The hydrophobic effect, and fluctuations: The long and the short of it COMMENTARY The hydrophobic effect, and fluctuations: The long and the short of it Erte Xia and Amish J. Patela,1 The hydrophobic effect, which is used to describe the aversion of oil for water or the affinity of oily objects for one another in water, plays an important role in diverse disciplines (1). For example, by segregating to the oil–water interface, amphiphilic molecules that possess both hydrophobic and hydrophilic groups can mitigate unfavorable oil–water interactions, thereby stabilizing emulsions and facilitating detergency. Be- cause roughly half the amino acids, which form the basic building blocks of proteins, are hydrophobic, the hydrophobic effect also plays a central role in bio- physics. Owing to its ubiquity and its multifaceted nature (2, 3), being able to accurately model the hydrophobic effect is both important and challenging. In PNAS, Vaikuntanathan et al. (4) provide important insights into the essential ingredients required in a minimal model of the hydrophobic effect. The hydrophobic effect characteristically mani- fests itself in very different ways at microscopic and macroscopic length scales (1). Macroscopically, the aversion of oil and water for one another is captured by the large surface tension associated with the oil– water interface. At this scale, the hydrophobic effect drives the minimization of the unfavorable interfacial area (for example, by the coalescence of oil droplets in water). Although the macroscopic hydrophobic effect is governed by the relatively straightforward physics of interfaces, it can nevertheless lead to complex phe- nomena, such as the nanobubble-mediated long- ranged forces between extended hydrophobic sur- faces (5), or the assembly of anisotropic particles at curved interfaces (6). Interestingly, the thermody- namic driving force for such hydrophobic assembly, which is dictated by interfacial tension, decreases with increasing temperature. In contrast, at the molecular scale, the driving force for hydrophobic assembly famously increases with increasing temperature (1). In fact, whenever biomo- lecular assembly occurs upon increasing temperature, in seeming disregard for entropic considerations, the hydrophobic effect is often the first suspect. A classic example is the cold denaturation of proteins. Whereas the denaturation or unfolding of proteins upon heating is common, and is ascribed to the configura- tional entropy of the macromolecule prevailing over the favorable energy of the folded state, certain pro- teins also denature upon cooling. Such cold denaturation is explained by the fact that the molecular hydropho- bic effect, which stabilizes the folded state of the pro- tein, is weakened upon cooling, causing the protein to unravel (7). To understand these contrasting manifestations of the hydrophobic effect, it is instructive to consider how water structure is perturbed near small (molecu- lar) and large (macroscopic) hydrophobic solutes. Near a macroscopic hydrophobic object, such as an oil droplet, interfacial water molecules are unable to participate in four hydrogen bonds like their bulk counterparts (8). This energetically unfavorable sce- nario is responsible for the large oil–water surface ten- sion and drives the minimization of interfacial area. In the vicinity of molecular hydrophobes such as meth- ane, the hydration waters are able to maintain all their hydrogen bonds, but accommodating the hydro- phobe severely constrains the hydrogen bonding net- work of water (1). This strain on the hydrogen bonding network is reflected in a large negative entropy of hydration; it is this entropic penalty that causes molec- ular hydrophobes to assemble more strongly as tem- perature is increased. Thus, hydrophobic hydration and assembly at the molecular level are not governed by interfacial physics, but are fundamentally different. Given these differences in hydrophobic hydration in the molecular and macroscopic limits, at what size do we cross over from one regime to the other? Theory, simulation, and experiment