You are on page 1of 7

Article

Cite This: J. Phys. Chem. C 2018, 122, 11385−11391 pubs.acs.org/JPCC

Water Adsorption on Hydrophilic and Hydrophobic Surfaces of


Silicon
Lei Chen,† Xin He,‡ Hongshen Liu,‡ Linmao Qian,*,† and Seong H. Kim*,†,‡

Tribology Research Institute, State Key Laboratory of Traction Power, Southwest Jiaotong University, Chengdu 610031, China

Department of Chemical Engineering and Materials Research Institute, The Pennsylvania State University, University Park,
Pennsylvania 16802, United States
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: The isotherm thickness and hydrogen-bonding interactions of


water layers adsorbed on hydrophilic and hydrophobic surfaces were quantified
and compared. The hydrophilic and hydrophobic surfaces were modeled with an
OH-terminated native oxide layer on silicon and a HF-etched silicon terminated
Downloaded via MONASH UNIV on July 24, 2022 at 14:14:52 (UTC).

with hydrogen, respectively. The silicon substrate allows the use of attenuated total
reflection infrared (ATR-IR) spectroscopy for quantitative measurement of
adsorbed water without interferences from the gas phase water. On the hydrophilic
Si−OH surface, the average thickness of the strongly hydrogen-bonded water layer
increases up to ∼2 molecular layers as relative humidity (RH) increases, beyond
which the weakly hydrogen-bonded structure is dominant. On the hydrophobic
Si−H surface, the adsorbed water layer consists predominantly of the weakly
hydrogen-bonded structure and its average thickness remains less than a
monolayer even at RH = 90%. The differences in the thickness and structure of adsorbed water layers on hydrophilic versus
hydrophobic surfaces found from ATR-IR measurements provide critical insights needed for better understanding of various
physical processes affected by water adsorption in ambient conditions.

■ INTRODUCTION
Most inorganic materials exposed to ambient air adsorb water,
surface.26 For example, simulations predicted that water
molecules on graphite do not have any hydrogen bond
and such adsorbed water layers can play a crucial role in interactions with the surface.27
biology,1,2 materials science,3,4 and tribology.5−7 The unique This article reports spectroscopic evidence on differences in
properties of water in the liquid state stem from its large dipole structural configuration of water molecules adsorbed on
moment, high polarizability, and hydrogen bonding capabil- hydrophilic and hydrophobic surfaces of silicon. The thickness
ity.8,9 In the case of water adsorbed on solids, additional and hydrogen-bonding interactions of water adsorbed on
interactions with the solid surface make its property or hydrophilic native oxide surface (terminated with Si−OH,
structure deviate from the liquid water.10,11 Understanding water contact angle < 5°) and hydrophobic HF-etched surface
the structure of interfacial water is of fundamental interest in (terminated with SiH, water contact angle ≈ 83°) were
many physical processes such as separation of hydrophilic measured as a function of RH using attenuated total reflectance
surfaces or macromolecules in water,12 fabrication of uniform infrared (ATR-IR) spectroscopy. The silicon substrate allows
films in atomic layer deposition,13 or control of surface wear the use of ATR-IR spectroscopy for measurement of adsorbed
and interfacial friction of solid materials.14−16 water without interferences from the gas phase water. Also,
The isotherm thickness and structure of adsorbed water films silicon is a technically important material in various scientific
can vary substantially depending on the surface chemistry and and engineering applications.28−32 One advantage of ATR-IR
the relative humidity (RH).17 On hydrophilic silicon oxide over other techniques such as polarization-modulation
surfaces, the adsorbed water layer can form strongly hydrogen- reflection−absorption infrared spectroscopy (PM-RAIRS) is
bonded or ordered structures, which was often denoted as that it allows quantitative determination of the adsorbed water
solid-like water, ice-like water, or quasi-ice.18,19 The formation layer thickness.33 In the past, silane-based self-assembled
of ordered structures is in agreement with density functional monolayers (SAMs) were frequently used for modification of
theory (DFT) and molecular dynamics (MD) simulations.20−22 silicon surfaces; however, the use of SAMs complicates spectral
With the increase of RH (or partial pressure of water vapor), analysis due to ingress of water to subsurface sites.17,34,35 For
more disordered or liquid-like structure grows in the adsorbed this reason, the Si−OH surface of the native oxide on Si and the
water layer. The double-layer structure was also reported for
mica and metal surfaces.23−25 At hydrophobic surfaces, MD Received: February 22, 2018
simulations showed that water molecules have a diffusive Revised: May 2, 2018
behavior due to the lack of strong interactions with the Published: May 4, 2018

© 2018 American Chemical Society 11385 DOI: 10.1021/acs.jpcc.8b01821


J. Phys. Chem. C 2018, 122, 11385−11391
The Journal of Physical Chemistry C Article

Figure 1. ATR-IR spectra of the O−H stretching region of water adsorbed on (a) Si−OH and (b) Si−H surfaces at RHs of 15%, 30%, 45%, 60%,
75%, and 90%. The insets schematically show the Si−OH functional groups on a native oxide layer and the SiH groups on the HF-etched surface. Si,
blue; O, red; H, gray. The O−H stretching region is fitted with two components: strongly H-bonded water (blue lines centered at 3200−3275 cm−1)
and weakly H-bonded water (green lines centered at 3400−3450 cm−1). Note the log(1/R) full-scale for Si−OH in (a) is four times larger than that
for Si−H in (b). (c) ATR-IR spectra of the SiH stretch region of Si−H and Si−OH surfaces measured in dry nitrogen. Peak deconvolution is shown
in Figure S2 in the SI. (d) ATR-IR spectrum of bulk liquid water on the Si−OH surface.

