The Effect of Salt on the Water Structure at a Charged Solid Surface

Department of Chemistry, University of Victoria, Victoria, British Columbia, V8W 3 V6, Canada. J. Phys. Chem. Lett. , 0, (),. DOI: 10.1021/jz200251h@p...
0 downloads 0 Views 1MB Size
LETTER pubs.acs.org/JPCL

The Effect of Salt on the Water Structure at a Charged Solid Surface: Differentiating Second- and Third-order Nonlinear Contributions Kailash C. Jena, Paul A. Covert, and Dennis K. Hore* Department of Chemistry, University of Victoria, Victoria, British Columbia, V8W 3V6, Canada ABSTRACT: We have used visible-infrared sum-frequency generation spectroscopy to reveal fundamental characteristics of water structure at the fused silica surface. By studying a wide range of ionic strengths, from 0.05 mM to 4 M, we are able to comment on the contributions of second- and third-order nonlinearities to the spectroscopic response. Spectra obtained from extremely dilute salt concentrations provide evidence of the previously sought increasing surface charge with ionic strength. This is followed by a screening regime where the extent to which the surface field penetrates into the bulk is limited by the electrolyte. Data from intermediate salt concentrations reveal a few strongly ordered layers of water immediately adjacent to the surface. At high concentrations, we observe a significant disruption of solvent ordering. Together, the observation of these four distinct regimes provides a unified understanding of interfacial water structure in the presence of salt that consolidates previous reports in the literature. SECTION: Surfaces, Interfaces, Catalysis

I

nterfacial water structure plays an important role in many natural and industrial processes such as adsorption of proteins, ionic transport across membranes, soil formation, separation and purification techniques, and catalysis.16 In these application areas, the role of ions is crucial in screening electrostatic fields at charged solid interfaces and creating electrical double layers at airwater interfaces.712,14 As a result, many studies seek to address questions such as the following: Up to what depth are water molecules ordered? Do electrolytes contribute to the development or to the screening of the surface field? Does contact adsorption of ions significantly disturb the interfacial solvent structure? The answers to these questions have a profound impact on the subsequent adsorption, orientation, and conformation of molecules at charged interfaces. Recently, there have been many experimental and computational approaches addressing the issue of water structure at solid surfaces.711,1322 Among these studies, there has been some controversy surrounding the distance over which water is ordered at a charged surface. For example, X-ray scattering1922 and molecular dynamics simulations2326 indicate that structured water exists no further than ∼1 nm from the surface. On the other hand, data from nonlinear optical11,18,27 and atomic force measurements28 have suggested that water molecules may be structured up to the Debye length. Nonlinear optical spectroscopies such as electronic second harmonic generation (SHG)2937 and electronic/vibrational sum-frequency generation (SFG)3846 are particularly attractive for such investigations since, in the simplest case, the signals are expected to be dipole-forbidden for any molecules that are not structured in a polar manner. They therefore offer extreme sensitivity to interfaces, without relying on shallow bulk penetration of the beams. However, when these techniques are applied to charged interfaces, care must be taken in r 2011 American Chemical Society

the interpretation of the measured signals. At a charged surface, the presence of a strong electrostatic field at the interface acts as a third input field with zero frequency. In this letter we will rationalize how second- and third-order contributions to the electric susceptibility manifest themselves in the SFG signal as a function of ionic strength. By comparing data obtained over a wide range of salt concentrations, we are able to provide new insight into the interfacial water structure and reconcile previous results from the literature. In SHG and SFG experiments at neutral interfaces, signal is understood to result from the second-order susceptibly χ(2). However, it has been observed that the SHG signal can be enhanced by applying a DC electric field, and this has been attributed to the contribution of the third-order susceptibility χ(3).47,48 The above process is known as electric-field-induced second harmonic (EFISH) generation. In noncollinear geometries, the χ(2) and χ(3) signals have different phase-matching directions and so are easily distinguished.43,49 However, since the χ(3) contribution in EFISH comes from a static field E0, χ(2) and χ(3) signals are simultaneously detected as S  jχð2Þ Evis EIR þ χð3Þ Evis EIR E0 j2

