Skip to main content
Log in

The hydrophobic effect: is water afraid, or just not that interested?

  • Lecture Text
  • Published:
ChemTexts Aims and scope Submit manuscript

Abstract

Our understanding of the hydrophobic effect has advanced greatly since 1990, with the help of experimental, theoretical, and computer simulation results. The key hydrophobic signature of positive ∆C°P and negative ∆S° at room temperature has been interpreted in light of the importance of solvent cavity creation, solvent-excluded volume, and solute–water intermolecular forces, along with some unusual thermodynamic properties of pure water. Application of the hydrophobic effect to the hydration of small nonpolar solutes, protein folding stability, and protein–ligand binding is discussed in detail in this review, with an emphasis on thermodynamic analyses and interpretations.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1

Modified from ref. [3]

Fig. 2
Fig. 3
Fig. 4
Fig. 5

Data from ref. [93]

Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Data from ref. [94]

Similar content being viewed by others

Notes

  1. H and ∆S for reaction (3) are approximately zero [3].

  2. Sets of related compounds (e.g., noble gases, alkanes, benzene derivatives, and even proteins) are believed to demonstrate enthalpy and entropy convergence, at TH,conv and TS,conv, temperatures at which all members of the set have the same values of hydration (or o/w transfer) enthalpy or entropy, respectively. These temperatures are distinct from TH and TS. Further discussion can be found in the Appendix.

  3. In the older literature, some authors measured and plotted the temperature dependence of the solubility equilibrium constant (e.g., Ostwald partition coefficient) instead of ∆G° vs. T. As seen in Fig. 2b, the ln(Ksp) vs. T curve crosses the T-axis at the same points, TG,lo and TG,lo, but reaches its minimum value at TH, not TS. This occurs because the slope of the ln(Ksp) vs. T curve is zero at the minimum, and ∂ln(Ksp)/∂T  = ∂(∆G°/T)/∂T = ∆H°/T.

  4. A nonclassical hydrophobic effect (∆H-driven, rather than ∆S-driven) has been observed for large nonpolar solutes that are transferred into water from the pure solid phase (as opposed to the liquid or vapor phase) [24, 25].

  5. An American physical chemist (1881–1983) who specialized in solution chemistry and “won virtually every major prize in the field of chemistry except the Nobel Prize”.

  6. Walter Kauzmann (1916–2009), an American biophysical chemist, specialized in protein folding.

  7. It is important to note that Baldwin used the mole fraction = 1 standard state, rather than the more appropriate 1 M standard state recommended by Ben Naim [113]. TS would be much lower using the molarity standard state.

  8. An update of this plot including 100 proteins gave similar results: R2 values of 0.35 and 0.29; convergence temperatures of 63 °C for enthalpy, 60 °C for entropy (data not shown).

References

  1. Silverstein TP (1998) The real reason why oil and water don’t mix. J Chem Educ 75:116

    CAS  Google Scholar 

  2. Solomons TWG, Fryhle CB (2011) Organic chemistry, 10th edn. Wiley, Hoboken, p 81

    Google Scholar 

  3. Herzfeld J (1991) Understanding hydrophobic behavior. Science 253:88–89

    CAS  PubMed  Google Scholar 

  4. Dill KA, Privalov PL, Gill SJ, Murphy KP (1990) The meaning of hydrophobicity. Science 250:297–299

    CAS  PubMed  Google Scholar 

  5. Harris RC, Pettitt BM (2016) Reconciling the understanding of ‘hydrophobicity’with physics-based models of proteins. J Phys Condens Matter 28:083003

    PubMed  PubMed Central  Google Scholar 

  6. Yaminsky VV, Vogler EA (2001) Hydrophobic hydration. Curr Opin Colloid Interface Sci 6:342–349

    CAS  Google Scholar 

  7. Baldwin RL (2013) The new view of hydrophobic free energy. FEBS Lett 587:1062–1066

    CAS  PubMed  Google Scholar 

  8. Baldwin RL (2012) Gas–liquid transfer data used to analyze hydrophobic hydration and find the nature of the Kauzmann–Tanford hydrophobic factor. Proc Natl Acad Sci 109:7310–7313

    CAS  PubMed  Google Scholar 

  9. Baldwin RL (2013) Properties of hydrophobic free energy found by gas–liquid transfer. Proc Natl Acad Sci 110:1670–1673

    CAS  PubMed  Google Scholar 

  10. Murphy KP, Privalov PL, Gill SJ (1990) Common features of protein unfolding and dissolution of hydrophobic compounds. Science 247:559–561

    CAS  PubMed  Google Scholar 

  11. Dec SF, Gill SJ (1984) Heats of solution of gaseous hydrocarbons in water at 25 °C. J Solut Chem 13:27–41

    CAS  Google Scholar 

  12. Ben-Amotz D (2016) Water-mediated hydrophobic interactions. Annu Rev Phys Chem 67:617–638

    CAS  PubMed  Google Scholar 

  13. Murphy KP (1994) Hydration and convergence temperatures: on the use and interpretation of correlation plots. Biophys Chem 51:311–326

