) = NADH(aq) + acetaldehyde(aq). (10)

The apparent equilibrium constant for the overall biochemical reaction (9) is

K' = c(NADH)c(acetaldehyde)/{c(NAD)c(ethanol)}. (11)

The reference reaction corresponding to reaction (9) is:

NAD-(aq) + ethanol(aq) = NADH2-(aq) + acetaldehyde(aq) + H+(aq). (12)

The equilibrium constant for this reference reaction is:

Kc(ref) = c(NADH2-)c(acetaldehyde)c(H+)/{c(NAD-)c(ethanol)co}. (13)

In some cases, however, investigators have written:

NAD(aq) + ethanol(aq) = NADH(aq) + acetaldehyde(aq) + H+(aq). (14)

This is neither a chemical reference reaction nor an (overall) biochemical reaction. It is an incorrect representation because it implies that exactly one hydrogen ion was produced in the reaction. This may not always be the situation. Secondly, the equation is not electrically balanced (it would only appear to be balanced if one were to write NAD+ instead of NAD). In papers where this reaction has been written as in Eq. (14), the investigators have often given the quantity [c(NADH)c(acetaldehyde)c(H+)/{c(NAD)c(ethanol)co}] which is equal to [K'c(H+)/co]. Since this quantity is not an apparent equilibrium constant, the standard transformed Gibbs energy of reaction should not be calculated from it. Thus, when investigators have given numerical values of the apparent equilibrium constant K according to Eq. (11), we have included the results in these tables. However, when the only result reported is K'c(H+)/co, it is also given in these tables along with the pH range over which the measurements were performed. However, we have rewritten the reaction and, when possible, calculated the apparent equilibrium constant in accord with Eq. (10). The oxidized form of nicotinamide-adenine dinucleotide, which is generally written as NAD+ in the literature, does not have an electrical charge of +1 and, for this reason, a superscript + has not been attached to the NAD. The charges attached to the NAD and NADH in reaction (12) are those that would be expected for the predominant ionic forms at pH = 7. The assignment of these charges is based upon an examination of the structures of these substances and a knowledge of the acidity constants of the reactants and products (see below).

For reaction (10) and many similar reactions, the quantity K'c(H+)/co has frequently been found to be independent of pH. The reason for this is that NAD has a pK of 3.88,18,19 NADH a pK of 4.46,19 and the ethanol and acetaldehyde are not ionized unless placed in extremely alkaline solutions. Thus, at neutral pH, the predominant species in aqueous solution are NAD-, NADH2-, ethanol0, and acetaldehyde0. These same species should also be predominant for 5 < pH < 10. Because of this, the quantity [c(NADH)c(acetaldehyde)c(H+)/{c(NAD)c(ethanol)co}] = [K'c(H+)/co] is very nearly equal to Kc(ref). The situation for NADP/NADPH coupled reactions is similar to that for NAD/NADH reactions. The only difference is that NADP also has a pK at 6.1 due to the ionization of the phosphate group attached at the 2'-hydroxyl position in the adenosine.20 On the basis of structure, one would expect the pKs for the ionization of this phosphate group to be nearly equal for both NADP and NADPH. Thus, the statements made above regarding the thermodynamics of NAD/NADH coupled reactions should also be true for NADP/NADPH coupled reactions.

The review of Miller and Smith-Magowan10 contains a discussion of some thermodynamic pathways leading to the standard Gibbs energy and enthalpy changes for the following reactions:

NAD-(aq) + H2(g) = NADH2-(aq) + H+(aq), (15)

NADP3-(aq) + H2(g) = NADPH4-(aq) + H+(aq). (16)

For reaction (15) at T = 298.15 K and Im = 0.1 mol kg-1, Miller and Smith-Magowan10 give DeltarGo = 20.0 kJ mol-1 and DeltarHo = -32.1 kJ mol-1. For reaction (16) at T = 298.15 K and Im = 0.1 mol kg-1, they10 give DeltarGo = 21.1 kJ mol-1 and DeltarHo = -26.7 kJ mol-1. Here, we have rewritten the corresponding equations in Miller and Smith-Magowan's Table 7 in accordance with the preceding discussion concerning the notation for NAD and NADP type species. For additional information related to the thermodynamics of these two important reactions, the reader is also referred to the study by Burton21 and the review by Rekharsky et al.22

5. Some Aspects and Uses of Thermodynamic Data on Biochemical Reactions

While a full discussion of the applications of thermodynamic data would be beyond the scope of this database, it seems useful to briefly indicate the utility of the information presented in this database. The primary motivation for many of those performing thermodynamic studies on biochemical reactions is to determine the position of equilibrium of the reaction(s) studied as well as to establish the reversibility of the reaction. This information is concisely summarized in terms of the apparent equilibrium constants given herein. These apparent equilibrium constants can be conveniently used to calculate the extent of reaction under the stated set of conditions and thus can be very useful to engineers concerned with the optimization of product yield in bioreactors. The enthalpy changes accompanying these reactions are also needed to know how much heating or cooling is required to keep a bioreactor at its proper temperature. To perform this calculation one needs to know both the standard transformed enthalpy of reaction DeltarH'o, the change in the hydrogen ion binding DeltarN(H+), and the enthalpy of protonation of the buffer(s) in the bioreactor. In general, both DeltarH'o and DeltarN(H+) are functions of temperature, ionic strength, and metal ion concentration.