ð1Þ

Here the χ(2) signal originates from those molecules that are asymmetrically orientated at the interface. The χ(3) signal has contributions from isotropic bulk water molecules (χ(3) iso ) and from oriented molecules due to the static electric field.8,50 In Figure 1a we illustrate that, under the electric dipole approximation, χ(2)  N2ÆR(2)æ = 0 in centrosymmetric environments, Received: February 24, 2011 Accepted: April 12, 2011 Published: April 15, 2011 1056

dx.doi.org/10.1021/jz200251h | J. Phys. Chem. Lett. 2011, 2, 1056–1061

The Journal of Physical Chemistry Letters

LETTER

Figure 1. An illustration of the relationship between χ(2) and χ(3) in the case of an (a) isotropic environment, (b) polar ordered environment, and (c) ordered environment in the absence of polarity. Arrows indicate the direction of the water dipoles.

Figure 3. Integrated intensity of all spectra shown in Figure 2, normalized with respect to the one acquired before salt addition. The top axis is drawn according to eq 2.

Figure 2. Sum frequency response as a function of the infrared energy. The black spectrum corresponds to the pure fused silicawater interface at pH 6, before any salt addition. Spectra in series A (cyan) correspond to dilute (4.8  1054.7  104 M) NaCl solutions; series B (red) 9.2  1044.7  102 M; series C (green) 0.131.1 M; series D (blue) 1.74.1 M NaCl.

where R(2) is the second-order molecular polarizability. In (3) is the third-order contrast χ(3)  N3ÆR(3)æ = χ(3) iso 6¼ 0, where R molecular polarizability. Here N2 refers to those molecules with a net polar (noncentrosymmetric) orientation, and N3 refers to all molecules experiencing the surface-originating field E0, regardless of whether they are aligned. We emphasize that while the secondorder susceptibility requires a polar orientation of water molecules, the third-order susceptibility is merely enhanced from its isotropic value when the water molecules are oriented (Figure 1b). Figure 1c illustrates that, even when there is a high degree of orientational order, χ(2) = 0 in the absence of polarity, since this again represents a centrosymmetric environment. The field present at aqueousoxide interfaces and its interaction with water dipoles has been well-studied, and is understood to promote ordering of water molecules at the interface.79,12,14,27,46,51,52 In studies of the fused silicawater interface, Ong et al. performed a thorough investigation in which they varied the solution pH, ionic strength, and temperature of the solution.8 Among their conclusions, their observation of an enhanced SHG signal attributed to an intrinsic interfacial electric field E0 is of central interest to this work. Du et al. studied the quartzwater interface using vibrationally resonant SFG spectroscopy.7 At pH values corresponding to a charged quartz surface, the addition of salt was observed to lower the SFG intensity. At neutral pH there was no change in the water structure, even at NaCl concentrations as high as 0.5 M. Considering this surface-originating field and its interaction with

water molecules, Yeganeh et al. used SFG to measure the isoelectric point of the Al2O3water interface by varying the pH of the solution.46 Similarly, Gragson and Richmond studied the molecular alignment and hydrogen bonding at charged airwater and CCl4water interfaces as a function of surface charge density, ionic strength, and temperature.11 EftekhariBafrooei and Borguet compared the OH vibrational lifetime at a neutral and a charged silica surface.27 A shorter lifetime at the charged surface was attributed to a greater number of solvation shells available for energy dissipation in a deeper interfacial region. In all of the above studies, the presence of E0 at a charged surface results in a greater depth over which water molecules are aligned, an increased orientation of interfacial water, and a χ(3) contribution to the signal. However, the relative contribution of χ(2) and χ(3) to the observed spectra is still an open question. From our data over a wide range of ionic strengths, we have identified several regimes that reveal the depth over which molecules respond to electrolyte addition, the balance between charge development and screening, and the relative contribution of second- and third-order optical nonlinearities to the spectral response. Figure 2 shows SFG spectra of the fused silicawater interface at various ionic strengths. The black spectrum corresponds to the neat interface, before any salt addition. In Figure 3 we plot the integrated SFG intensity, normalized with respect to this spectrum. In our discussion of the physical and chemical processes responsible for the trends in the data shown in Figure 3, two models of interfacial charge distribution are considered. At low ionic strength, the GouyChapman diffuse charge model4,11,53 describes the distribution of ionic species in the vicinity of a charged surface. The Debye length, which results from this model, describes the extent to which the electrolyte screens the surface field and may be calculated as rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1000εR RT 0:3  pffiffi ½nm ð2Þ ld ¼ 2 2 8πNA ε0 I I where εR is the bulk dielectric constant, R is the gas constant, T is the absolute temperature, NA is Avogadro’s number, and ε0 is the permittivity of free space. Values of ld obtained from this relationship are indicated in the top axis of Figure 3. At high 1057