Si−H surface obtained right after the HF etching were used to at the surface of fused quartz are hydroxylated, then the areal
model hydrophilic and hydrophobic surface chemistry, density of OH groups is roughly 5/nm2.38
respectively. The ATR-IR analysis of these surfaces quantita- A Thermo-Nicolet Nexus 670 spectrometer equipped with a
tively shows how the isotherm thickness and hydrogen-bonding MCT-A detector and a Harrick ATR setup was used to collect
interactions in the adsorbed water layer vary with the surface ATR-IR spectra of water layers on those two surfaces as RH
chemistry of silicon-based materials. varied from 0% (dry nitrogen) to 90% at room temperature


(20.8 ± 0.5 °C). After cleaning, the silicon crystal was promptly
EXPERIMENTAL DETAILS mounted to the ATR assembly and purged with dry nitrogen
until there was no change in the OH stretch and H2O bending
A double-side-polished p-doped Si(100) wafer (∼725 μm vibration regions. The single beam spectrum collected in dry
thick) was cut into the ATR crystal shape (50 mm × 10 mm, nitrogen (RH = 0%) was used as a background spectrum for all
with 45° bevel-cut and polish in both ends for IR entrance and subsequent data collections. The spectra were collected with a 4
exit). The total reflection of IR beams from the probe surface cm−1 resolution and averaged over 100 scans. At an incident
was ∼35 times. One crystal was prepared by cleaning with the angle of 45°, the effective penetration depth of the evanescent
RCA-1 solution (aqueous solution with H2O2 and NH3, 80 wave was 482 nm at 1635 cm−1.39
°C), rinsing with DI water, and drying with nitrogen. After that,
the crystal was exposed to UV/O3 for 30 min, which produced
an organic-contaminant-free native oxide surface.36 This sample
■ RESULTS AND DISCUSSION
The ATR-IR spectra of adsorbed water layers measured during
is named as “Si−OH”. Another ATR crystal was treated with the stepwise increase of RH from 9% to 90% on the hydrophilic
wet etching in hydrofluoric acid (∼40 wt % HF in aqueous Si−OH and hydrophobic Si−H surfaces are reported in Figure
solution) for about 5 min, followed by rinsing with DI water 1a,b. The SiH stretching vibration region of the ATR-IR spectra
and drying with nitrogen. This crystal is denoted as “Si−H”. X- (Figure 1c) confirms the presence of SiH (2082 cm−1) and
ray photoelectron spectroscopy (XPS) analysis confirmed that SiH2 (2103 and 2121 cm−1) species at the Si−H surface.40,41
the oxygen content at the surface is below the detection limit After water adsorption measurement at 90% RH, the area of
(see Figure S1 in the Supporting Information, SI). The water SiH stretching vibration peak is still >98% of the peak area
contact angles were measured to be <5° on the Si−OH surface measured before the 9% RH exposure (see Figure S2 in
and ∼83° on the Si−H surface (measured immediately after Supporting Information). This result implies that the oxidation
HF treatment). When the silicon oxide surface is saturated with of the Si−H surface due to exposure to water vapor at room
hydroxyl groups, the water contact angle decreases to 0°.17 temperature is negligible over the duration of the measurement
Using the Cassie equation, it is estimated that the Si−OH time (typically, ∼5 h for a full series from 9% RH to 90% RH
surface is ∼98% saturated with silanol groups.37 If all Si atoms measurements), and most SiH groups remain intact. For
11386 DOI: 10.1021/acs.jpcc.8b01821
J. Phys. Chem. C 2018, 122, 11385−11391
The Journal of Physical Chemistry C Article

