<<

View metadata, citation and similar papers at core.ac.uk brought to you by CORE

provided by Caltech Authors

pubs.acs.org/JPCL

Superacid Chemistry on Mildly Acidic Water † ‡ † † Shinichi Enami, Logan A. Stewart, Michael R. Hoffmann, and Agustín J. Colussi*, † W. M. Keck Laboratories, California Institute of Technology, Pasadena, California 91125, United States, and ‡ Department of Physics, University of California, Santa Barbara, California 93106, United States

ABSTRACT The mechanism of transfer across water-hydrophobic media boundaries is investigated in experiments in which the of gaseous n-hexanoic (PCOOH) upon collision with liquid water microjets is monitored by online electrospray mass spectrometry as a function of pH. Although PCOOH(aq) is þ ( K þ - ) ( ) ( ) a very weak p BH < 3 , PCOOH g is converted to PC OH 2 on pH < 4 water via a process that ostensibly retains some of the exoergicity of its gas-phase þ þ ( ) þ þ ΔG - -1 counterpart, PCOOH H3O =PC OH 2 H2O, < 22 kcal mol . The large ( ) þ ( ) þ kinetic isotope effects observed on H2O/D2O microjets, PC OH 2 /PC OH OD = ( ) þ ( ) þ 88 and PC OH OD /PC OD 2 = 156 at pD = 2, and their inverse dependences on pH indicate that PCOOH(g) hydronation on water (1) involves tunneling, (2) is faster than H-isotope exchange, and (3) is progressively confined to the outermost layers as water becomes more acidic. Proton transfers across steep water density gradients appear to be promoted by both dynamic and thermodynamic factors. SECTION Atmospheric, Environmental and Green Chemistry

( ) ( ) ( ) arious classes of chemical reactions have been found gas-phase basicities GB larger than H2O, GB B >GB H2O = -1 26 þ( ) to proceed faster and/or along different pathways at 158 kcal mol , and are readily protonated by H3O g ,as - water boundaries with hydrophobic phases.1 4 These evinced by the versatility of proton-transfer reaction mass V 27 þ ( ) phenomena have been tentatively ascribed to the peculiar spectrometry. The larger stabilities of H2O n 3 H3O versus 5-9 ( ) þ 28 structure oracidity of interfacial water. Herein, we advance H2O n 3 BH clusters, however, render the free energies of the notion that they are due both to the onset of favorable reactions that generate, annihilate, or interconvert charged thermochemistry and to the unique dynamics of proton species at water's surfaces (such as reaction R3) sensitive func- transfer under steep water density gradients. We point out that tions of n