dx.doi.org/10.1021/jz200251h |J. Phys. Chem. Lett. 2011, 2, 1056–1061

The Journal of Physical Chemistry Letters

LETTER

and the Stern model is a more appropriate description of the interface. In this model, the interfacial structure acts like a capacitor8 j0 ¼

4πσ0 d ε

ð3Þ

where j0 is the potential at the surface, ε is the dielectric constant, and d is the distance between the negatively charged surface and the cations. Prior to interpreting our data, careful consideration must be given to the coherence length lc38,54 specific to our experimental geometry. This quantity is defined as lc ¼

1 kðωSFG Þ þ kðωvis Þ þ kðωIR Þ

ð4Þ

where the wavevector in the refracted medium (salt solution) is given by qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ωi ð5Þ kðωi Þ ¼ nðωi Þ2  sin2 θi c

Figure 4. Proposed model of the balance between electrolyte screening of the surface electric field and charge-induced molecular order at the interface. The dashed line in panel a shows a function fit to empirical surface charge measurements (points) at the fused silicawater interface.57 The solid line in panel a represents the surface potential, calculated from the surface charge data at I e 0.13 M and by the Stern model at I > 0.13 M. Relative effects of interfacial order are shown in panel b by the solid lines for ÆR(2)æ and dotted lines for ÆR(3)æ. Contributions of the χ(2) (solid line) and χ(3) terms to the measured signal are shown in panel c. The solid line in panel d shows the modeled signal alongside our measured data (points).

ionic strengths (greater than 0.13 M) the surface charge reaches a level where the GouyChapman model is no longer suitable,

with θi being the refracted angle, c the speed of light, and i representing each of the beams. It is understood that the depth to which molecules contribute to the SFG process is limited to lc, regardless of the distance over which χ(2) or χ(3) are nonzero. In our case, lc = 48 nm at ωIR = 3000 cm1. This means that only the lowest ionic strength of region A is not sensitive to the effect of screening as ld > lc. We will now consider the four regions AD identified in Figure 3. The development of the surface charge, surface potential,53,57 ÆR(2)æ, ÆR(3)æ, χ(2), and χ(3)j0 with increasing ionic strength, along with the model predicted SFG intensities, is shown in Figure 4 to aid in this discussion. Region A. At ionic strengths less than 0.7 mM (indicated by the cyan data in Figure 2), there is no change in SFG intensity with increasing ionic strength. This region is clearly identified as the initial flat region in Figure 3. Since the static field penetrates into the bulk, it is reasonable to expect that screening by charged species in solution would reduce the relative contribution of the χ(3) term in eq 1. On the other hand, it has been established that increasing solution ionic strength promotes the development of a more negative charge at the surface,5557 as plotted in Figure 4a (points). This should increase the degree to which the first few layers of water are oriented, and thereby enhance the χ(2) and χ(3) contributions to the signal. Simulations of water structure next to charged interfaces show that, as the surface charge increases, water molecules are increasingly aligned adjacent to the interface, but are not oriented past ∼1.5 nm from the surface.2326 This balance between enhancement of χ(2) and χ(3) due to increased structure near the interface, and reduction of χ(3) due to screening is illustrated in Figure 4 and results in the region A plateau. In the original SHG ionic strength study by Eisenthal et al., the authors remarked that they did not observe evidence of the expected increasing surface charge with salt concentration.8 Since we start with much lower ionic strength, we are able to see a balance between screening and the development of additional surface charge that results in our region A plateau. Region B. Starting from a concentration of 0.7 mM, we observe that the signal drops as the ionic strength increases. This is accounted for in our model by continued development of the surface charge coupled with only a slight increase to a constant interfacial ordering. The net effect on the SFG signal is that the 1058