all agree that the cross-over occurs for solutes that are about 1 nm in size (2, 9–13). This length scale emerges from a prod- uct of the surface tension of water and its isothermal compressibility, properties that respectively influence the hydration of large and small solutes (1, 10, 12). The 1-nm length scale also plays a key role in numerous molecular assemblies; for example, most amino acid side chain residues are smaller than 1 nm, but folded proteins are larger than 1 nm. Being able to quantita- tively capture the length scale cross-over in hydrophobic aDepartment of Chemical and Biomolecular Engineering, University of Pennsylvania, Philadelphia, PA 19104 Author contributions: E.X. and A.J.P. wrote the paper. The authors declare no conflict of interest. See companion article on page E2224 in issue 16 of volume 113. 1To whom correspondence should be addressed. Email: amish.patel@seas.upenn.edu. www.pnas.org/cgi/doi/10.1073/pnas.1603014113 PNAS | April 26, 2016 | vol. 113 | no. 17 | 4549–4551 C O M M E N T A R Y D o w n lo a d e d a t C a rn e g ie M e llo n U n iv e rs ity o n A p ri l 5 , 2 0 2 1 http://crossmark.crossref.org/dialog/?doi=10.1073/pnas.1603014113&domain=pdf mailto:amish.patel@seas.upenn.edu www.pnas.org/cgi/doi/10.1073/pnas.1603014113 hydration is thus critical for understanding biophysical phenom- ena such as protein folding and aggregation, and must be a central requirement of any comprehensive theory of the hydrophobic effect. In a seminal contribution, Lum, Chandler, and Weeks (LCW) introduced the key theoretical principles necessary for reconciling hydrophobic hydration at small and large length scales (2). The LCW theory of hydrophobicity prescribed decomposing the spa- tial variations in water density, which arise from the presence of a hydrophobic solute, into two separate components. The first com- ponent, ρlðrÞ, is governed by interfacial physics and displays only gentle variations over molecular length scales; LCW used a Landau– Ginzburg density functional to describe the free energetics of this component. The second component, ρsðrÞ, incorporates the small-amplitude, short-wavelength fluctuations in molecular density. Importantly, the statistics of such small fluctuations in water density are Gaussian, a fact that underpins liquid-state theories (14, 15) and has been confirmed by molecular simulations (16). LCW theory leverages the Gaussian statistics of the small length scale fluctuations to integrate out ρsðrÞ, thereby providing a renormalized and coarse-grained description of the free ener- getics of ρlðrÞ. In particular, ρlðrÞ remains close to its bulk liquid value, ρbulk, within and near molecular solutes, because the small ρsðrÞ fluctuations are sufficient to vacate the region occupied by the solute. In contrast, vacating waters from a region larger than 1 nm (for the solute to occupy) requires nucleation of an interface, with ρlðrÞ vanishing within the solute region and approaching ρbulk outside. The original LCW paper used a variational formalism to evaluate ρlðrÞ in the mean-field limit (2). To incorporate fluctua- tions in ρlðrÞ, subsequent LCW-inspired treatments (17, 18) used simple models capable of capturing interfacial physics, such as the lattice gas model. In the lattice gas model, variations in density between adjacent lattice sites carry an energetic penalty, «, which along with the lattice spacing, δ, determines the interfacial tension, γ. However, the formation of a water–vapor interface at ambient conditions is associated not only with an energetic penalty, but also with entropically favorable interfacial fluctuations (Fig. 1A); indeed, the statistics of such long-wavelength capillary wave fluc- tuations are well understood (19). Vaikuntanathan and Geissler (20) recognized that for the lattice gas model to appropriately capture capillary wave fluctuations, only a small range of « values could be used. If « is too small, we approach the critical point, and the bulk liquid and vapor phases display large density fluctua- tions; however, if « is too large, interfacial fluctuations are sup- pressed. With