    CAS  Google Scholar 

  14. Tanford C (1980) Micelles: introduction. In: The hydrophobic effect: formation of micelles and biological membranes, 2nd edn. Wiley, New York, pp 42–58

  15. Lucas M, Bury R (1976) The influence of solute size on the thermodynamic parameter of transfer of a nonpolar hydrophobic solute from gas to water or from light to heavy water. J Phys Chem 80:999–1002

    CAS  Google Scholar 

  16. Chandler D (2005) Interfaces and the driving force of hydrophobic assembly. Nature 437:640–647

    CAS  PubMed  Google Scholar 

  17. Lazaridis T, Paulaitis ME (1994) Simulation studies of the hydration entropy of simple, hydrophobic solutes. J Phys Chem 98:635–642

    CAS  Google Scholar 

  18. Lum K, Chandler D, Weeks JD (1999) Hydrophobicity at small and large length scales. ACS, Washington

    Google Scholar 

  19. Ashbaugh HS, Truskett TM, Debenedetti PG (2002) A simple molecular thermodynamic theory of hydrophobic hydration. J Chem Phys 116:2907–2921

    CAS  Google Scholar 

  20. Ashbaugh HS, Pratt LR (2006) Colloquium: Scaled particle theory and the length scales of hydrophobicity. Rev Mod Phys 78:159

    CAS  Google Scholar 

  21. Ashbaugh HS (2009) Entropy crossover from molecular to macroscopic cavity hydration. Chem Phys Lett 477:109–111

    CAS  Google Scholar 

  22. Graziano G (2006) Scaled particle theory study of the length scale dependence of cavity thermodynamics in different liquids. J Phys Chem B 110:11421–11426

    CAS  PubMed  Google Scholar 

  23. Hillyer MB, Gibb BC (2016) Molecular shape and the hydrophobic effect. Annu Rev Phys Chem 67:307–329

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Matulis D, Bloomfield VA (2001) Thermodynamics of the hydrophobic effect. II. Calorimetric measurement of enthalpy, entropy, and heat capacity of aggregation of alkylamines and long aliphatic chains. Biophys Chem 93:53–65

    CAS  PubMed  Google Scholar 

  25. Matulis D (2001) Thermodynamics of the hydrophobic effect. III. Condensation and aggregation of alkanes, alcohols, and alkylamines. Biophys Chem 93:67–82

    CAS  PubMed  Google Scholar 

  26. Smith R, Tanford C (1973) Hydrophobicity of long chain n-alkyl carboxylic acids, as measured by their distribution between heptane and aqueous solutions. Proc Natl Acad Sci 70:289–293

    CAS  PubMed  Google Scholar 

  27. Solomonov BN, Sedov IA (2008) The hydrophobic effect Gibbs energy. J Mol Liq 139:89–97

    CAS  Google Scholar 

  28. Choudhury N, Pettitt BM (2006) Enthalpy–entropy contributions to the potential of mean force of nanoscopic hydrophobic solutes. J Phys Chem B 110:8459–8463

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Li IT, Walker GC (2011) Signature of hydrophobic hydration in a single polymer. Proc Natl Acad Sci 108:16527–16532

    CAS  PubMed  Google Scholar 

  30. Sharp KA, Madan B (1997) Hydrophobic effect, water structure, and heat capacity changes. J Phys Chem B 101:4343–4348

    CAS  Google Scholar 

  31. Muller N (1990) Search for a realistic view of hydrophobic effects. Acc Chem Res 23:23–28

    CAS  Google Scholar 

  32. Lee B, Graziano G (1996) A two-state model of hydrophobic hydration that produces compensating enthalpy and entropy changes. J Am Chem Soc 118:5163–5168

    CAS  Google Scholar 

  33. Godec A, Merzel F (2012) Physical origin underlying the entropy loss upon hydrophobic hydration. J Am Chem Soc 134:17574–17581

    CAS  PubMed  Google Scholar 

  34. Davis JG, Gierszal KP, Wang P, Ben-Amotz D (2012) Water structural transformation at molecular hydrophobic interfaces. Nature 491:582–585

    CAS  PubMed  Google Scholar 

  35. Grdadolnik J, Merzel F, Avbelj F (2017) Origin of hydrophobicity and enhanced water hydrogen bond strength near purely hydrophobic solutes. Proc Natl Acad Sci 114:322–327

    CAS  PubMed  Google Scholar 

  36. Matysiak S, Debenedetti PG, Rossky PJ (2011) Dissecting the energetics of hydrophobic hydration of polypeptides. J Phys Chem B 115:14859–14865

    CAS  PubMed  Google Scholar 

  37. Finney JL, Bowron DT, Daniel RM et al (2003) Molecular and mesoscale structures in hydrophobically driven aqueous solutions. Biophys Chem 105:391–409