comparison, the ATR-IR spectrum of liquid water is also shown


in Figure 1d. Note that the spectral feature of the bulk liquid is
not sensitive to the surface chemistry since the penetration
depth of evanescent wave into liquid water is several orders of
magnitude larger than the thickness of the adsorbed water
layer.42
The difference in hydrogen-bonding interactions of water
molecules adsorbed on Si−OH and Si−H surfaces are
manifested as variances in peak shape of the O−H stretch
mode in the range from 3000 to 3800 cm−1 (Figure 1a,b).43,44
As the strength of hydrogen bonding interaction increases, the
O−H stretching peak position shows a larger red-shift.45
Although the exact peak shape is a complicated function of
hydrogen bond dynamics,46 the O−H stretch mode of water in Figure 2. Area ratio of peak-I (centered at 3200−3275 cm−1) and
this region can be fitted with two main components: (I) the peak-II (centered at 3400−3450 cm−1) as a function of RH on the
strongly hydrogen-bonded configuration at 3200−3275 cm−1 hydrophilic Si−OH and hydrophobic Si−H surfaces (estimated from
and (II) the weakly hydrogen-bonded configuration at 3400− Figure 1a,b). The solid lines are drawn to guide eyes. The dotted line
marks the area ratio of those two peaks for bulk liquid water
3450 cm−1.37,46,47 The IR spectrum of ice water measured at
(estimated from Figure 1d). The ATR-IR measurements were
<273 K has a major peak centered at ∼3220 cm−1 (same as repeated 5 times for the Si−OH surface and 4 times for the Si−H
peak-I) with a relatively narrow full-width-half-maximum surface.
(fwhm) of ∼275 cm−1.48,49 Based on this comparison, the
peak-I component is often called the ice-like or solid-like
structure; in short, it just means more ordered or networked water Not only the structural configuration but also the thickness
due to strong hydrogen bond interactions. Then, the peak-II of adsorbed water depend greatly on the outermost surface
component could be called the disordered structure to contrast chemistry.17 In ATR-IR, the thickness of the adsorbed water
with the peak-I component. layer (d) can be estimated by comparing the integrated areas of
This argument can be used to analyze the ATR-IR spectral adsorbed species (Aads) with the peak area of bulk water (Abulk)
features of adsorbed water on hydrophilic Si−OH and with consideration of the penetration depth (dbulk) of the
hydrophobic Si−H surfaces. On the Si−OH surface, peak-I evanescent IR beam:53
(blue line in Figure 1a) is dominant over peak-II (green line) at A ads A λ
low RH conditions, implying that most adsorbed water d= dbulk = ads
Abulk Abulk 2π (n1 sin θ − n2 2)1/2
2 2
(1)
molecules have strong hydrogen-bonding interactions and
assume the solid-like configuration. As RH increases, the Here, λ is the wavelength of IR, θ is the incident angle of the IR
weakly hydrogen-bonded disordered structure (peak-II) grows beam (45°), and n1 and n2 are the refractive indices of the
along with the solid-like configuration. This is in good substrate and bulk water, respectively. The exact refractive
agreement with previous reports on the water adsorption on index of the adsorbed layer is not known, but it is assumed here
silicon oxide surface50 and other hydrophilic surfaces.51,52 In to be the same as the bulk value. The adsorbed water layer
contrast, the Si−H surface shows much larger intensity of the constitutes with the more-ordered (peak-I) and disordered
peak-II component (green line in Figure 1b) compared to the (peak-II) components; so the total thickness of absorbed water
peak-I component (blue line) in the entire RH range studied layer (d) can be expressed as the sum of the ordered water
(Figure 1b), indicating that the weakly hydrogen-bonded thickness (dordered) and disordered water thickness (ddisordered).
disordered structure is dominant at all times. The thickness of each component was calculated using eq 1
Comparison with the ATR-IR spectrum of bulk liquid water with the refractive index of bulk ice (n2 = 1.30)54 and the area
(Figure 1d) reveals that the structural configurations of the of peak-I for dordered; and the refractive index of bulk liquid (n2 =
adsorbed water layers on the hydrophilic Si−OH and 1.33)55 and the area of peak-II for ddisordered. Figure 3 displays
hydrophobic Si−H surfaces (Figure 1a,b) are drastically the thickness (d) of the adsorbed water layer on the Si−OH
different from that of the liquid water.17 The relative ratio of and Si−H surfaces. The calculated thickness is compared with
peak-I to peak-II is calculated from the areas of the the molecular diameter of water (2.82 Å),50 which could be
deconvoluted peaks at various RH conditions. This ratio for taken as the thickness of one monolayer.
the bulk liquid water is ∼0.72 in Figure 1d (marked with a On the hydrophilic Si−OH surface (Figure 3a), the total
dotted line in Figure 2). With the increase of RH, the I/II ratio thickness (d) of the adsorbed water layer grows rapidly to
of the water layer adsorbed on Si−OH decreases from ∼3 at about a full monolayer at RH ≈ 9%, then slowly in the
RH = 9% to ∼1.5 at RH = 90% (Figure 2). This result indicates intermediate RH region (for example, 20−70%), and then fast
that on the hydrophilic Si−OH surface, the water layer mostly again as RH approaches saturation (up to ∼4 molecular layers
assumes the strongly hydrogen-bonded solid-like network at at 90% RH). Based on the calculated thickness of the ordered
low RHs, and then its structure gradually changes toward the (dordered) and disordered (ddisordered) components, the structural
bulk liquid network as RH approaches the saturation point. In configuration of the adsorbed water layer could be deduced.
contrast, the I/II ratio of the water layer on the Si−H surface The first monolayer formed at low RH is highly ordered; this
remains relatively constant at 0.3−0.5 in the entire RH range must be due to strong hydrogen bond interactions with the
(Figure 2). On the hydrophobic Si−H surface, the solid-like surface OH groups. The ordered structure induced by the
structure is minor, and its relative abundance does not change interactions with the surface appears to propagate up to ∼2
with RH. These results are congruent with MD simulation layers, consistent with the water configuration at a fully
results.27 hydroxylated surface calculated in MD simulations.21 A similar
11387 DOI: 10.1021/acs.jpcc.8b01821
J. Phys. Chem. C 2018, 122, 11385−11391
The Journal of Physical Chemistry C Article