or, equivalently, of water density, Fw,particularlyabout F ≈ 29,30 these two conditions should generally arise during catalytic w 0. The empirical eq E1 encodes such functionality - events within enzyme active sites.10 19 ð Þþ þð Þ f þð Þþ ð ÞðÞ In the laboratory, Bronsted€ acidity scales are based on proton B s H3O s BH s H2O s R3 activity relative to a realizable standard state, or the posi- ΔG ðF Þ¼ΔG þðΔG - ΔG ÞF R < R < ð Þ tion of equilibria among conjugated acid/base pairs. Uniform 3 w 1 2 1 w 0 1 E1 proton activity/chemical potential across phases in thermo- Δ ( The exoergicity of gas-phase proton transfers, G1 = dynamic equilibrium ensures that local acidity defined, e.g., ( ) - ( ) as the pH of bulk water relative to a local standard state) GB H2O GB B < 0, is therefore expected to be exquisitely is a global acidity, even at water boundaries of widths com- tunable and eventually reversed with great economy of mensurate with the range of intermolecular forces.20,21 The motion within interfacial layers. Our report underscores the importance of this phenomenon. acidity of interfacial water is therefore identical to the acidity ( ) of bulk water (in the same standard state). Alternatively, A is In a recent study, trimethylamine, Me3N g , a strong base ( þ ) ( ( € ) both in bulk water pKBH =9.8 and in the gas phase GB = said to be a stronger Bronsted acid than B in any given -1) þ( ) medium if would bind preferentially to B. Because 227 kcal mol , could be captured as Me3NH s on water þ only at pH < 4 (rather than at pH < pKBHþ) in a process hydronium, H3O , can protonate most organic compounds þ( ) in the gas phase (reaction R1)22,23 but only relatively strong clearly controlled by the unavailability of interfacial H3O s bases in bulk water (reaction R2)24,25 in the pH>4 range rather than by unfavorable thermo- chemistry.5 It must be emphasized that such finding strictly þ þ þ BðgÞþH3O ðgÞ f BH ðgÞþH2OðgÞðR1Þ implies that H3O actually emerges to the surface of pH < 4 ( ) water, where it can protonate impinging Me3N g molecules þ þ ( BðaqÞþH3O ðaqÞ f BH ðaqÞþH2OðlÞðR2Þ during collision times not that the surface of 4 < pH < 7 it is apparent that the extent of protonation of Bronsted€ and bases on water's surface is determined not only by Received Date: October 12, 2010 (global) acidity but, critically, by their hydration status Accepted Date: December 1, 2010 therein. It is a remarkable fact that most nonalkanes have Published on Web Date: December 03, 2010