    CAS  PubMed  Google Scholar 

  38. Silverstein TP, Slade K (2019) Effects of macromolecular crowding on biochemical systems. J Chem Educ 96:2476–2487

    CAS  Google Scholar 

  39. Ball P (2008) Water as an active constituent in cell biology. Chem Rev 108:74–108

    CAS  PubMed  Google Scholar 

  40. Brini E, Fennell CJ, Fernandez-Serra M et al (2017) How water’s properties are encoded in its molecular structure and energies. Chem Rev 117:12385–12414

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Lazaridis T (2001) Solvent size vs cohesive energy as the origin of hydrophobicity. Acc Chem Res 34:931–937

    CAS  PubMed  Google Scholar 

  42. Blokzijl W, Engberts JB (1993) Hydrophobic effects. Opinions and facts. Angew Chem Int Ed Engl 32:1545–1579

    Google Scholar 

  43. Lucas M (1976) Size effect in transfer of nonpolar solutes from gas or solvent to another solvent with a view on hydrophobic behavior. J Phys Chem 80:359–362

    CAS  Google Scholar 

  44. Lee B (1991) Solvent reorganization contribution to the transfer thermodynamics of small nonpolar molecules. Biopolym Orig Res Biomol 31:993–1008

    CAS  Google Scholar 

  45. Pratt LR, Pohorille A (1992) Theory of hydrophobicity: transient cavities in molecular liquids. Proc Natl Acad Sci 89:2995–2999

    CAS  PubMed  Google Scholar 

  46. Wallqvist A, Covell DG (1996) On the origins of the hydrophobic effect: observations from simulations of n-dodecane in model solvents. Biophys J 71:600–608

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Durell SR, Wallqvist A (1996) Atomic-scale analysis of the solvation thermodynamics of hydrophobic hydration. Biophys J 71:1695–1706

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Cerdeiriña CA, Debenedetti PG (2018) Water’s thermal pressure drives the temperature dependence of hydrophobic hydration. J Phys Chem B 122:3620–3625

    PubMed  Google Scholar 

  49. Ashbaugh HS, Bukannan H (2020) Temperature, pressure, and concentration derivatives of nonpolar gas hydration: impact on the heat capacity, temperature of maximum density, and speed of sound of aqueous mixtures. J Phys Chem B 124:6924–6942

    CAS  PubMed  Google Scholar 

  50. Graziano G (2003) Comment on “A simple molecular thermodynamic theory of hydrophobic hydration”[J. Chem. Phys. 116, 2907 (2002)]. J Chem Phys 119:10448–10449

    CAS  Google Scholar 

  51. Cabani S, Gianni P, Mollica V, Lepori L (1981) Group contributions to the thermodynamic properties of non-ionic organic solutes in dilute aqueous solution. J Solut Chem 10:563–595

    CAS  Google Scholar 

  52. Fisicaro E, Compari C, Braibanti A (2004) Entropy/enthalpy compensation: hydrophobic effect, micelles and protein complexes. Phys Chem Chem Phys 6:4156–4166

    CAS  Google Scholar 

  53. Graziano G (2009) Hydration entropy of polar, nonpolar and charged species. Chem Phys Lett 479:56–59

    CAS  Google Scholar 

  54. Graziano G (2019) Contrasting the hydration thermodynamics of methane and methanol. Phys Chem Chem Phys 21:21418–21430

    CAS  PubMed  Google Scholar 

  55. Edsall JT (1935) Apparent molal heat capacities of amino acids and other organic compounds. J Am Chem Soc 57:1506–1507

    CAS  Google Scholar 

  56. Silverstein KA, Haymet ADJ, Dill KA (2000) The strength of hydrogen bonds in liquid water and around nonpolar solutes. J Am Chem Soc 122:8037–8041

    CAS  Google Scholar 

  57. Klotz IM (1958) Protein hydration and behavior. Science 128:815–822

    CAS  PubMed  Google Scholar 

  58. Kauzmann W (1959) Some factors in the interpretation of protein denaturation. In: Advances in protein chemistry. Elsevier, Amsterdam, pp 1–63

  59. Kozak JJ, Knight WS, Kauzmann W (1968) Solute–solute interactions in aqueous solutions. J Chem Phys 48:675–690

    CAS  Google Scholar 

  60. Némethy G, Scheraga HA (1962) Structure of water and hydrophobic bonding in proteins. I. A model for the thermodynamic properties of liquid water. J Chem Phys 36:3382–3400

    Google Scholar 

  61. Némethy G, Scheraga HA (1962) Structure of water and hydrophobic bonding in proteins. II. Model for the thermodynamic properties of aqueous solutions of hydrocarbons. J Chem Phys 36:3401–3417

    Google Scholar 

  62. Tanford C (1962) Contribution of hydrophobic interactions to the stability of the globular conformation of proteins. J Am Chem Soc 84:4240–4247

    CAS  Google Scholar 

  63. Tanford C (1978) The hydrophobic effect and the organization of living matter. Science 200:1012–1018

    CAS  PubMed  Google Scholar 

  64. Hildebrand JH (1979) Is there a “hydrophobic effect”? Proc Natl Acad Sci USA 76:194

    CAS  PubMed  Google Scholar 

  65. Gill SJ, Dec SF, Olofsson G, Wadsö I (1985) Anomalous heat capacity of hydrophobic solvation. J Phys Chem 89:3758–3761