thickness of the adsorbed water molecules from ATR-IR is


nearly half a monolayer (Figure 3b), the water clusters near
defect sites must be three-dimensional (3D) islands, instead of
two-dimensional patches covering a large area of the surface.
The structural configuration of the adsorbed water layers
implicated by the ATR-IR data is schematically illustrated in
Figure 4. On the hydrophilic Si−OH surface, the water
molecules in the ordered layer are under strong influences of
the surface hydroxyl groups, and the disordered layer exists at
the air-side of the adsorbate layer (Figure 4a). This
configuration is congruent with the structures previously
proposed from other spectroscopic observations19,49,60 and
MD simulations.61,62 The ordered water is likely to be a fully
Figure 3. Adsorption isotherm of adsorbed water on (a) hydrophilic connected network at the monolayer coverage, based on MD
Si−OH and (b) hydrophobic Si−H surfaces. The circle and triangle simulations predicting the formation of 2D ice-like water
symbols represent the total thickness of the adsorbed water layer layer.63,64 In the case of the hydrophobic Si−H surface, the
calculated based on eq 1. The solid line is drawn to guide eyes. The adsorbed water layer is very little, and they are likely to exist in
average and standard deviation values of adsorbed water thickness are
estimated from the ATR-IR measurements repeated 5 times on the clusters.65 Those clusters may be formed near defect sites that
Si−OH surface and 4 times on the Si−H surface. Note that the may have some polarity or contain OH groups. Based on the
thickness is an estimation based on eq 1; the value will vary depending spectral shape of the OH stretch mode, the structure of these
on how dbulk is defined in eq 1. The number of layers is calculated by clusters appears to be highly disordered or dynamically
dividing the average thickness with the average diameter of water changing, regardless of RH conditions of the environment
molecule (2.82 Å). (Figure 2).66 The total averaged thickness is less than the full
monolayer coverage even at 90% RH.67 This is due to the lack
trend was also observed on metal and metal oxide surfaces both of hydrogen bonding interactions with the Si−H surface.
in experimental measurements and MD simulations, except the The ATR-IR analysis of water adsorption provides the
degree of order in the first monolayer.55−58 The disordered structure and thickness information on the physisorbed water
structure on the hydrophilic Si−OH surface becomes layers on the hydrophilic Si−OH and hydrophobic Si−H
significant when RH increases above 60%. Because the ordered surfaces in equilibrium with the gas phase water, which could be
structure is induced by the surface OH groups, the disordered related to various physical processes occurring such surfaces.
structure is expected to be formed on top of the ordered layer, For example, the protein adsorption on biomaterials is a
not under the ordered layer. function of surface wettability.68 Hydrophilic surfaces with a
On the hydrophobic Si−H surface (Figure 3b), the water contact angle less than 60° show a resistance to protein
calculated average thickness of the adsorbed water layer adsorption, whereas a hydrophobic surface can readily adsorb
remains less than half a monolayer in RH ≤ 75% conditions. proteins in aqueous solution.68 This could be attributed to the
Previously, PM-RAIRS found that water does not adsorb on the
hydrogen bond strength of water molecules at or near the
CH3-terminated hydrophobic SAMs on gold (water contact
surface (Figure 4). At the hydrophilic surface, water molecules
angle >100°) even at near-saturation RH.33 In the case of the
Si−H surface, the silicon substrate with high polarizability may are associated into a strongly hydrogen bonded network, which
reduce the water contact angle to 83°, compared to the CH3- may not be displaced easily. In contrast, the weakly associated
terminated SAM surface. However, this may not be sufficient to disordered water layers near the hydrophobic surface could be
hold a full monolayer of water on the Si−H surface (Figure 3b). easily displaced by protein molecules, changing the interfacial
Considering the lack of hydrogen bonding interactions between energy. Furthermore, the lack of hydrogen bonding interactions
the water molecule and the Si−H surface, the physisorbed between water molecules and a hydrophobic surface may be the
water molecules are likely to form clusters via hydrogen main cause for the formation of a lower density “depleted”
bonding interactions with each other.59 Such clusters may form interface at the hydrophobic surface in liquid water.69,70 The
around defect sites at the Si−H surface (such as a trace amount energetic instability of such interfaces may play a significant role
of Si−OH groups). Since the concentration of oxidized Si sites in collapsing or coagulation of hydrophobic particles in aqueous
is below the detection limit of XPS (Figure S1) and the average solution.71,72

Figure 4. Schematic illustrating the structural of adsorbed water layer on (a) hydrophilic Si−OH and (b) hydrophobic Si−H surfaces. The dotted
yellow lines represent hydrogen bonds. Note that illustration is not to scale. In (a), the water layer adsorbed on the hydrophilic Si−OH surface
grows up to full coverage at RH < 10% with highly ordered structures. Inset in (b) shows the schematic of clusters of water molecules on the
hydrophobic Si−H surface.