r 2010 American Chemical Society 3488 DOI: 10.1021/jz101402y |J. Phys. Chem. Lett. 2010, 1, 3488–3493 pubs.acs.org/JPCL

Scheme 1. Creation of a Microjeta in the Spraying Chamber of an Electrospray Mass Spectrometer by Injecting Acidified Water through an Electrically Grounded Nebulizer

a The microjet is briefly exposed to hexanoic acid vapors before it is broken up (at ∼10 μs) into charged microdroplets by the fast nebulizer gas. Upon subsequent solvent evaporation, in excess are ultimately ejected from the microdroplets via field desorption and detected by mass spectrometry within 1 ms. water is necessarily “basic”). On the basis of the preceding proportional to concentrations (up to a few mM) in the þ( ) 5,34 considerations, herein, we explore the possibility that H3O s interfacial layers of the microjets prior to their breakup. ( ) could protonate bases weaker than Me3N g , that is, whether See Experimental Methods and Figure S1 in the Supporting þ( ) þ( ) H3O s is a stronger proton donor than H3O aq of the Information, SI, for further details. The present study focuses ( ) þ ( ) same nominal acidity. In other words, we inquire whether on the evolution of PC OH 2 m/z =117 ESI signals as the thermochemistry of reaction R3 approaches that of functions of pH and solvent deuteration. ( ) þ reaction R1 or R2 on water's surface. Figure 1 shows the PC OH 2 ESI mass spectral m/z =117 ( ) ( ) We chose hexanoic acid, PCOOH, GB PCOOH = signal intensities, I11 7 , detected on 1 aqueous microjets -1 26 31 187 kcal mol >GB(H2O), pKBHþ < -3, as a suitable exposed to 77 ppbv PCOOH(g) and (2) microjets containing probe in experiments where reactive events on the surface 1mMPCOOH(aq), as functions of the pH of the bulk liquid. If ( ) ( ) ( ) of aqueous microjets exposed to PCOOH g /N2 g mixtures all PCOOH g collisions with the surface of microjets yielded þ for ∼10 μs are monitored by online electrospray ionization PC(OH)2 , that is, its uptake coefficient were γ ≈ 1, we esti- ( ) μ -1 11 ( ) þ( ) -2 mass spectrometry ESI-MS . A flow of 50 L min of water mate that <1.5 10 PC OH 2 s cm would be gener- (whose pH had been adjusted using concentrated HCl) was ated upon exposure to 77 ppbv PCOOH(g) for 10 μs.5 As - injected as a liquid microjet via an electrically grounded a reference, a 1 mM interfacial layer that is δ=1 10 7 cm - pneumatic nozzle (bore diameter: 100 μm) into the spraying deep contains 6 1010 PCOOH(aq) cm 2. It is apparent chamber of an ESI mass spectrometer held at 1 atm and that the reactive uptake of PCOOH(g) via protonation ap- ( ) ( ) ( ) 293 K Scheme 1 . The fast nebulizer gas N2 soon tore up proaches 100% efficiency, whereas predissolved PCOOH aq the outer layers of the microjet into microdroplets carrying (pKBHþ < -3) is minimally protonated at pH > 1. We infer 32 ( ) ( ) þ( ) ion excesses of either sign. Excess ions were eventually that PCOOH g is directly captured as PC OH 2 s on the þ( ) field-ejected to the gas phase from evaporating micro- surface of pH < 4 water via proton transfer from H3O s droplets, mass-analyzed, and detected within 1 ms. The before it becomes hydrated. Thermochemical considerations ( ) þ PC OH 2 ESI-MS signals detected in these experiments suggest that most nonalkane gases could be similarly proto- therefore correspond to species produced on the outermost nated on the surface of pH < 4 water, albeit, perhaps, at layers of the microjet upon brief exposure to PCOOH(g).5,33 different rates. Competitive substrate protonation versus We had previously verified that this setup operates as a hydration at water-hydrophobic media interfaces is the key þ( ) 35 linear transfer device, that is, ESI-MS signals are directly condition for H3O s to manifest itself as a superacid.