    CAS  Google Scholar 

  66. Head-Gordon T (1995) Is water structure around hydrophobic groups clathrate-like? Proc Natl Acad Sci 92:8308–8312

    CAS  PubMed  Google Scholar 

  67. Xu H, Dill K (2005) Water’s hydrogen bonds in the hydrophobic effect: a simple model. J Phys Chem B 109:23611–23617

    CAS  PubMed  Google Scholar 

  68. Rezus YLA, Bakker HJ (2007) Observation of immobilized water molecules around hydrophobic groups. Phys Rev Lett 99:148301

    CAS  PubMed  Google Scholar 

  69. Irudayam SJ, Henchman RH (2010) Solvation theory to provide a molecular interpretation of the hydrophobic entropy loss of noble-gas hydration. J Phys: Condens Matter 22:284108

    Google Scholar 

  70. Ashbaugh HS, Barnett JW, Saltzman A et al (2016) Communication: stiffening of dilute alcohol and alkane mixtures with water. J Chem Phys 145:201102

    PubMed  Google Scholar 

  71. Wu X, Lu W, Streacker LM et al (2018) Methane hydration-shell structure and fragility. Angew Chem Int Ed 57:15133–15137

    CAS  Google Scholar 

  72. Galamba N (2013) Water’s structure around hydrophobic solutes and the iceberg model. J Phys Chem B 117:2153–2159

    CAS  PubMed  Google Scholar 

  73. Baldwin RL (2014) Dynamic hydration shell restores Kauzmann’s 1959 explanation of how the hydrophobic factor drives protein folding. Proc Natl Acad Sci 111:13052–13056

    CAS  PubMed  Google Scholar 

  74. Meister K, Strazdaite S, DeVries AL et al (2014) Observation of ice-like water layers at an aqueous protein surface. Proc Natl Acad Sci 111:17732–17736

    CAS  PubMed  Google Scholar 

  75. Brotzakis ZF, Voets IK, Bakker HJ, Bolhuis PG (2018) Water structure and dynamics in the hydration layer of a type III anti-freeze protein. Phys Chem Chem Phys 20:6996–7006

    CAS  PubMed  Google Scholar 

  76. Kim J, Tian Y, Wu J (2015) Thermodynamic and structural evidence for reduced hydrogen bonding among water molecules near small hydrophobic solutes. J Phys Chem B 119:12108–12116

    CAS  PubMed  Google Scholar 

  77. Graziano G (2014) Comment on “Water’s structure around hydrophobic solutes and the Iceberg Model”. J Phys Chem B 118:2598–2599

    CAS  PubMed  Google Scholar 

  78. Dec SF, Bowler KE, Stadterman LL et al (2006) Direct measure of the hydration number of aqueous methane. J Am Chem Soc 128:414–415

    CAS  PubMed  Google Scholar 

  79. Buchanan P, Aldiwan N, Soper AK et al (2005) Decreased structure on dissolving methane in water. Chem Phys Lett 415:89–93

    CAS  Google Scholar 

  80. Kinoshita M (2008) Molecular origin of the hydrophobic effect: analysis using the angle-dependent integral equation theory. J Chem Phys 128:01B612

    Google Scholar 

  81. Islam N, Flint M, Rick SW (2019) Water hydrogen degrees of freedom and the hydrophobic effect. J Chem Phys 150:014502

    PubMed  Google Scholar 

  82. Bergman DL, Lynden-Bell RM (2001) Is the hydrophobic effect unique to water? The relation between solvation properties and network structure in water and modified water models. Mol Phys 99:1011–1021

    CAS  Google Scholar 

  83. Lee B (1994) Enthalpy-entropy compensation in the thermodynamics of hydrophobicity. Biophys Chem 51:271–278

    CAS  PubMed  Google Scholar 

  84. Madan B, Lee B (1994) Role of hydrogen bonds in hydrophobicity: the free energy of cavity formation in water models with and without the hydrogen bonds. Biophys Chem 51:279–289

    CAS  PubMed  Google Scholar 

  85. Bernal JD (1939) Structure of proteins. Nature 143:663–667

    Google Scholar 

  86. Dill KA (1990) Dominant forces in protein folding. Biochemistry 29:7133–7155

    CAS  PubMed  Google Scholar 

  87. Livingstone JR, Spolar RS, Record MT Jr (1991) Contribution to the thermodynamics of protein folding from the reduction in water-accessible nonpolar surface area. Biochemistry 30:4237–4244

    CAS  PubMed  Google Scholar 

  88. Pace CN, Fu H, Fryar KL et al (2011) Contribution of hydrophobic interactions to protein stability. J Mol Biol 408:514–528

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Pace CN, Shirley BA, McNutt M, Gajiwala K (1996) Forces contributing to the conformational stability of proteins. FASEB J 10:75–83