11388 DOI: 10.1021/acs.jpcc.8b01821


J. Phys. Chem. C 2018, 122, 11385−11391
The Journal of Physical Chemistry C Article

The strongly hydrogen-bonded water structure would play a


pivotal role in interfacial water flow. MD simulations showed
■ AUTHOR INFORMATION
Corresponding Authors
that a hydrodynamic slip of water can occur even on
hydrophilic surfaces depending on the surface density of *E-mail: linmao@swjtu.edu.cn.
hydrophilic sites.73 If the OH groups on the Si−OH surface are *E-mail: shkim@engr.psu.edu.
assumed to be distributed in a hexagonal array pattern (a close- ORCID
packed array with equal distance among each other), then the Seong H. Kim: 0000-0002-8575-7269
areal density of 5/nm2 would correspond to the distance Notes
between the nearest OH groups of ∼0.48 nm.38 This is ∼1.7
The authors declare no competing financial interest.
times the size of the water molecule.50 Because the silica surface
is known to be a nonslip surface for water, this distance might
be too far to allow the slip of water molecules among adjacent
OH sites,73 but it is close enough to induce a strongly hydrogen
■ ACKNOWLEDGMENTS
This work was financially supported by the Natural Science
bonded network covering the entire surface, as evidenced by Foundation of China (51505391, 51527901) and the
ATR-IR (Figures 1−3). Fundamental Research Funds for the Central Universities
Also, the structure of the adsorbed water layer is found to be (2682016CX026). ATR-IR measurements were carried out
important in nanotribology. The adhesion and friction forces with the support from the National Science Foundation of the
measured at nanoscale are found to be strongly dependent on USA (Grant No. DMR-1609107 and CMMI-1435766).


the amount of ordered water layer formed around the annulus
of the contact area on the hydrophilic surface, which is a REFERENCES
function of RH in the environment. However, they are much
(1) Israelachvili, J.; Wennerström, H. Role of Hydration and Water
smaller and relatively independent of RH on the hydrophobic
Structure in Biological and Colloidal Interactions. Nature 1996, 379,
surface because the adsorbed water layer is so thin (less than a 219−225.
monolayer in average, Figure 3) and its structure does not (2) Klein, J. Repair or Replacement-A Joint Perspective. Science 2009,
change (Figure 2).74−76 Another example is water-assisted 323 (5910), 47−48.
mechanochemical reactions of silicon materials. Compared to (3) Moeremans, B.; Cheng, H. W.; Hu, Q. Y.; Garces, H. F.; Padture,
the disordered water layer, the ordered water layer appears to N. P.; Renner, F. U.; Valtiner, M. Lithium-ion Battery Electrolyte
facilitate mechanochemical reactions that lead to hydrolysis of Mobility at Nano-confined Graphene Interfaces. Nat. Commun. 2016,
silicon oxide surfaces, which is manifested as surface wear.77,78 7, 12693.
A similar mechanism may pertain to the mechanochemical wear (4) Stamenkovic, V. R.; Strmcnik, D.; Lopes, P. P.; Markovic, N. M.
Energy and Fuels from Electrochemical Interfaces. Nat. Mater. 2017,
of multicomponent silicate glass materials.79 The mechano-
16, 57−69.
chemical wear of hydrophilic silicon increases with the growth (5) Raviv, U.; Giasson, S.; Kampf, N.; Gohy, J.; Jérôme, R.; Klein, J.
of the ordered water layer, but this is somewhat suppressed Lubrication by Charged Polymers. Nature 2003, 425, 163−165.
upon the growth of disordered water layers at high RH.80,81 (6) Klein, J. Hydration Lubrication. Friction 2013, 1 (1), 1−23.


(7) Chen, J. Y.; Ratera, I.; Park, J. Y.; Salmeron, M. Velocity
CONCLUSIONS Dependence of Friction and Hydrogen Bonding Effects. Phys. Rev. Lett.
2006, 96, 236102.
The structure and isotherm thickness of physisorbed water (8) Ben-Naim, A. Hydrophobic Interactions; Plenum Press: New York,
layers strongly depend on the surface chemistry of the substrate 1980.
and RH in the surrounding environment. ATR-IR analysis of (9) Raschke, T. M. Water Structure and Interactions with Protein
the adsorbed water in equilibrium with the gas phase provided Surfaces. Curr. Opin. Struct. Biol. 2006, 16, 152−159.
the spectroscopic evidence for the strongly hydrogen-bonded (10) Fecko, C. J.; Eaves, J. D.; Loparo, J. J.; Tokmakoff, A.; Geissler,
ordered network of water molecules on the hydrophilic Si−OH P. L. Ultrafast Hydrogen-Bond Dynamics in the Infrared Spectroscopy
of Water. Science 2003, 301, 1698−1702.
surface and the lack or rarity of such structures on the (11) Wang, H. J.; Xi, X. K.; Kleinhammes, A.; Wu, Y. Temperature-
hydrophobic Si−H surface. As RH increases, the thickness of Induced Hydrophobic-Hydrophilic Transition Observed by Water
the ordered water layer on the hydrophilic Si−OH surface Adsorption. Science 2008, 322, 80−83.
reaches a full monolayer at <10% RH and continues growing up (12) Chandler, D. Interfaces and the Driving Force of Hydrophobic
to about two monolayers. In contrast, the disordered structure Assembly. Nature 2005, 437, 640−647.
component grows slowly with RH, reaching a monolayer (13) Strempel, V. E.; Naumann d’Alnoncourt, R.; Driess, M.;
coverage at RH > 60%, and increases fast as RH approaches the Rosowski, F. Atomic Layer Deposition on Porous Powders with in situ
saturation value. The water layer on the hydrophobic Si−H Gravimetric Monitoring in a Modular Fixed Bed Reactor Setup. Rev.
Sci. Instrum. 2017, 88, 074102.
surface is highly disordered and less than a monolayer
(14) Wang, X. D.; Guo, J.; Chen, C.; Chen, L.; Qian, L. M. A Simple
(probably forming isolated clusters) even at RH = 90%. Method to Control Nanotribology Behaviors of Monocrystalline