the optimally parameterized lattice gas for ρlðrÞ, the authors’ LCW-inspired model was able to quantitatively capture, without any adjustable parameters, the length scale cross-over in the hydration free energies of idealized hydrophobes of various shapes. Remarkably, their model was also able to capture subtle features of the interfacial free energetics, such as a preference for droplets over bubbles (20, 21), highlighting the importance of capillary fluctua- tions even for nanoscale interfaces (Fig. 1 B and C). In their recent study, Vaikuntanathan et al. (4) subject their LCW-inspired model to another stringent test, that of quantifying the statistics of water density fluctuations; such fluctuations, char- acterized by the probability, PvðNÞ, of observing N waters in a probe volume, v, have provided numerous insights into the mul- tifaceted nature of the hydrophobic effect (3, 16, 23, 24). For example, low-N fluctuations are significantly enhanced near a hy- drophobic surface, situating the interfacial waters at the edge of dewetting transition (Fig. 1D) and making them sensitive to pertur- bations (2, 24). Using their LCW-inspired model, Vaikuntanathan et al. (4) obtain not just qualitative, but also quantitative, agreement with molecular simulations, in their estimates of PvðNÞ for probe volumes in bulk water, at interfaces, and in hydrophobic confinement. Furthermore, this quantitative agree- ment is lost if either the small length-scale Gaussian fluctuations or the large length-scale interfacial fluctuations are not included in the model. By identifying capillary fluctuations as an essential ingredient of a minimal model of the hydrophobic effect, Vaikuntanathan et al. (4) have laid the groundwork for studying aqueous systems using a coarse-grained representation of water, which has the accuracy of molecular simulations but exacts a tiny fraction of the computational cost. Given the importance of solvent fluctua- tions in modulating dewetting barriers and the kinetics of hydro- phobic interactions (25–27), their model could conceivably be used to study not only the thermodynamics, but also the kinetics of complex self-assemblies. Another important application is the study of biomolecules, which are too large to model using explicit water. Along with hydrophobic interactions, electrostatic interac- tions also play an important role in the biomolecular context. We hope that the success of LCW-inspired models motivates the de- velopment of minimal models that efficiently incorporate fluctua- tions in both the short- and long-wavelength water polarization response to charged moieties (28, 29). Acknowledgments This work was supported by National Science Foundation Grants UPENN MRSEC DMR 11-20901 and CBET 1511437, and the Charles E. Kaufman Foundation Grant KA2015-79204. A B C D Fig. 1. Long-wavelength interfacial fluctuations can be seen in simulation snapshots of (A) the liquid–vapor interface, (B and C) nanoscopic vapor bubbles in bulk water, and (D) at a hydrophobic surface. The work of Vaikuntanathan et al. (4) and Vaikuntanathan and Geissler (20) highlights the importance of such fluctuations and provides a way to incorporate them into a minimal model of the hydrophobic effect. The average interface in each case is either flat (A and D) or spherical (B and C), and the color represents deviation away from the average interface toward (red) or away (blue) from water (cyan). The instantaneous interfaces were computed using the algorithm outlined in ref. 22. 4550 | www.pnas.org/cgi/doi/10.1073/pnas.1603014113 Xi and Patel D o w n lo a d e d a t C a rn e g ie M e llo n U n iv e rs ity o n A p ri l 5 , 2 0 2 1 www.pnas.org/cgi/doi/10.1073/pnas.1603014113 1 Chandler D (2005) Interfaces and the driving force of hydrophobic assembly. Nature 437(7059):640–647. 2 Lum K, Chandler D, Weeks JD (1999) Hydrophobicity at small and large length scales. J Phys Chem B 103:4570–4577. 3 Patel AJ, et al. (2011) Extended surfaces modulate hydrophobic interactions of neighboring solutes. Proc Natl Acad Sci USA 108(43):17678–17683. 