    CAS  PubMed  Google Scholar 

  90. Harano Y, Kinoshita M (2005) Translational-entropy gain of solvent upon protein folding. Biophys J 89:2701–2710

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Lesser GJ, Rose GD (1990) Hydrophobicity of amino acid subgroups in proteins. Proteins Struct Funct Bioinform 8:6–13

    CAS  Google Scholar 

  92. Li L, Li C, Zhang Z, Alexov E (2013) On the dielectric “constant” of proteins: smooth dielectric function for macromolecular modeling and its implementation in DelPhi. J Chem Theory Comput 9:2126–2136

    CAS  PubMed  PubMed Central  Google Scholar 

  93. Rees DC, Robertson AD (2001) Some thermodynamic implications for the thermostability of proteins. Protein Sci 10:1187–1194

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Pica A, Graziano G (2016) Shedding light on the extra thermal stability of thermophilic proteins. Biopolymers 105:856–863

    CAS  PubMed  Google Scholar 

  95. Graziano G (2014) On the mechanism of cold denaturation. Phys Chem Chem Phys 16:21755–21767

    CAS  PubMed  Google Scholar 

  96. Graziano G (2015) The Gibbs energy cost of cavity creation depends on geometry. J Mol Liq 211:1047–1051

    CAS  Google Scholar 

  97. Graziano G (2019) Some clues on the conformational stability of globular proteins. Trends Pept Protein Sci 4:1–1

    Google Scholar 

  98. Graziano G (2008) Is there a relationship between protein thermal stability and the denaturation heat capacity change? J Therm Anal Calorim 93:429–438

    CAS  Google Scholar 

  99. Sawle L, Ghosh K (2011) How do thermophilic proteins and proteomes withstand high temperature? Biophys J 101:217–227

    CAS  PubMed  PubMed Central  Google Scholar 

  100. Razvi A, Scholtz JM (2006) A thermodynamic comparison of HPr proteins from extremophilic organisms. Biochemistry 45:4084–4092

    CAS  PubMed  Google Scholar 

  101. Privalov PL, Gill SJ (1988) Stability of protein structure and hydrophobic interaction. In: Advances in protein chemistry. Elsevier, Amsterdam, pp 191–234

  102. Robertson AD, Murphy KP (1997) Protein structure and the energetics of protein stability. Chem Rev 97:1251–1268

    CAS  PubMed  Google Scholar 

  103. Seong S-Y, Matzinger P (2004) Hydrophobicity: an ancient damage-associated molecular pattern that initiates innate immune responses. Nat Rev Immunol 4:469–478

    CAS  PubMed  Google Scholar 

  104. Ohtaka H, Schön A, Freire E (2003) Multidrug resistance to HIV-1 protease inhibition requires cooperative coupling between distal mutations. Biochemistry 42:13659–13666

    CAS  PubMed  Google Scholar 

  105. Biela A, Sielaff F, Terwesten F et al (2012) Ligand binding stepwise disrupts water network in thrombin: enthalpic and entropic changes reveal classical hydrophobic effect. J Med Chem 55:6094–6110

    CAS  PubMed  Google Scholar 

  106. Fernández-Vidal M, White SH, Ladokhin AS (2011) Membrane partitioning: “classical” and “nonclassical” hydrophobic effects. J Membr Biol 239:5–14

    PubMed  Google Scholar 

  107. Ross PD, Subramanian S (1981) Thermodynamics of protein association reactions: forces contributing to stability. Biochemistry 20:3096–3102

    CAS  Google Scholar 

  108. Bingham RJ, Findlay JB, Hsieh S-Y et al (2004) Thermodynamics of binding of 2-methoxy-3-isopropylpyrazine and 2-methoxy-3-isobutylpyrazine to the major urinary protein. J Am Chem Soc 126:1675–1681

    CAS  PubMed  Google Scholar 

  109. Meyer EA, Castellano RK, Diederich F (2003) Interactions with aromatic rings in chemical and biological recognition. Angew Chem Int Ed 42:1210–1250

    CAS  Google Scholar 

  110. Syme NR, Dennis C, Phillips SE, Homans SW (2007) Origin of heat capacity changes in a “nonclassical” hydrophobic interaction. ChemBioChem 8:1509–1511

    CAS  PubMed  PubMed Central  Google Scholar 

  111. Setny P, Baron R, McCammon JA (2010) How can hydrophobic association be enthalpy driven? J Chem Theory Comput 6:2866–2871

    CAS  PubMed  PubMed Central  Google Scholar 

  112. Privalov PL (1979) Stability of proteins small globular proteins. In: Advances in protein chemistry. Elsevier, Amsterdam, pp 167–241

  113. Ben-Naim A (1978) Standard thermodynamics of transfer. Uses and misuses. J Phys Chem 82:792–803

    CAS  Google Scholar 

  114. Murphy KP, Gill SJ (1990) Group additivity thermodynamics for dissolution of solid cyclic dipeptides into water. Thermochim Acta 172:11–20