Apparent equilibrium constants obtained from studies of in vitro systems can also be used to calculate the position of equilibrium in metabolic processes involving several reactions. Glycolysis is probably the best example of a situation where data are available and for which such calculations have been performed.14,23 The results of these calculations can then be compared with information on the concentrations of the various substrates obtained from the analysis of in vivo systems. This comparison can provide valuable insight into the chemical machinery of living systems.

For many biochemical reactions, the apparent equilibrium constant is a function of temperature, pH, pX, and ionic strength. Thus, when performing Hess' Law and thermochemical cycle calculations, it is necessary that the data for all of the reactions in such a calculation refer to the same set of conditions. The dependencies of apparent equilibrium constants and standard transformed enthalpies of reaction on the conditions of reaction can be very complex and the reduction of such results to a common standard state generally requires auxiliary information on the binding of protons and metal ions to the various reactants as well as information on or assumptions about the activity coefficients of the species in solution. Calculations of this type have been performed by Kuby and Noltmann,24 Alberty,25,26 Guynn, Gelberg, and Veech,27 Langer et al.,28 Goldberg and Tewari,13 and others.

Tables of standard formation properties29 have proven to be a useful way of generalizing upon and presenting thermodynamic data for many chemical substances. However, tables of this type have been prepared for only limited classes of biochemical substances.6,10,30 It has recently been shown31 how it is possible to prepare tables of standard transformed formation properties for biochemical reactants (i.e. sums of species) as distinct from standard formation properties for individual biochemical species. The adenosine 5'-triphosphate series was used as a prototype for this purpose. Thus, it appears likely and desirable that several different types of thermodynamic tables will eventually appear in the literature. Clearly, the larger the scope of such tables, the more useful they are for calculating thermodynamic quantities for reactions which have not been the subject of a direct investigation.

6. This Website

The information contained in this website has been published in the Journal of Physical and Chemical Reference Data as a series of five reviews.32-37 Comments on the thermodynamic database, including additions and corrections, should be sent to Dr. Robert N. Goldberg, Biotechnology Division, National Institute of Standards and Technology, Gaithersburg, MD 20899 U.S.A.; e-mail: robert.goldberg@nist.gov; telephone: 1-301-975-2584; FAX: 1-301-330-3447. Comments dealing with the structure of the website should be sent to Dr. T. N. Bhat, Center for Advanced Research in Biotechnology, 9600 Gudelsky Drive, Rockville, MD 20850 U.S.A.; e-mail:bhat@nist.gov; telephone: 301-738-6279; FAX: 301-738-6255.

In citing this work please refer to:
Goldberg RN, Tewari YB, Bhat TN, "Thermodynamics of Enzyme-Catalyzed Reactions -a Database for Quantitative Biochemistry", Bioinformatics 2004;20(16):2874-2877.

7. References

1. H. A. Krebs and H. L. Kornberg, with an appendix by K. Burton, "A Survey of the Energy Transformations in Living Matter" (Springer-Verlag, Berlin, 1957).

2. M. R. Atkinson and R. K. Morton, in "Comparative Biochemistry", edited by M. Florkin and H. S. Mason (Academic Press, New York, 1960), Vol. 2; pp. 1-95.

3. T. E. Barman, "Enzyme Handbook" (Springer-Verlag, New York, 1969), Vols. I and II.

4. T. E. Barman, "Enzyme Handbook" (Springer-Verlag, New York, 1974), Supplement I.

5. H. D. Brown, in "Biochemical Microcalorimetry", edited by H. D. Brown (Academic Press, New York, 1969); pp. 149-164.

6. R. C. Wilhoit, in "Biochemical Microcalorimetry", edited by H. D. Brown (Academic Press, New York, 1969); pp. 33-81, 305-317.