r 2010 American Chemical Society 3489 DOI: 10.1021/jz101402y |J. Phys. Chem. Lett. 2010, 1, 3488–3493 pubs.acs.org/JPCL

( ) Figure 1. Protonation of hexanoic acid PCOOH on water. Figure 2. ESI mass spectrum of hydronated hexanoic acid, m/z = þ m z þ þ þ PCOOH2 ESI-MS / =117 signalþ intensities as functions of pH. 117 [PC(OH) ],118[PC(OH)(OD) ], and 119 [PC(OD) ],onH O/ Blue circles: Signals from PCOOH produced on aqueous micro- ( ) 2 2 (2 ) (2 ) μ D2O 5/95 microjets at pD = 3.2 exposed to 77 ppbv PCOOH g . jets exposed to 77 ppbv PCOOH g for 10 s. The blueþ line is a See text for details. titration curve with pH1/2 = 2.5. Tan triangles: PCOOH2 signals from 1 mM PCOOH(aq) microjets. All experiments are in 1 atm of N2(g) at 293 K.

Scheme 2. Protonated Hexanoic Acid Tautomers (A and B) and the Product (C) of Their Collisionally Induced Dissociations (CIDs)

In this regard, we would like to point out that the probability of ( ) CH3COOD H-isotope exchange on neutral H2O l is larger than that of its uptake,36 suggesting that the hydration of polar Figure 3. Kinetic isotope effects (KIEs) for the hydronation of acetic acid is already slower than interfacial H-exchange. ( ) ( ) gaseous hexanoic acid PCOOH on H2O/D2O 5/95 microjets In our experiments, PCOOH(g) protonation via reaction R3 exposed to 77 ppbv PCOOH(g) as functions of pD. Tan downward 37 ( ) I I ( ) leads to oxonium (A) or carboxonium (B) species, which triangles and curve: KIE = 95/5 117/ 118 corrected .Blueupward 0 ( ) I ( ) I ( ) fragment to the oxonium C (m/z =89) via neutral CO losses triangles and curve: KIE = 95/5 118 corrected / 119 corrected . þ See text for details. (Scheme 2)38 rather than into the acylium PCO (m/z = 99, 39 0 0 not shown) plus H2O, as reported elsewhere. Noticeably, KIE = F /0.053 values calculated from the ratio F0 ( )( )þ ( ) þ ( ) ( ) Figure 2 shows the mass spectrum obtained by exposing =PCOH OD /PC OD 2 = I11 8 corrected /I11 9 corrected ( ) [ ( ) þ H2O/D2O 5:95 microjets at pD = 3.2 pD = pH measured become increasingly larger than KIE at lower pD, revealing 0.4] to PCOOH(g). It is apparent that the ratio of the that proton transfer becomes progressively faster than F ( ) þ ( )( )þ ≈ isotopologues' raw signals, =PCOH 2 /PC OH OD (neutral) H-atom exchange in more acidic water. 2.78, is much larger than that expected from the ratio of Our observations are consistent with the direct protonation H/D abundances, H/D = 5/95 = 0.053, or from a process of PCOOH(g) on the surface of water by minimally hydrated 40 þ displaying a normal kinetic H-isotope effect KIE < 12, H3O (s)(Figure 4). The thermochemistry of reaction R1 F <12 0.053 = 0.64. Notice that I11 9 :I11 8 :I11 7 /1:0.105: ultimately merges into that of reaction R2 via to the cumula- ( ) þ 0.003 signal intensities should be expected for fully ex- tive hydration energy differences between H2O n 3 H3O and ( ) ( ) þ 28 changed hydronated PCOOH. KIE values, KIE = 0.053 H2O m 3 PC OH 2 cluster ions. The exothermicity/exoergi- F( ) (Δ ≈ Δ ≈ ) corrected , calculated from ratios of corrected experimen- city S1 S2 0 of reaction R1 gradually decreases F( ) ( ) (F ) tal signal intensities, corrected = I11 7 /I11 8 corrected ,as as conditions shift from gas-phase w =0 to bulk water ( ) (F -3) functions of pD are shown in Figure 3. I11 8 corrected takes w=1 g cm . The final outcome is that proton exchange 13 into account the C contribution of I11 7 to I11 8 signals, between fully hydrated species becomes endothermic/ ( ) ( ) - ( ) I11 8 corrected = I11 8 measured 0.065 I11 7 measured . endoergic. In essence, the protonation of PCOOH(aq) requires

r 2010 American Chemical Society 3490 DOI: 10.1021/jz101402y |J. Phys. Chem. Lett. 2010, 1, 3488–3493 pubs.acs.org/JPCL