    CAS  Google Scholar 

  115. Murphy KP, Gill SJ (1991) Solid model compounds and the thermodynamics of protein unfolding. J Mol Biol 222:699–709

    CAS  PubMed  Google Scholar 

  116. Baldwin RL (1986) Temperature dependence of the hydrophobic interaction in protein folding. Proc Natl Acad Sci 83:8069–8072

    CAS  PubMed  Google Scholar 

  117. Garde S, Hummer G, Garcia AE et al (1996) Origin of entropy convergence in hydrophobic hydration and protein folding. Phys Rev Lett 77:4966

    CAS  PubMed  Google Scholar 

  118. Lee B, Richards FM (1971) The interpretation of protein structures: estimation of static accessibility. J Mol Biol 55:379–384

    CAS  PubMed  Google Scholar 

  119. Chothia C (1984) Principles that determine the structure of proteins. Annu Rev Biochem 53:537–572

    CAS  PubMed  Google Scholar 

  120. Bello J (1978) Tight packing of protein cores and interfaces: relation to conservative amino acid sequences and stability of protein–protein interaction. Int J Pept Protein Res 12:38–41

    CAS  PubMed  Google Scholar 

  121. Huang DM, Chandler D (2000) Temperature and length scale dependence of hydrophobic effects and their possible implications for protein folding. Proc Natl Acad Sci 97:8324–8327

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

I wish to express profound gratitude to Giuseppe Graziano and Henry Ashbaugh, both of whom have helped me to grapple with the recent literature, and patiently answered my many questions. Bob Gennis also read over the first draft and gave me some cogent editing suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Todd P. Silverstein.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix

Table 8 Unfolding thermodynamic parameters for curves plotted in Figs. 7 and 8

Enthalpy, entropy convergence for hydrophobic hydration

In his massive and influential 1979 review of protein stability [112], one of the observations that Privalov made was that at about 100–110 °C, compact globular proteins converged on the same value of specific enthalpy of unfolding (cal/g). He found a similar convergence for the specific entropy of unfolding around 110 °C, although the effect was not nearly as clear.

In 1986, Robert Baldwin published a keystone paper [116] in which he compared the thermodynamics of liquid hydrocarbon neat transfer to the unfolding of the enzyme lysozyme. He found that for six hydrocarbons (benzene, toluene, ethylbenzene, cyclohexane, n-pentane, and n-hexane), TH and TS were approximately the same: TH = 22 ± 5 °C, and TS = 113 ± 3 °C.Footnote 7 Baldwin then applied these values to the temperature-dependent unfolding of hen lysozyme by splitting protein stability into a hydrophobic effect (HE) and a non-hydrophobic effect (NHE, ∆Xobs = ∆XHE + ∆XNHE), and assuming that (1) ∆C°P(U) is entirely due to the HE, and (2) for the effect of hydrophobicity on protein unfolding, TH and TS are identical to the liquid hydrocarbon neat transfer values. Using Eqs. (4) and (6) (main text), Baldwin calculated HE ∆H°298 and ∆S°298, and then by subtracting from the net observed ∆H°298 and ∆S°298, he calculated the NHE values. Both NHE values were temperature-independent: ∆H°298(U, NHE) =  + 217.8 ± 0.7 kJ/mol, and ∆S°298(U, NHE) =  + 2294 ± 3 J/K/mol. The former value was interpreted as the cost, upon unfolding, of breaking polar interactions (e.g., H-bonds), and the latter the increase in conformational entropy, i.e., the release of polypeptide chain positioning constraints.

In 1990, Murphy and Gill took this analysis a step further, defining the notion of “convergence” temperatures for enthalpy and entropy [10, 13, 114, 115]. They started by applying Baldwin’s HE/NHE breakdown to the temperature dependence of enthalpy, considering a series of homologous compounds (e.g., noble gases, alkanes, alcohols, amines) which shared the same functional group and differed only in their hydrophobicity (e.g., size of noble gas or number of carbons). The hydration enthalpy can then be envisioned as comprising a hydrophobic term that increases with solute hydrophobicity, and a non-hydrophobic term that is independent of hydrophobicity. Enthalpy convergence would then occur at some temperature, TH,conv, at which all of the homologous compounds in the series have the same hydration enthalpy, ∆Hconv. Thus, at the convergence temperature, ∆H° of hydration depends only on the nature of the series (i.e., the functional group), and not on the individual compound. With this in mind, the temperature dependence of ∆H (main text Eq. 2), can be cast as

$$ \begin{aligned} \Delta H^\circ (T) & = \Delta H^\circ (T_{{{\text{ref}}}} ) \, + \, \Delta C^\circ_{P} (T{-}T_{{{\text{ref}}}} ) \\ & = \Delta H^\circ (T_{{H{\text{,conv}}}} ) \, + \, \Delta C^\circ_{P} (T{-}T_{{H,{\text{conv}}}} ). \\ \end{aligned} $$
(19)

Normally, one would use Eq. (19) along with measured values of ∆H° at various temperatures to determine ∆C°P. If, however, one already knows ∆C°P for each compound in the series, then measured values of ∆H° at a specific reference temperature (e.g., 298 K) can be used in the following version of Eq. (19):

$$ \Delta H^\circ_{298} = \Delta H^\circ (T_{{H,{\text{conv}}}} ) + \Delta C^\circ_{P} (298 \, {-}T_{{H,{\text{conv}}}} ). $$
(20)