7. R. K. Thauer, K. Jungermann, and K. Decker, Bacteriol. Rev. 41, 100 (1977).

8. M. V. Rekharsky, A. M. Egorov, G. L. Gal'chenko, and I. V. Berezin, Thermochim. Acta 46, 89 (1981).

9. R. N. Goldberg and Y. B. Tewari, J. Phys. Chem. Ref. Data 18, 809 (1989).

10. S. L. Miller and D. Smith-Magowan, J. Phys. Chem. Ref. Data 19, 1049 (1990).

11. E. C. Webb, "Enzyme Nomenclature 1992" (Academic Press, San Diego, 1992).

12. E. S. Domalski, W. H. Evans, and E. D. Hearing, "Heat Capacities and Entropies of Organic Compounds in the Condensed Phase," J. Phys. Chem. Ref. Data 13, Suppl. 1 (1984).

13. R. N. Goldberg and Y. B. Tewari, Biophys. Chem. 40, 241 (1991).

14. R. A. Alberty, Biophys. Chem. 42, 117 (1992).

15. R. A. Alberty, Biophys. Chem. 43, 239 (1992).

16. R. A. Alberty, J. Biol. Chem. 243, 1337 (1968).

17. R. C. Phillips, P. George, and R. J. Rutman, J. Biol. Chem. 244, 3330 (1969).

18. C. E. Moore, Jr. and A. L. Underwood, Analyt. Biochem. 29, 149 (1969).

19. W. T. Yap, B. F. Howell, and R. Schaffer, National Institute of Standards and Technology, unpublished results.

20. R. M. C. Dawson, D. C. Elliot, W. H. Elliot, and K. M. Jones, "Data for Biochemical Research" (Oxford University Press, Oxford, 1986)

21. K. Burton, Biochem. J. 143, 365 (1974).

22. M. V. Rekharsky, G. L. Gal'chenko, A. M. Egorov, and I. V. Berezin, in "Thermodynamic Data for Biochemistry and Biotechnology", edited by H. J. Hinz ( Springer-Verlag, Berlin, 1986); pp. 431-444.

23. S. Minakami and H. Yoshikawa, J. Biochem. (Tokyo) 59, 139 (1966).

24. S. A. Kuby and E. A. Noltmann in "The Enzymes", Volume 6, edited by P. D. Boyer, H. Lardy, and K. Myrbäck (Academic Press, New York, 1962) pp. 515 - 603.

25. R. A. Alberty, J. Biol. Chem., 244, 3290 (1969).

26. R. A. Alberty, Proc. Natl. Acad. Sci. U.S.A. 88, 3268 (1991).

27. R. W. Guynn, H. J. Gelberg, and R. L. Veech, J. Biol. Chem. 248, 6957 (1973).

28. R. S. Langer, C. R. Gardner, B. K. Hamilton, and C. K. Colton, AIChE J. 23, 1 (1977).

29. D. D. Wagman, W. H. Evans, V. B. Parker, R. H. Schumm, I. Halow, S. M. Bailey, K. L. Churney, and R. L. Nuttall, "The NBS Tables of Chemical Thermodynamic Properties", J. Phys. Chem. Ref. Data, 11, Supplement No. 2 (1982).

30. R. N. Goldberg and Y. B. Tewari, J. Phys. Chem. Ref. Data 18, 809 (1989).

31. R. A. Alberty and R. N. Goldberg, Biochemistry 31, 10610 (1992).

32. R. N. Goldberg, Y. B. Tewari, D. Bell, K. Fazio, and E. Anderson, "Thermodynamics of enzyme-catalyzed reactions: Part 1. Oxidoreductases," J. Phys. Chem. Ref. Data, 22, 515 (1993).

33. R. N. Goldberg and Y. B. Tewari, "Thermodynamics of enzyme-catalyzed reactions: Part 2. Transferases," J. Phys. Chem. Ref. Data, 23, 547 (1994).

34. R. N. Goldberg and Y. B. Tewari, "Thermodynamics of enzyme-catalyzed reactions: Part 3. Hydrolases," J. Phys. Chem. Ref. Data, 23, 1035 (1994).

35. R. N. Goldberg and Y. B. Tewari, "Thermodynamics of enzyme-catalyzed reactions: Part 4. Lyases," J. Phys. Chem. Ref. Data, 24, 1669 (1995).

36. R. N. Goldberg and Y. B. Tewari, "Thermodynamics of enzyme-catalyzed reactions: Part 5. Isomerases and ligases," J. Phys. Chem. Ref. Data, 24, 1765 (1995).

37. R. N. Goldberg, "Thermodynamics of enzyme-catalyzed reactions: Part 6 - 1999 update," J. Phys. Chem. Ref. Data, 28, 931 (1999).

38. T. N. Bhat, "Migration from Static to Dynamic Web Interface - Enzyme Thermodynamics Database as an Example," Proceedings of the The 9th World Multi-Conference on Systemics, Cybernetics and Informatics (WMSCI2005), July 10-13 Orlando, Florida, USA, 69 - 74.

 Back to main page