to/from interfacial acceptors/donors. Because acidity is essentially a relative concept, it is not immediately obvious how the protonating ability of water's surface could be deduced from its structure. In plain terms, water's surfaces are not more or less acidic than bulk water, and the degree of protonation/deprotonation of specific species on water's surfaces is not univocally determined by their molecular configurations. Enzyme function arises not only from the structure of their active sites but from the nature and role of near-surface water.53 Attempts at deducing the acidity of water's surface from its structure are therefore akin to expecting to hear the sound of one hand clapping. The KIE associated with the hydronation of PCOOH(g) at the air/water interface is about an order of magnitude larger than that expected from semiclassical transition-state theory - for gas-phase reactions, KIE < 12.54 57 Rigorous analysis of kinetic isotope effects in homogeneous liquid phases involves, Figure 4. Enthalpy diagram for the protonation of hexanoic acid þ however, extensive conformational averaging of solvent PCOOH by H3O under increasing solvation. Data adapted from 40 28 modes. A self-consistent interpretation of the KIE results Meot-Ner, by assuming (1) similar solvation enthalpies for pro- tonated hexanoic acid-water and protonated acetic acid-water of Figure 3 of proton transfer at the air-water interface should ( ) ΔH ( þ( )) - -1 41,46 clusters and 2 hydration H3O g = 115 kcal mol . deal, in addition, with decreasing solvent participation along the reaction coordinate. Such a task is beyond the scope of this strongly acidic conditions because the hydration energy of report,58 but we notice that proton transfers in enzymatic Δ ( þ) - -1 41 59 Hhyd H3O = 115 kcal mol , significantly reactions often display “abnormally” large KIEs and hint ( ) þ 24,42 exceeds those of larger ions, such as PC OH 2 . A that this circumstance may be associated with the steep water þ( ) similar situation will develop whenever H3O s encounters density gradients in which they take place. The increasingly 43 proton acceptors on the surface of pH < 4 water. The vast larger and diverging KIE and KIE0 values observed at lower pH 26 database on gas-phase proton affinities anticipates that further suggest that these behaviors could result from proton þ( ) H3O s could protonate both strong bases, such as trimethyl- transfers across progressively thinner reaction barriers as 5 þ ( þ ) ( ) amine pKBH =9.8 , and very weak ones, such as PCOOH aq H3O (s) emerges to the outermost water layers. Within the ( þ - ) pKBH < 3 , alkenes and carbonyls, that is, the protonating context of present findings, previous statements in the litera- þ( ) þ( ) potential of H3O s greatly exceeds that of H3O aq of the ture reflecting that “most reactions that take place in solution same acidity. Because the protonation of carboxylic acids in do not occur in the gas phase and vice versa”and that “clearly, 31,37 þ condensed media requires superacidic conditions, H3O - a quantitative knowledge of solvation forces might lead to the ( ) s formally behaves as a superacid in present experiments. understanding and manipulation of enormous new areas of Thus, phenomenological evidence, thermodynamics, and chemistry”43 and that “it is a common misconception to the fact that the isoelectric point of water falls in the 2 < consider enzymatic reactions as actually taking place `in 44 - pH < 4 range at all inert hydrophobic interfaces suggest water'”3,12,14,15,59 61 seem apropos and more farsighted than that the protonation of neutral species approaching water ever. from hydrophobic media should be a general phenomenon. Summing up, we report that gaseous hexanoic acid mole- At this point, it should be apparent that the statements cules are readily protonated upon collision with the surface of “ ” hydronium emerges to the surface of pH < 4 water and mildly acidic pH < 4 water in a process that displays exceed- “ ” interfacial hydronium behaves as a superacid are not ingly large kinetic H-isotope effects. We ascribe these phe- contradictory because they refer to different phenomena. nomena to the action of minimally solvated interfacial The negative charge of droplets and bubbles above the hydronium as “superacidic” proton donors and suggest that isoelectric point strictly implies that hydroxide approaches they may be common to chemistry at water interfaces with ( the air/water interface closer than hydronium by reasons most hydrophobic media. that have been analyzed elsewhere)44,45 but provides no evidence about whether hydronium or hydroxide are available SUPPORTING INFORMATION AVAILABLE Additional data to incoming proton acceptors/donors. Whether the surface of and experimental details. This material is available free of charge water is basic in the 4 < pH < 7 range, that is, whether it can via the Internet at http://pubs.acs.org. accept protons from strong and weak acids, remains to be 5 demonstrated experimentally. AUTHOR INFORMATION The water's surface is not a “surface” but a highly inhomo- geneous, discrete layer where water density vanishes within a Corresponding Author: nanometer. Therefore, the presence of hydronium/hydroxide *To whom correspondence should be addressed. ions somewhere in the interfacial layer, as inferred from 47-52 calculations or surface-specific spectroscopies, does ACKNOWLEDGMENT S.E. appreciates the Japan Society for the Pro- not imply that these species can actually proton transfer motion of Sciences Postdoctoral Fellowships for Research Abroad.

r 2010 American Chemical Society 3491 DOI: 10.1021/jz101402y |J. Phys. Chem. Lett. 2010, 1, 3488–3493 pubs.acs.org/JPCL