dx.doi.org/10.1021/jz200251h |J. Phys. Chem. Lett. 2011, 2, 1056–1061

The Journal of Physical Chemistry Letters χ(2) term in eq 1 now remains constant, while the χ(3) term decreases, being dominated by an increased screening of E0 throughout the region.5557 These results are in agreement with those observed by Eisenthal et al.8 As we approach Debye lengths of ∼2 nm, we are now near the extent of the outermost ordered noncentrosymmetric water layers. We propose that the slower drop in signal toward the end of region B is a sign that we are entering a region near the surface where there is a more significant contribution from χ(2). Region C. A second plateau in the SFG signal is observed between approximately 0.1 and 1 M ionic strength. We propose that this plateau is the effect of two phenomena. First, χ(2) dominates the signal in this region as a result of the short penetration distance of the surface field into bulk, resulting in a small relative χ(3) contribution. Second, this region marks the transition from the GouyChapman model of surface to the Stern model where the surface potential now remains constant. As a result, both the second- and third-order terms remain constant over this region. This means that the (already small) contribution of χ(3) remains constant over the entire range of ionic strengths in region C, further supporting the plateau feature observed in this region. Region D. At ionic strengths greater than 1.1 M, the signal drops rapidly. As this behavior abruptly follows the region C plateau, we believe that the hydrogen bonding environment near the interface is disturbed at these high salt concentrations. This results in a less-ordered environment near the surface and hence both χ(2) and the ordered component of χ(3) decrease rapidly. A similar behavior has been observed in the case of the airwater interface at high salt concentrations.10 The perturbation may be partially due to water displacement upon ion contact adsorption, thereby disrupting the highly ordered water layers immediately adjacent to the surface. If all polar ordering were to be disrupted, we would be left with only the isotropic contribution of χ(3). In summary, we have varied the concentration of salt to measure key characteristics of water structure at a charged solid surface. At extremely low ionic strength, we observe a plateau in the spectroscopic response. This is attributed to the expected but previously unobserved balance between screening and the development of additional surface charge for fused silica. At ionic strengths greater than 0.7 mM, a rapid drop in signal occurs as the surface field is now effectively screened. Near the end of region B, the trend gives way to a slower decrease as χ(2) dominates, approaching an interfacial depth of about 2 nm. Eventually we reach the limit of the ionic cloud diffuse charge model, and switch to a capacitor model at high ionic strength. Here, both χ(2) and χ(3) contributions remain constant with increasing salt concentration. Finally, at very high ionic strength, we observe a disruption in the hydrogen bonding network on account of contact adsorption. Our findings suggest that a few layers of water are strongly ordered near the charged surface. These ordered layers contribute to the second-order response that relies on symmetry breaking. The third-order response may be generated from a much greater depth, limited by the penetration of the surface field due to screening by the electrolyte ions. This highlights the necessity of considering χ(3) contributions in nonlinear optical studies of charged interfaces. Additional experiments would be able to further support our claims. For example, varying the electrolyte concentration and species over a range of pH values would be valuable. For these proposed experiments, resolving the amplitude and phase of the SFG signal under a variety of beam polarizations would also assist

LETTER

Figure 5. Detailed view of the experimental geometry. A Teflon cell inside an aluminum block contains the salt water solutions. The solutionfused silica interface consists of a fused silica prism connected to the Teflon cell by a fluoropolymer O-ring. Beams enter the prism and then reflect from the solutionprism interface.

in the further refinement of a model for surface water structure. It is noted that, while SFG experiments provide insight on the distribution of local hydrogen bonding environments, nonresonant SHG experiments would also provide an efficient and robust manner for obtaining pertinent data for these models.