*
ASSOCIATED CONTENT
S Supporting Information
Silicon. J. Appl. Phys. 2016, 119, 044304.
(15) Chen, L.; Kim, S. H.; Wang, X. D.; Qian, L. M. Running-in
Process of Si-SiO2/SiO2 Pair at Nanoscale–Sharp Drops in Friction
and Wear rate During Initial Cycles. Friction 2013, 1 (1), 81.
The Supporting Information is available free of charge on the (16) Chen, L.; Yang, Y. J.; He, H. T.; Kim, S. H.; Qian, L. M. Effect of
ACS Publications website at DOI: 10.1021/acs.jpcc.8b01821. Coadsorption of Water and Alcohol Vapor on the Nanowear of
Silicon. Wear 2015, 332, 879−884.
XPS measurements on Si−H and Si−OH surfaces, ATR- (17) Asay, D. B.; Barnette, A. L.; Kim, S. H. Effects of Surface
IR analysis of the SiH stretch region of the Si−H surface Chemistry on Structure and Thermodynamics of Water Layers at
(PDF) Solid-Vapor Interfaces. J. Phys. Chem. C 2009, 113, 2128−2133.

11389 DOI: 10.1021/acs.jpcc.8b01821


J. Phys. Chem. C 2018, 122, 11385−11391
The Journal of Physical Chemistry C Article

(18) Midya, U. S.; Bandyopadhyay, S. Hydration Behavior at the Ice- using Ultraviolet Cleaning and HF Etching. J. Appl. Phys. 1988, 64,
Binding Surface of the Tenebrio molitor Antifreeze Protein. J. Phys. 3516.
Chem. B 2014, 118, 4743−4752. (41) Chabal, Y. J.; Higashi, G. S.; Raghavachari, K. Infrared
(19) Verdaguer, A.; Weis, C.; Oncins, G.; Ketteler, G.; Bluhm, H.; Spectroscopy of Si(111) and Si(100) Surfaces after HF Treatment:
Salmeron, M. Growth and Structure of Water on SiO2 Films on Si Hydrogen Termination and Surface Morphology. J. Vac. Sci. Technol.,
Investigated by Kelvin Probe Microscopy and in Situ X-ray A 1998, 7, 2104.
Spectroscopies. Langmuir 2007, 23, 9699−9703. (42) Nguyen, T.; Byrd, E.; Lin, C. J. A Spectroscopic Technique for
(20) Yang, J. J.; Meng, S.; Xu, L. F.; Wang, E. G. Ice Tessellation on a in situ Measurement of Water at the Coating/Metal Interface. J. Adhes.
Hydroxylated Silica Surface. Phys. Rev. Lett. 2004, 92, 146102. Sci. Technol. 1991, 5, 697−709.
(21) Argyris, D.; Cole, D. R.; Striolo, A. Hydration Structure on (43) Scherer, J. R. Advances in Infrared and Raman Spectroscopy;
Crystalline Silica Substrates. Langmuir 2009, 25 (14), 8025−8035. Heyden: London, 1978.
(22) Argyris, D.; Cole, D. R.; Striolo, A. Dynamic Behavior of (44) Wagner, R.; Benz, S.; Mohler, O.; Saathoff, H.; Schnaiter, M.;
Interfacial Water at the Silica Surface. J. Phys. Chem. C 2009, 113, Schurath, U. Mid-infrared Extinction Spectra and Optical Constants of
19591−19600. Supercooled Water Droplets. J. Phys. Chem. A 2005, 109, 7099−7112.
(23) Odelius, M.; Bernasconi, M.; Parrinello, M. Two Dimensional (45) Dey, A.; Mondal, S. I.; Sen, S.; Ghosh, D.; Patwari, G. N.
Ice Adsorbed on Mica Surface. Phys. Rev. Lett. 1997, 78, 2855−2858. Electrostatics Determine Vibrational Frequency Shifts in Hydrogen
(24) Zhao, G. T.; Tan, Q. Y.; Xiang, L.; Cai, D.; Zeng, H. B.; Yi, H.;
Bonded Complexes. Phys. Chem. Chem. Phys. 2014, 16, 25247.
Ni, Z. H.; Chen, Y. F. Structure and Properties of Water Film
(46) Mallamace, F.; Broccio, M.; Corsaro, C.; Faraone, A.; Majolino,
Adsorbed on Mica Surfaces. J. Chem. Phys. 2015, 143, 104705.
D.; Venuti, V.; Liu, L.; Mou, C. Y.; Chen, S. H. Evidence of the
(25) Mu, R. T.; Zhao, Z. J.; Dohnálek, Z.; Gong, J. L. Structural
Motifs of Water on Metal Oxide Surfaces. Chem. Soc. Rev. 2017, 46, Existence of the Low-density Liquid Phase in Supercooled, Confined
1785−1806. Water. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 424−428.
(26) Willard, A. P.; Chandler, D. The Molecular Structure of the (47) Ewing, G. E. Thin Film Water. J. Phys. Chem. B 2004, 108,
Interface between Water and a Hydrophobic Substrate is Liquid-Vapor 15953.
Like. J. Chem. Phys. 2014, 141, 18C519. (48) Irvine, W. M.; Pollack, J. B. Infrared Optical Properties of Water
(27) Gordillo, M. C.; Martl,́ J. Molecular Dynamics Description of a and Ice Spheres. Icarus 1968, 8, 324.
Layer of Water Molecules on a Hydrophobic Surface. J. Chem. Phys. (49) Du, Q.; Freysz, E.; Shen, Y. R. Surface Vibrational Spectroscopic
2002, 117, 3425. Studies of Hydrogen Bonding and Hydrophobicity. Science 1994, 264,
(28) Tilli, M.; Lindroos, V.; Airaksinen, V. M.; Franssila, S.; Paulasto- 826−828.
Krockel, M.; Lehto, A.; Motooka, T. Handbook of Silicon Based MEMS (50) Asay, D. B.; Kim, S. H. Evolution of the Adsorbed Water Layer