This work was supported by the U.S. National Science Foundation, (22) Brodbelt, J. S. Analytical Applications of Ion-Molecule Reac- Grant AGS-0964853. tions. Mass Spectrom. Rev. 1997, 16,91–110 . (23) Uggerud, E. Physical Organic Chemistry of the Gas Phase. Reactivity Trends for Organic Cations. Mod. Mass Spectrom. REFERENCES 2003, 225,3–36. (24) Anslyn, E. V.; Dogherty, D. A. Modern Physical Organic (1) Narayan, S.; Muldoon, J.; Finn, M. G.; Fokin, V. V.; Kolb, H. C.; Chemistry; University Science Books: Sausalito, CA, 2006. Sharpless, K. B. “On Water”: Unique Reactivity of Organic (25) Arnett, E. M. Quantitative Comparisons of Weak Organic Compounds in Aqueous Suspension. Angew. Chem., Int. Ed. Bases. Prog. Phys. Org. Chem. 1963, 1, 223–403. 2005, 44, 3275–3279. (26) Hunter, E. P. L.; Lias, S. G. Evaluated Gas Phase Basicities and (2) Acevedo, O.; Jorgensen, W. L. Advances in Quantum and Proton Affinities of Molecules: An Update. J. Phys. Chem. Ref. Molecular Mechanical (QM/MM) Simulations for Organic and Data 1998, 27,413–656. Enzymatic Reactions. Acc. Chem. Res. 2010, 43, 142–151. (27) Blake, R. S.; Monks, P.S.; Ellis, A. M. Proton-Transfer Reaction (3) Shapiro, N.; Vigalok, A. Highly Efficient Organic Reactions Mass Spectrometry. Chem. Rev. 2009, 109,861–896. “On Water”, “In Water”, and Both. Angew. Chem., Int. Ed. (28) Meot-Ner, M. The Ionic Hydrogen Bond. Chem. Rev. 2005, 2008, 47, 2849–2852. 105,213–284. (4) Poynor, A.; Hong, L.; Robinson, I. K.; Granick, S.; Zhang, Z.; (29) Shamay, E. S.; Buch, V.; Parrinello, M.; Richmond, G. L. At the Fenter, P. A. How Water Meets a Hydrophobic Surface, Phys. Water's Edge: Nitric Acid as a Weak Acid. J. Am. Chem. Soc. Rev. Lett. 2006, 97, doi 10.1103/PhysRevLett.97.266101. 2007, 129, 12910–12911. (5) Enami, S.; Hoffmann, M. R.; Colussi, A. J. Proton Availability – (30) Jungwirth, P.; Tobias, D. J. Ions at the Air/Water Interface. at the Air/Water Interface. J. Phys. Chem. Lett. 2010, 1, 1599 – 1604. J. Phys. Chem. B 2002, 106, 6361 6373. - (6) Jung, Y. S.; Marcus, R. A. Protruding Interfacial OH Groups (31) Zalewski, R. I. Acid Base Properties of Carboxylic Acids, and 'On-Water' Heterogeneous Catalysis, J. Phys., Condens. Esters and Amides. In The Chemistry of Acid Derivatives; Patai, , pp 305-369. Matter 2010, 22, doi 28411710.1088/0953-8984/22/28/ S., Ed.; Wiley: Chichester, U.K., 1992; Vol. 2 284117. (32) Zilch, L. W.; Maze, J. T.;Smith, J. W.; Ewing, G. E.; Jarrold, M. F. (7) Jung, Y.S.; Marcus, R. A. on the Theory of Organic Catalysis on Charge Separation in the Aerodynamic Breakup of Micro- Water. J. Am. Chem. Soc. 2007, 129, 5492–5502. meter-Sized Water Droplets. J. Phys. Chem. A 2008, 112, – (8) Marcus, R. A. Spiers Memorial Lecture Interplay of Theory 13352 13363. and Computation in Chemistry: Examples from On-Water (33) Enami, S.; Hoffmann, M. R.; Colussi, A. J. Prompt Formation Organic Catalysis, Enzyme Catalysis, and Single-Molecule of Organic Acids in Pulse Ozonation of Terpenes on Aqueous – Fluctuations. Faraday Discuss. 