Thus, a plot of ∆H°298 vs. ∆C°P for each compound in the series will give a straight line with slope = (298 − Tconv) and intercept = ∆H°(Tconv). This is sometimes referred to as an MPG (Murphy–Privalov–Gill) plot. Because ∆H°(Tconv) characterizes the entire series, it must apply to what remains the same in the series, i.e., the non-hydrophobic component of the observed net ∆H°; thus, it would be different for a series of alkanes vs. alcohols vs. amines. On the other hand, Tconv should be the same for alkanes, alcohols, and amines because they all differ by the same hydrophobic parameter, i.e., the number of carbons.

A similar analysis of entropy leads to this adaptation of Eq. (3) in the main text:

$$ \Delta S^\circ_{298} = \Delta S^\circ (T_{{S,{\text{conv}}}} ) + \Delta C^\circ_{P} \cdot \ln (298/T_{{S,{\text{conv}}}} ). $$
(21)

The MPG plot here would be ∆S°298 vs. ∆C°P, with slope = ln(298/TS,conv), and intercept = ∆S°(TS,conv). As with enthalpy, ∆S°(Tconv) characterizes the entire series, so it must apply to the non-hydrophobic component of the observed net ∆S°.

For example, MPG plots of Baldwin’s liquid hydrocarbon neat transfer data [116] do show a linear correlation (Fig. 11), although the relationship is not very strong for ∆H (R2 = 0.6). From the slopes, the convergence temperatures are calculated to be TS,conv = 106 ± 5 °C; TH,conv = 40 ± 6 °C. Note that these temperatures are close to, but not identical to TS and TH.

Fig. 11
figure 11

Data from ref. [116]

MPG plots of liquid hydrocarbon neat transfer. a Entropy; b enthalpy; liquids are benzene, toluene, ethylbenzene, cyclohexane, n-pentane, and n-hexane

Figure 11a, with its excellent linearity, suggests that entropy convergence for the neat transfer of this series of liquid hydrocarbons does occur. The relative uncertainty in the convergence temperature is good, at only 4% (= 100 × 4.7/106). (For ∆S°(TS,conv), however, it is 70%!) Thus, we might expect to see a fairly robust isosbestic point in the ∆S° vs. T plot for neat transfer (Fig. 12). Sadly, that is not actually the case.

Fig. 12
figure 12

Data from ref. [116]

Temperature dependence of the entropy of neat transfer of liquid hydrocarbons

The first thing to note from Fig. 12 is that for all six liquids, TS, the temperature at which each curve crosses the T-axis, lies between 382 and 390 K; this matches the average value reported above (113 ± 3 °C). Most of the curves cross each other (shared values of ∆S°) between about 360 and 390 K; this also matches the calculated convergence temperature from the MPG plot, 106 ± 5 °C. However, there are several intersections below 360 K, and several more above 400 K. The lack of a clear isosbestic point suggests that the entropic convergence behavior is not nearly as robust as one might have expected from the MPG plot. A similar “smeared” isosbestic is observed in the ∆H° vs. T curve intersections, stretching from 300 to 345 K (data not shown). At least some of this “blurring” of the expected isosbestic point has been shown to be due to the temperature dependence of ∆CP [117]. Beyond this though, one may question the assumptions in the convergence derivation, namely, that the hydrophobic and non-hydrophobic contributions can be separated out and that the former really are identical for every member of the homologous series of compounds. Graziano [32] and Pratt [117] have supported the existence of entropy convergence for families of small solutes, but rejected enthalpy convergence (Graziano, personal communication).

Murphy, Privalov, and Gill published an extremely influential paper in 1990 in which they compared MPG plots of Baldwin’s liquid hydrocarbon data with nonpolar gas hydration, solid → water transfer of nonpolar cyclic peptides, and protein unfolding [10]. They found that the slopes of all four of the ∆S° vs. ∆C°P plots were nearly identical, ranging from − 0.23 ± 0.04 to − 0.28 ± 0.01. Thus TS,conv for all four of these sets of compounds lie in a fairly narrow range between 102 ± 15 and 121 ± 6 °C; this pointed to the “dominant role that water [and the hydrophobic effect] play in determining the [entropy] of hydration of [all of] these compounds” [10].

Murphy, Privalov, and Gill drew another interesting conclusion from the intercepts of their MPG plots [10]. They determined ∆S°(TS,conv) to be − 78.5 ± 2.5 J/K/mol for nonpolar gas hydration vs. − 6 ± 4 J/K/mol for liquid neat transfer; this is expected, because gases lose much more freedom of motion upon transfer to the aqueous phase than liquids do. Meanwhile, ∆S°(TS,conv) was quite similar for protein unfolding (18 ± 1 J/K/mol) and the transfer of solid nonpolar cyclic peptides into water (16 ± 1 J/K/mol). This corroborated previous findings that in terms of packing and freedom of motion, the protein interior is best modeled as a solid organic compound, rather than a liquid [118,119,120].