Materials and Technologies; Elsevier: London, 2010. Structure on Silicon Oxide at Room Temperature. J. Phys. Chem. B
(29) Kim, S. H.; Asay, D. B.; Dugger, M. T. Nanotribology and 2005, 109, 16760−16763.
MEMS. Nano Today 2007, 2 (5), 22−29. (51) Thiel, P. A.; Madey, T. E. The Interaction of Water with Solid-
(30) Chen, L.; Wen, J.; Zhang, P.; Yu, B.; Chen, C.; Ma, T.; Lu, X.; surfaces: Fundamental Aspects. Surf. Sci. Rep. 1987, 7, 211.
Kim, S. H.; Qian, L. Nanomanufacturing of silicon surface with a single (52) Beaglehole, D.; Christenson, H. K. Vapor Adsorption on Mica
atomic layer precision via mechanochemical reactions. Nat. Commun. and Silicon: Entropy Effects, Layering, and Surface Forces. J. Phys.
2018, 9, 1542. Chem. 1992, 96, 3395−3403.
(31) Qi, Y. Q.; Chen, L.; Jiang, S. L.; Yu, J. X.; Yu, B. J.; Qian, L. M. (53) Barnette, A. L.; Kim, S. H. Coadsorption of n-Propanol and
Investigation of Silicon Wear against Non-porous and Micro-porous Water on SiO2: Study of Thickness, Composition, and Structure of
SiO2 Spheres in Water and in Humid air. RSC Adv. 2016, 6, 89627− Binary Adsorbate Layer Using Attenuated Total Reflection Infrared
89634. (ATR-IR) and Sum Frequency Generation (SFG) Vibration Spec-
(32) Chen, L.; Xiao, C.; He, X.; Yu, B. J.; Kim, S. H.; Qian, L. M. troscopy. J. Phys. Chem. C 2012, 116, 9909−9916.
Friction and Tribochemical Wear Behaviors of Native Oxide Layer on (54) Warren, S. G. Optical Constants of Ice from the Ultraviolet to
Silicon at Nanoscale. Tribol. Lett. 2017, 65, 139. the Microwave. Appl. Opt. 1984, 23 (8), 1206−1225.
(33) Tu, A.; Kwag, H. R.; Barnette, A. L.; Kim, S. H. Water (55) Daimon, M.; Masumura, A. Measurement of the Refractive
Adsorption Isotherms on CH3-, OH-, and COOH-Terminated Index of Distilled Water from the Near-infrared Region to the
Organic Surfaces at Ambient Conditions Measured with PM-RAIRS. Ultraviolet region. Appl. Opt. 2007, 46 (18), 3811−3820.
Langmuir 2012, 28, 15263−15269. (56) Phan, A.; Ho, T. A.; Cole, D. R.; Striolo, A. Molecular Structure
(34) Schwendel, D.; Hayashi, T.; Dahint, R.; Pertsin, A.; Grunze, M.; and Dynamics in Thin Water Films at Metal Oxide Surfaces:
Steitz, R.; Schreiber, F. Interaction of Water with Self-assembled Magnesium, Aluminum, and Silicon Oxide Surfaces. J. Phys. Chem. C
Monolayers: Neutron Reflectivity Measurements of the Water Density
2012, 116, 15962−15973.
in the Interface region. Langmuir 2003, 19 (6), 2284−2293. (57) Phan, A.; Cole, D. R.; Striolo, A. Liquid Ethanol Simulated on
(35) Scatena, L. F.; Brown, M. G.; Richmond, G. L. Water at
Crystalline Alpha Alumina. J. Phys. Chem. B 2013, 117, 3829−3840.
Hydrophobic Surfaces: Weak Hydrogen Bonding and Strong
(58) Carrasco, J.; Hodgson, A.; Michaelides, A. A Molecular
Orientation Effects. Science 2001, 292, 908−912.
Perspective of Water at Metal Interfaces. Nat. Mater. 2012, 11,
(36) Graubner, V.-M.; Jordan, R.; Nuyken, O.; Schnyder, B.; Lippert,
T.; Kotz, R.; Wokaun, A. Photochemical Modification of Cross-linked 667−674.
Poly (Dimethylsiloxane) by Irradiation at 172 nm. Macromolecules (59) Björneholm, O.; et al. Water at Interfaces. Chem. Rev. 2016, 116,
2004, 37, 5936. 7698−7726.
(37) Cassie, A. B. D. Contact Angles. Discuss. Faraday Soc. 1948, 3, (60) Nijem, N.; Canepa, P.; Kaipa, U.; Tan, K.; Roodenko, K.;
11. Tekarli, S.; Halbert, J.; Oswald, I. W. H.; Arvapally, R. K.; Yang, C.;
(38) Banerjee, J.; Kim, S. H.; Pantano, C. G. Elemental Areal Density Thonhauser, T.; Omary, M. A.; Chabal, Y. J. Water Cluster
Calculation and Oxygen Speciation for Flat Glass Surfaces using x-ray Confinement and Methane Adsorption in the Hydrophobic Cavities
Photoelectron Spectroscopy. J. Non-Cryst. Solids 2016, 450, 185−193. of a Fluorinated Metal−Organic Framework. J. Am. Chem. Soc. 2013,
(39) Du, Q.; Superfine, R.; Freysz, E.; Shen, Y. R. Vibrational 135, 12615−12626.
Spectroscopy of Water at the Vapor/water Interface. Phys. Rev. Lett. (61) Aarts, I. M. P.; Pipino, A. C. R.; Hoefnagels, J. P. M.; Kessels, W.
1993, 70, 2313−2316. M. M.; van de Sanden, M. C. M. Quasi-Ice Monolayer on Atomically
(40) Takahagi, T.; Nagai, I.; Ishitani, A.; Kuroda, H.; Nagasawa, Y. Smooth Amorphous SiO2 at Room Temperature Observed with a
The Formation of Hydrogen Passivated Silicon Single-crystal Surfaces High-Finesse Optical Resonator. Phys. Rev. Lett. 2005, 95, 166104.