2010, 145,9–14. Surfaces. J. Phys. Chem. Lett. 2010, 1, 2374 2379. (9) Acevedo, O.; Armacost, K. Claisen Rearrangements: Insight (34) Cheng, J.; Psillakis, E.; Hoffmann, M. R.; Colussi, A. J. Acid into Solvent Effects and “On Water” Reactivity from QM/MM Dissociation versus Molecular Association of Perfluoroalkyl Simulations. J. Am. Chem. Soc. 2010, 132, 1966–1975. Oxoacids: Environmental Implications. J. Phys. Chem. A (10) Enami, S.; Hoffmann, M. R.; Colussi, A. J. Absorption of 2009, 113, 8152–8156. – - Inhaled NO2. J. Phys. Chem. B 2009, 113, 7977 7981. (35) Nathanson, G. M. Molecular Beam Studies of Gas Liquid (11) Li, C. J.; Trost, B. M. Green Chemistry for Chemical Synthesis. Interfaces. Annu. Rev. Phys. Chem. 2004, 55,231–255. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 13197–13202. (36) Shi, Q.; Li, Y. Q.; Davidovits, P.; Jayne, J. T.; Worsnop, D. R.; (12) Mase, N.; Barbas, C. F. In Water, On Water, and By Water: Mozurkewich, M.; Kolb, C. E. Isotope Exchange for Gas-Phase Mimicking Nature's Aldolases with Organocatalysis and Acetic Acid and Ethanol at Aqueous Interfaces: A Study of Water. Org. Biomol. Chem. 2010, 8, 4043–4050. Surface Reactions. J. Phys. Chem. B 1999, 103,2417–2430. (13) Chanda, A.; Fokin, V. V. Organic Synthesis “On Water. Chem. (37) Olah, G. A. Superacid Chemistry, 2nd ed.; Wiley: Hoboken, NJ, Rev. 2009, 109, 725–748. 2009. (14) Kiss, G.; Rothlisberger, D.; Baker, D.; Houk, K. N. Evaluation (38) Dookeran, N. N.; Yalcin, T.; Harrison, A. G. Fragmentation and Ranking of Enzyme Designs. Protein Sci. 2010, 19, 1760– Reactions of Protonated Alpha-Amino Acids. J. Mass Spectrom. 1773. 1996, 31, 500–508. (15) Mehler, E. L.; Fuxreiter, M.; Simon, I.; Garcia-Moreno, E. B. (39) Middlemiss, N. E.; Harrison, A. G. Structure and Fragmenta- The Role of Hydrophobic Microenvironments in Modulating tion of Protonated Carboxylic Acids in the Gas-Phase. Can. J. – – pKa Shifts in Proteins. Proteins 2002, 48, 283 292. Chem. 1979, 57, 2827 2833. (16) McEvoy, J. P.; Brudvig, G. W. Structure-Based Mechanism of (40) Truhlar, D. G. Tunneling in Enzymatic and Nonenzymatic Photosynthetic Water Oxidation. Phys. Chem. Chem. Phys. Hydrogen Transfer Reactions. J. Phys. Org. Chem. 2010, 23, 2004, 6, 4754–4763. 660–676. (17) Beattie, J. K.; McErlean, C. S. P.; Phippen, C. B. W. The (41) Gal, J. F.; Maria, P. C.; Raczynska, E. D. Thermochemical Mechanism of On-Water Catalysis. Chem.;Eur. J. 2010, 16, Aspects of Proton Transfer in the Gas Phase. J. Mass Spectrom. 8972–8974. 2001, 36, 699–716. (18) Ball, P. Water as an Active Constituent in Cell Biology. Chem. (42) Bohm, S.; Exner, O. Basicity of Carboxylic Acids: Resonance Rev. 2008, 108,74–108. in the Cation and Substituent Effects. New J. Chem. 2005, 29, (19) Friedman, R. Proton Transfer on the Molecular Surface 336–342. of Proteins and Model Systems. Isr. J. Chem. 2009, 49, (43) Arnett, E. M. Gas-Phase Proton Transfer ; A Breakthrough 149–153. for Solution Chemistry. Acc. Chem. Res. 1973, 6, 404–409. (20) Rowlinson, J. S. Thermodynamics of Inhomogeneous Systems. (44) Beattie, J. K.; Djerdjev, A. N.; Warr, G. G. The Surface of Neat Pure Appl. Chem. 1993, 65,873–882. Water is Basic. Faraday Discuss. 2009, 141,31–39. (21) Widom, B. Structure of Interfaces from Uniformity of the (45) Cheng, J.; Hoffmann, M. R.; Colussi, A. J. Anion Fractionation Chemical Potential. J. Stat. Phys. 1978, 19, 563–574. and Reactivity at Air/Water:Methanol Interfaces. Implications