’ EXPERIMENTAL METHODS A 1064 nm Nd:YAG laser with 25 ps pulse width and 10 Hz repetition rate (Ekspla PL2241) is doubled for use as one of our pump beams (ωvis) and also to create a tunable infrared beam (ωIR = 27503750 cm1, 5 cm1 step size) in a BBO/AgGaS2based OPA/OPG/DFG parametric generator (Ekspla PG501). The visible and IR pump beams are spatially and temporally overlapped at the fused silicawater interface at angles of 63 for the IR beam (diameter 0.5 mm, energy 200 μJ/pulse at 3000 cm1) and 66 for the visible beam (diameter 1 mm, energy 110 μJ/pulse), averaging 50 shots per data point. As shown in Figure 5, an IR-grade fused silica prism (Del Mar Photonics, CA) is clamped to a Teflon water cell with a fluoropolymer O-ring (Marco Rubber, NH) creating a water-tight seal. The SFG signal generated at the prismsolution interface was collected by a photo multiplier tube after passing through a 532-nm notch filter and monochromator in order to remove the contribution of the reflected fundamental beams. The prism was cleaned before each experiment by dipping into a concentrated solution of sulfuric acid containing 0.1% nitric acid, followed by copious rinsing in 18 MΩ 3 cm deionized water (Nanopure, Barnstead Thermo) and was allowed to dry covered before use. The same procedure was followed for the cleaning of all glassware and the Teflon sample cell. This Nanopure water was also used for the experiments and to prepare the NaCl (ACP, Montreal Canada) solutions. Since our working pH (≈ 6.0 ( 0.1) is well above the surface pH at zero charge (∼3), the fused silica surface is negatively charged.7,27,51 Since the surface is only partially charged at pH 6, this allows the surface charge state equilibrium to shift toward increased production of SiO upon the addition of salt. All spectra for this study were collected with p-polarized infrared and s-polarized visible beams incident at the fused silicawater interface; the s-component of the SFG response was recorded as a function of the infrared energy. The shapes of the spectra were corrected for local field effects according to the polarization of the beams and the incident geometry. As we scan the energy of the infrared beam, an increase in the signal results from resonance of χ(2) and χ(3) with vibrational 1059

dx.doi.org/10.1021/jz200251h |J. Phys. Chem. Lett. 2011, 2, 1056–1061

The Journal of Physical Chemistry Letters transitions of the interfacial water OH stretching modes. As a result of the complexity resulting from the multitude of hydrogenbonding environments between 27503750 cm1, the small changes we wish to capture, and the relative lack of features observed at high salt concentrations, all of our comparisons are made by integrating the spectral intensity, and normalizing with respect to the integrated intensity prior to salt addition.

’ AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected]. Phone: 250-721-7168. Fax: 250-7217147.

’ ACKNOWLEDGMENT We wish to thank the Natural Science and Engineering Research Council of Canada (NSERC) for support of this science with a Discovery Grant. Equipment was purchased with assistance from the Canadian Foundation for Innovation (CFI) Leaders Opportunity Fund, and the British Columbia Knowledge Development Fund (BCKDF). We thank Prof. Eric Borguet for fruitful discussions. ’ REFERENCES (1) Richmond, G. L. Molecular Bonding and Interactions at Aqueous Surfaces as Probed by Vibrational Sum Frequency Spectroscopy. Chem. Rev. 2002, 102, 2693–2724. (2) Kim, J.; Somorjai, G. A. Molecular Packing of Lysozyme, Fibrinogen, and Bovine Serum Albumin on Hydrophilic and Hydrophobic Surfaces Studied by InfraredVisible Sum Frequency Generation and Fluorescence Microscopy. J. Am. Chem. Soc. 2003, 125, 3150–3158. (3) Boman, F. C.; Mosorrafiti, M. J.; Gibbs, J. M.; Stepp, B. R.; Salazar, A. M.; Nguyen, S. T.; Geiger, F. M. DNA Single Strands Tethered to Fused Quartz/Water Interfaces Studied by Second Harmonic Generation. J. Am. Chem. Soc. 2005, 127, 15368–15369. (4) Bockris, J. O.; Reddy, A. K. N. Modern Electrochemistry; Plenum Press: New York, 1970. (5) Shen, Y. R.; Ostroverkhov, V. Sum-Frequency Vibrational Spectroscopy on Water Interfaces: Polar Orientation of Water Molecules at Interfaces. Chem. Rev. 2006, 106, 1140–1154. (6) Hopkins, A. J.; McFearin, C. L.; Richmond, G. L. Investigations of the SolidAqueous Interfaces with Vibrational Sum Frequency Spectroscopy. Curr. Opin. Solid State Mater. Sci. 2005, 9, 19–27. (7) Du, Q.; Freysz, E.; Shen, Y. R. Vibrational Spectra of Water Molecules at Quartz/Water Interfaces. Phys. Rev. Lett. 1994, 72, 238–241. (8) Ong, S.; Zhao, X.; Eisenthal, K. B. Polarization of Water Molecules at a Charged Interface: Second Harmonic Studies of the Silica/Water Interface. Chem. Phys. Lett. 1992, 191, 327–335. (9) Jena, K. C.; Hore, D. K. Variation of Ionic Strength Reveals the Interfacial Water Structure at a Charged Mineral Surface. J. Phys. Chem. C 2009, 113, 15364–15372. (10) Schnitzer, C.; Baldelli, S.; Shultz, M. J. Sum Frequency Generation of Water on NaCl, NaNO3, KHSO4, HCl, HNO3, and H2SO4 Aqueous Solutions. J. Phys. Chem. B 2000, 104, 585–590. (11) Gragson, D. E.; Richmond, G. L. Potential Dependent Alignment and Hydrogen Bonding of Water Molecules at Charged Air/Water and CCl4/Water Interfaces. J. Am. Chem. Soc. 1998, 120, 366–375. (12) Liu, D.; Ma, G.; Levering, L. M.; Allen, H. C. Vibrational Spectroscopy of Aqueous Sodium Halide Solutions and AirLiquid Interfaces: Observation of Increased Interfacial Depth. J. Phys. Chem. B 2004, 108, 2252–2260. (13) Toney, M. F.; Howard, J. N.; Richer, J.; Borges, G. L.; Gordon, J. G.; Melroy, O. R.; Wiesler, D. G.; Yee, D.; Sorensen, L. B. Voltage-