To give some idea of the linearity of the MPG plots in the 1990 Science paper, Fig. 13 plots the nonpolar gas hydration and protein unfolding results. Note that once again, the linear fits are good (R2 = 0.89 for gas hydration, 0.96 for protein unfolding), and the slopes are fairly close (− 0.068 and − 0.077). In light of how influential the 1990 Murphy, Privalov, and Gill paper was, it is surprising that no one seems to have checked to see whether these gases and proteins actually demonstrated isosbestic behavior in their ∆H° vs. T plots. In Fig. 14 I have plotted the five members of each series that lie closest to the linear fit line in the MPG plots. As with liquid hydrocarbon neat transfer (Fig. 12), many of the crossings lie in a fairly narrow range of temperature (355–365 K for A and 370–380 K for B), but a number of crossings occur far outside this range.

Fig. 13
figure 13

MPG enthalpy plots of a nonpolar gas hydration; and b protein unfolding; data from ref. [101]. Without giving any reason, MPG omitted the parvalbumin unfolding data point (red) from their published plot

Fig. 14
figure 14

Data from ref. [101]

Temperature dependence of the enthalpy of a nonpolar gas hydration, and b protein unfolding

Perhaps the most interesting part of this story about convergence came to light in the years after 1990, as more protein unfolding thermodynamic data became available. Twelve proteins were originally tabulated in Privalov and Gill’s 1988 review [101], 11 of which were included in their 1990 MPG plot [10]. It turned out that not only was the omitted point (suspiciously?) far off the fit line (Fig. 14b), but the 11 points included were extraordinarily unusual in their linearity.

This was discovered in 1997, when Robertson and Murphy published a review in which they tabulated unfolding thermodynamic data for 65 globular proteins of known structure [102]. The MPG plots in this paper had R2 values of only 0.36 and 0.33; unsurprisingly, the convergence temperatures calculated from the slopes of these plots (66 °C for enthalpy, 65 °C for entropy) were much different from those in the 1990 paper (102 °C for enthalpy, 106 °C for entropy).Footnote 8 Robertson and Murphy’s conclusion from these disappointing MPG plots was that the convergence behavior for this much larger set of unfolded proteins was “not very compelling.” Huang and Chandler [121], using Lum–Chandler–Weeks theory [18], showed that the convergence temperature should decrease with increasing protein size, thus “one may not expect to observe a convergence temperature for the entropy of unfolding for all proteins.” Considering the diversity of protein groups that are exposed to water upon unfolding, as well as the change in shape, it is not surprising that proteins do not behave like a set of homologous compounds all with the same functional group. (And as we have seen, the convergence behavior even in this best of cases is not as clear as one might hope.) This failure to observe convergence behavior in the larger data sets has been recognized by some authors [20], but has been ignored by others [99].

It is important to conclude this discussion by stressing that enthalpy and entropy convergence do not generally exist for protein unfolding, and the existence of enthalpy convergence for hydrophobic hydration can be questioned as well.

Entropy is a measure of freedom of motion

One way to measure entropy is given by Boltzmann’s entropy equation,

$$ S = R \cdot \ln (W), $$
(22)

where R is the gas law constant, 8.314 J/K/mol, and W is the number of ways that a system can be arranged. For an ideal gas, W can be a measure of positions within a volume of space available to the particle. Excluding the particle from a portion of the volume will lower W and thus decrease entropy.

We can demonstrate this by combining the first and second laws of thermodynamics to get, for a fully reversible reaction,

$$ {\text{d}}U\left( {{\text{internal}}\;{\text{energy}}} \right) = T \cdot {\text{d}}S - P \cdot {\text{d}}V. $$
(23)

Furthermore, since dU = CV·dT, Eq. (23) becomes

$$ {\text{d}}S = C_{V} \left( {\frac{{{\text{d}}T}}{T}} \right) + \left( \frac{P}{T} \right){\text{d}}V. $$
(24)

Using the ideal gas law (PV = nRT P/T = R/\(\overline{V}\)), we get

$$ {\text{d}}S = C_{V} \left( {\frac{{{\text{d}}T}}{T}} \right) + R \cdot \frac{{{\text{d}}V}}{{\overline{V}}}. $$
(25)

For an isothermal process, dT = 0, so

$$ {\text{d}}S = R \cdot \frac{{{\text{d}}V}}{{\overline{V}}}. $$
(26)

For the reversible isothermal compression of an ideal gas, dV is negative and thus dS would be negative as well. In other words, restricting the available space curtails freedom of motion, which in turn lowers entropy.

Although the situation is more complex in the liquid (and solid) phase, the general conclusion remains the same: excluding solvent from a volume of space (e.g., a cavity) decreases the entropy of the solvent.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Silverstein, T.P. The hydrophobic effect: is water afraid, or just not that interested?. ChemTexts 6, 26 (2020). https://doi.org/10.1007/s40828-020-00117-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s40828-020-00117-8

Keywords

Navigation