r 2010 American Chemical Society 3492 DOI: 10.1021/jz101402y |J. Phys. Chem. Lett. 2010, 1, 3488–3493 pubs.acs.org/JPCL

for the Origin of Hofmeister Effects. J. Phys. Chem. B 2008, 112, 7157–7161. (46) Mejias, J. A.; Lago, S. Calculation of the Absolute Hydration þ - Enthalpy and Free Energy of H and OH . J. Chem. Phys. 2000, 113, 7306–7316. (47) Buch, V.; Milet, A.; Vacha, R.; Jungwirth, P.; Devlin, J. P. Water Surface Is Acidic. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 7342– 7347. (48) Tian, C. S.; Shen, Y.R. Structure and Charging of Hydrophobic Material/Water Interfaces Studied by Phase-Sensitive Sum- Frequency Vibrational Spectroscopy. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 15148–15153. (49) Tarbuck, T. L.; Ota, S. T.; Richmond, G. L. Spectroscopic Studies of Solvated Hydrogen and Hydroxide Ions at Aqueous Surfaces. J. Am. Chem. Soc. 2006, 128, 14519–14527. (50) Petersen, P.B.; Saykally, R. J. Is the Liquid Water Surface Basic or Acidic? Macroscopic vs. Molecular-Scale Investigations. Chem. Phys. Lett. 2008, 458, 255–261. (51) Iuchi, S.; Chen, H. N.; Paesani, F.; Voth, G. A. Hydrated Excess Proton at Water-Hydrophobic Interfaces. J. Phys. Chem. B 2009, 113,4017–4030. (52) Kudin, K. N.; Car, R. Why are Water-Hydrophobic Interfaces Charged?. J. Am. Chem. Soc. 2008, 130,3915–3919. (53) Whitesides, G. M.; Lipomi, D. J. Soft Nanotechnology: “Struc- ture” vs. “Function. Faraday Discuss. 2009, 143, 373–384. (54) Zhao, X. G.; Tucker, S. C.; Truhlar, D. G. Solvent and Secondary Kinetic Isotope Effects for the Microhydrated SN2 Reaction Of -( ) – Cl H2O n with CH3Cl. J. Am. Chem. Soc. 1991, 113,826 832. (55) Kiefer, P. M.; Hynes, J. T. Theoretical Aspects of Tunneling Proton Transfer Reactions in a Polar Environment. J. Phys. Org. Chem. 2010, 23, 632–646. (56) Masgrau, L.; Ranaghan, K. E.; Scrutton, N. S.; Mulholland, A. J.; Sutcliffe, M. J. Tunneling and Classical Paths for Proton Transfer in an Enzyme Reaction Dominated by Tunneling: Oxidation of Tryptamine by Aromatic Amine Dehydrogenase. J. Phys. Chem. B 2007, 111,3032–3047. (57) Duong, M.C P.T.; Kim, Y.Theoretical Studies for the Rates and Kinetic Isotope Effects of the Excited-State Double Proton Transfer in the 1:1 7-Azaindole:H2O Complex Using Varia- tional Transition State Theory Including Multidimensional Tunneling. J. Phys. Chem. A 2010, 114, 3403–3410. (58) Bothma, J. P.; Gilmore, J. B.; McKenzie, R. H. The Role of Quantum Effects in Proton Transfer Reactions in Enzymes: Quantum Tunneling in a Noisy Environment?. New J. Phys. 2010, 12, 055002. (59) Basran, J.; Sutcliffe, M. J.; Scrutton, N. S. Enzymatic H-Transfer Requires Vibration-Driven Extreme Tunneling. Biochemistry 1999, 38,3218–3222. (60) Kemp, D. S. How to Promote Proton-Transfer. Nature 1995, 373, 196–197. (61) Pal, S. K.; Peon, J.; Zewail, A. H. Biological Water at the Protein Surface: Dynamical Solvation Probed Directly with Femto- second Resolution. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 1763–1768.

r 2010 American Chemical Society 3493 DOI: 10.1021/jz101402y |J. Phys. Chem. Lett. 2010, 1, 3488–3493