LETTER

Dependent Distribution of Water Molecules at the Electrode-Electrolyte Interface. Nature 1994, 368, 444–446. (14) Yang, Z.; Li, Q.; Chou, K. C. Structures of Water Molecules at the Interfaces of Aqueous Salt Solutions and Silica: Cation Effects. J. Phys. Chem. C 2009, 113, 8201–8205. The authors have observed a decrease in the water SFG response upon adding 104 M NaCl solution. There is no obvious reason for the different behavior we observed; perhaps this is due to a different fused silica surface structure that may result from a different cleaning procedure. (15) Nihonyanagi, S.; Ye, S.; Uosaki, K. Sum Frequency Generation (SFG) Study on the Molecular Structures at the Interfaces between Quartz Modified with Amino-Terminated Self-Assembled Monolayer and Electrolyte Solutions of Various pH and Ionic Strength. Electrochim. Acta 2001, 46, 3057–3061. (16) Tian, C.; Ji, N.; Waychunas, G. A.; Shen, Y. R. Interfacial Structures of Acidic and Basic Aqueous Solutions. J. Am. Chem. Soc. 2008, 130, 13033–13039. (17) Mucha, M.; Frigato, T.; Levering, L. M.; Allen, H. C.; Tobias, D. J.; Dang, L. X.; Jungwirth, P. Unified Molecular Picture of the Surfaces of Aqueous Acid, Base, and Salt Solutions. J. Phys. Chem. B 2005, 109, 7617–7623. (18) Chen, X.; Yang, T.; Kataoka, S.; Cremer, P. S. Specific Ion Effects on Interfacial Water Structure near Macromolecules. J. Am. Chem. Soc. 2007, 129, 12272–12279. (19) Fenter, P.; Sturchio, N. C. MineralWater Interfacial Structures Revealed by Synchrotron X-ray Scattering. Prog. Surf. Sci. 2004, 77, 171–258. (20) Zhang, Z.; Fenter, P.; Cheng, L.; Sturchio, N. C.; Bedzyk, M. J.; Predota, M.; Bandura, A.; Kubicki, J. D.; Lvov, S. N.; Cummings, P. T.; Ion Adsorption at the RutileWater Interface: Linking Molecular and Macroscopic Properties. Langmuir 2004, 20, 4954–4969. (21) Park, C.; Fenter, P. A.; Nagy, K. L.; Sturchio, N. C. Hydration and Distribution of Ions at the MicaWater Interface. Phys. Rev. Lett. 2006, 97, 016101. (22) Park, C.; Fenter, P. A.; Sturchio, N. C.; Nagy, K. L. Thermodynamics, Interfacial Structure, and pH Hysteresis of Rbþ and Sr2þ Adsorption at the Muscovite (001)Solution Interface. Langmuir 2008, 24, 13993–14004. (23) Xu, D.; Leng, Y.; Chen, Y.; Li, D. Water Structures near Charged (100) and (111) Silicon Surfaces. Appl. Phys. Lett. 2009, 94, 201901. (24) Dimitrov, D. I.; Raev, N. D. Molecular Dynamics of the Electrical Double Layer at the 1 M KCl Solution/Hg Electrode Interface. J. Electroanal. Chem. 2000, 486, 1–8. (25) Lee, S. H.; Rossky, P. J. A Comparison of the Structure and Dynamics of Liquid Water at Hydrophobic and Hydrophilic Surfaces - A Molecular Dynamics Simulation Study. J. Chem. Phys. 1994, 100, 3334–3345. (26) Yeh, I.-C.; Berkowitz, M. L. Dielectric Constant of Water at High Electric Field: Molecular Dynamics Study. J. Chem. Phys. 1999, 110, 7935–7942. (27) Eftekhari-Bafrooei, A.; Borguet, E. Effect of Surface Charge on the Vibrational Dynamics of Interfacial Water. J. Am. Chem. Soc. 2009, 131, 12034–12035. (28) Teschke, O.; Ceotto, G.; de Souza, E. F. Interfacial water dielectric-permittivity-profile measurements using atomic force microscopy. Phys. Rev. E 2001, 64, 011605. (29) Shen, Y. R. Optical 2nd Harmonic-Generation at Interfaces. Annu. Rev. Phys. Chem. 1989, 40, 327–350. (30) Eisenthal, K. B. Liquid Interfaces Probed by Second-Harmonic and Sum-Frequency Spectroscopy. Chem. Rev. 1996, 96, 1343–1360. (31) Vogel, V.; Shen, Y. R. Air/Liquid Interfaces and Adsorbed Molecular Monolayers Studied with Nonlinear Optical Techniques. Annu. Rev. Mater. Sci. 1991, 21, 515–534. (32) Geiger, F. M. Second Harmonic Generation, Sum Frequency Generation, and χ(3): Dissecting Environmental Interfaces with a Nonlinear Optical Swiss Army Knife. Annu. Rev. Phys. Chem. 2009, 60, 61–83. (33) Bloembergen, N.; Pershan, P. S. Light Waves at the Boundary of Nonlinear Media. Phys. Rev. 1962, 128, 606–622. 1060