4 Vaikuntanathan S, Rotskoff G, Hudson A, Geissler PL (2016) Necessity of capillary modes in a minimal model of nanoscale hydrophobic solvation. Proc Natl Acad Sci USA 113(16):E2224–E2230. 5 Ducker WA, Mastropietro D (2016) Forces between extended hydrophobic solids: Is there a long-range hydrophobic force? Curr Opin Colloid Interface Sci 22:51–58. 6 Cavallaro M, Jr, Botto L, Lewandowski EP, Wang M, Stebe KJ (2011) Curvature-driven capillary migration and assembly of rod-like particles. Proc Natl Acad Sci USA 108(52):20923–20928. 7 Matysiak S, Debenedetti PG, Rossky PJ (2012) Role of hydrophobic hydration in protein stability: A 3D water-explicit protein model exhibiting cold and heat denaturation. J Phys Chem B 116(28):8095–8104. 8 Lee CY, McCammon JA, Rossky PJ (1984) The structure of liquid water at an extended hydrophobic surface. J Chem Phys 80:4448–4455. 9 Stillinger FH (1973) Structure in aqueous solutions of nonpolar solutes from the standpoint of scaled-particle theory. J Solution Chem 2:141–158. 10 Ashbaugh HS, Pratt LR (2006) Colloquium: Scaled particle theory and the length scales of hydrophobicity. Rev Mod Phys 78:159. 11 Southall NT, Dill KA (2000) The mechanism of hydrophobic solvation depends on solute radius. J Phys Chem B 104:1326–1331. 12 Rajamani S, Truskett TM, Garde S (2005) Hydrophobic hydration from small to large lengthscales: Understanding and manipulating the crossover. Proc Natl Acad Sci USA 102(27):9475–9480. 13 Li IT, Walker GC (2011) Signature of hydrophobic hydration in a single polymer. Proc Natl Acad Sci USA 108(40):16527–16532. 14 Pratt LR, Chandler D (1977) Theory of the hydrophobic effect. J Chem Phys 67:3683–3704. 15 Chandler D (1993) Gaussian field model of fluids with an application to polymeric fluids. Phys Rev E Stat Phys Plasmas Fluids Relat Interdiscip Topics 48(4): 2898–2905. 16 Hummer G, Garde S, García AE, Pohorille A, Pratt LR (1996) An information theory model of hydrophobic interactions. Proc Natl Acad Sci USA 93(17):8951–8955. 17 ten Wolde PR, Sun SX, Chandler D (2002) Model of a fluid at small and large length scales and the hydrophobic effect. Phys Rev E Stat Nonlin Soft Matter Phys 65(1 Pt 1):011201. 18 Varilly P, Patel AJ, Chandler D (2011) An improved coarse-grained model of solvation and the hydrophobic effect. J Chem Phys 134(7):074109. 19 Weeks JD (1977) Structure and thermodynamics of the liquid–vapor interface. J Chem Phys 67:3106–3121. 20 Vaikuntanathan S, Geissler PL (2014) Putting water on a lattice: The importance of long wavelength density fluctuations in theories of hydrophobic and interfacial phenomena. Phys Rev Lett 112(2):020603. 21 Sedlmeier F, Netz RR (2012) The spontaneous curvature of the water-hydrophobe interface. J Chem Phys 137(13):135102. 22 Willard AP, Chandler D (2010) Instantaneous liquid interfaces. J Phys Chem B 114(5):1954–1958. 23 Godawat R, Jamadagni SN, Garde S (2009) Characterizing hydrophobicity of interfaces by using cavity formation, solute binding, and water correlations. Proc Natl Acad Sci USA 106(36):15119–15124. 24 Patel AJ, et al. (2012) Sitting at the edge: How biomolecules use hydrophobicity to tune their interactions and function. J Phys Chem B 116(8):2498–2503. 25 Setny P, Baron R, Michael Kekenes-Huskey P, McCammon JA, Dzubiella J (2013) Solvent fluctuations in hydrophobic cavity-ligand binding kinetics. Proc Natl Acad Sci USA 110(4):1197–1202. 26 Remsing RC, et al. (2015) Pathways to dewetting in hydrophobic confinement. Proc Natl Acad Sci USA 112(27):8181–8186. 27 Tiwary P, Mondal J, Morrone JA, Berne BJ (2015) Role of water and steric constraints in the kinetics of cavity-ligand unbinding. Proc Natl Acad Sci USA 112(39): 12015–12019. 28 Hummer G, Pratt LR, García AE (1997) Multistate gaussian model for electrostatic solvation free energies. J Am Chem Soc 119:8523–8527. 29 Remsing RC, Liu S, Weeks JD (2016) Long-ranged contributions to solvation free energies from theory and short-ranged models. Proc Natl Acad Sci USA 113:2819–2826. Xi and Patel PNAS | April 26, 2016 | vol. 113 | no. 17 | 4551 D o w n lo a d e d a t C a rn e g ie M e llo n U n iv e rs ity o n A p ri l 5 , 2 0 2 1