dx.doi.org/10.1021/jz200251h |J. Phys. Chem. Lett. 2011, 2, 1056–1061

The Journal of Physical Chemistry Letters (34) Petersen, P. B.; Saykally, R. J. On the Nature of Ions at the Liquid Water Surface. Annu. Rev. Phys. Chem. 2006, 57, 333–364. (35) Zhao, X.; Ong, S.; Eisenthal, K. B. Polarization of Water Molecules at a Charged Interface. Second Harmonic Studies of Charged Monolayers at the Air/Water Interface. Chem. Phys. Lett. 1993, 202, 513–520. (36) Stack, A. G.; Higgins, S. R.; Eggleston, C. M. Point of Zero Charge of a CorundumWater Interface Probed with Optical Second Harmonic Generation (SHG) and Atomic Force Microscopy (AFM): Non-stoichiometric Approaches to Oxide Surface Charge. Geochim. Cosmochim. Acta 2001, 65, 3055–3063. (37) Salafsky, J. S.; Eisenthal, K. B. Protein Adsorption at Interfaces Detected by Second Harmonic Generation. J. Phys. Chem. B 2000, 104, 7752–7755. (38) Wei, X.; Hong, S.-C.; Zhuang, X.; Goto, T.; Shen, Y. R. Nonlinear Optical Studies of Liquid Crystal Alignment on a Rubber Polyvinyl Alcohol Surface. Phys. Rev. E 2000, 62, 5160–5172. (39) Briggman, K. A.; Stephenson, J. C.; Wallace, W. E.; Richter, L. J. Absolute Molecular Orientational Distribution of the Polystyrene Surface. J. Phys. Chem. B 2001, 105, 2785–2791. (40) Wang, J.; Buck, S.; Chen, Z. Sum Frequency Generation Vibrational Spectroscopy Studies on Protein Adsorption. J. Phys. Chem. B 2002, 106, 11666–11672. (41) 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. (42) Scatena, L. F.; Brown, M. G.; Richmond, G. L. Water at Hydrophobic Surfaces: Weak Hydrogen Bonding and Strong Orientation Effects. Science 2001, 292, 908–912. (43) Lambert, A. G.; Davies, P. B.; Neivandt, D. J. Implementing the Theory of Sum Frequency Generation Vibrational Spectroscopy: A Tutorial Review. Appl. Spectrosc. Rev. 2005, 40, 103–145. (44) Wang, J.; Buck, S. M.; Even, M. A.; Chen, Z. Molecular Responses of Proteins at Different Interfacial Environments Detected by Sum Frequency Generation Vibrational Spectroscopy. J. Am. Chem. Soc. 2002, 124, 13302–13305. (45) Watry, M. R.; Tarbuck, T. L.; Richmond, G. L. Vibrational SumFrequency Studies of a Series of Phospholipid Monolayers and the Associated Water Structure at the Vapor/Water Interface. J. Phys. Chem. B 2003, 107, 512–518. (46) Yeganeh, M. S.; Dougal, S. M.; Pink, H. S. Vibrational Spectroscopy of Water at Liquid/Solid Interfaces: Crossing the Isoelectric Point of a Solid Surface. Phys. Rev. Lett. 1999, 83, 1179–1182. (47) Lee, C. H.; Chang, R. K.; Bloembergen, N. Nonlinear Electroreflectance in Silicon and Silver. Phys. Rev. Lett. 1967, 18, 167– 170. (48) Aktsipetrov, O. A.; Fedyanin, A. A.; Dadap, J. I.; Downer, M. C. DC-Electric-Field-Induced Second Harmonic Generation Studies of Surfaces and Buried Interfaces of Column IV Semiconductors. Laser Phys. 1996, 6, 1142–1151. (49) Eichler, H.; Gunter, P.; Pohl, D. Laser Induced Dynamic Gratings; Springer: Berlin, 1986. (50) de Beer, A. G. F.; Campen, R. K.; Roke, S. Separating Surface Structure and Surface Charge with Second-Harmonic and Sum-Frequency Scattering. Phys. Rev. B 2010, 82, 235431. (51) Kataoka, S.; Gurau, M. C.; Albertorio, F.; Holden, M. A.; Lim, S. M.; Yang, R. D.; Cremer, P. S. Investigation of Water Structure at the TiO2/Aqueous Interface. Langmuir 2004, 20, 1662–1666. (52) Kim, J.; Kim, G.; Cremer, P. S. Investigations of Polyelectrolyte Adsorption at the Solid/Liquid Interface by Sum Frequency Spectroscopy: Evidence for Long-Range Macromolecular Alignment at Highly Charged Quartz/Water Interfaces. J. Am. Chem. Soc. 2002, 124, 8751–8756. (53) Morel, F. M. M. Principles of Aquatic Chemistry; John Wiley & Sons: New York, 1983. (54) Su, X.; Lianos, L.; Shen, Y. R.; Somorjai, G. A. Surface-Induced Ferroelectric Ice on Pt(111). Phys. Rev. Lett. 1998, 80, 1533–1536. (55) Ahmed, S. M. Dissociation of Oxide Surfaces at the LiquidSolid Interface. Can. J. Chem. 1966, 44, 1663–1670.

LETTER

(56) Abenroth, R. P. Behavior of a Pyrogenic Silica in Simple Electrolytes . J. Colloid Interface Sci. 1970, 34, 591–596. (57) Kitamura, A.; Fujiwara, K.; Yamamoto, T.; Nishikawa, S.; Moriyama, H. Analysis of Adsorption Behavior of Cations onto Quartz Surface by Electrical Double-Layer Model. J. Nucl. Sci. Technol. 1999, 36, 1167–1175.

1061

dx.doi.org/10.1021/jz200251h |J. Phys. Chem. Lett. 2011, 2, 1056–1061