11390 DOI: 10.1021/acs.jpcc.8b01821


J. Phys. Chem. C 2018, 122, 11385−11391
The Journal of Physical Chemistry C Article

(62) Smirnov, K. S. A Molecular Dynamics Study of the Interaction


of Water with the External Surface of Silicalite-1. Phys. Chem. Chem.
Phys. 2017, 19, 2950.
(63) Gupta, P. K.; Meuwly, M. Dynamics and Vibrational
Spectroscopy of Water at Hydroxylated Silica Surfaces. Faraday
Discuss. 2014, 167, 329.
(64) Cimas, Á .; Tielens, F.; Sulpizi, M.; Gaigeot, M. P.; Costa, D. The
Amorphous Silica−Liquid Water Interface Studied by ab Initio
Molecular Dynamics (AIMD): Local Organization in Global Disorder.
J. Phys.: Condens. Matter 2014, 26, 244106.
(65) Odelius, M.; Bernasconi, M.; Parrinello, M. Two Dimensional
Ice Adsorbed on Mica Surface. Phys. Rev. Lett. 1997, 78, 2855−2858.
(66) Nauta, K.; Miller, R. E. Formation of Cyclic Water Hexamer in
Liquid Helium: The Smallest Piece of Ice. Science 2000, 287, 293−295.
(67) Yang, J. J.; Meng, S.; Xu, L. F.; Wang, E. G. Water Adsorption
on Hydroxylated Silica Surfaces Studied using the Density Functional
Theory. Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 71, 035413.
(68) Ohba, T.; Kanoh, H.; Kaneko, K. Cluster-growth-induced Water
Adsorption in Hydrophobic Carbon Nanopores. J. Phys. Chem. B
2004, 108, 14964−14969.
(69) Vogler, E. A. Protein Adsorption in Three Dimensions.
Biomaterials 2012, 33, 1201−1237.
(70) Mezger, M.; Reichert, H.; Schöder, S.; Okasinski, J.; Schröder,
H.; Dosch, H.; Palms, D.; Ralston, J.; Honkimäki, V. High-resolution
in situ x-ray Study of the Hydrophobic Gap at the Water−Octadecyl-
Trichlorosilane Interface. Proc. Natl. Acad. Sci. U. S. A. 2006, 103 (49),
18401−18404.
(71) Mezger, M.; Sedlmeier, F.; Horinek, D.; Reichert, H.; Pontoni,
D.; Dosch, H. On the Origin of the Hydrophobic Water Gap: an X-ray
Reflectivity and MD Simulation Study. J. Am. Chem. Soc. 2010, 132,
6735−6741.
(72) Zhou, R. H.; Huang, X. H.; Margulis, C. J.; Berne, B. J.
Hydrophobic Collapse in Multidomain Protein Folding. Science 2004,
305, 1605.
(73) De Palma, R.; Peeters, S.; Van Bael, M. J.; Van den Rul, H.;
Bonroy, K.; Laureyn, W.; Mullens, J.; Borghs, G.; Maes, G. Silane
Ligand Exchange to Make Hydrophobic Superparamagnetic Nano-
particles Water-dispersible. Chem. Mater. 2007, 19, 1821−1831.
(74) Ho, T. A.; Papavassiliou, D. V.; Lee, L. L.; Striolo, A. Liquid
Water can Slip on a Hydrophilic Surface. Proc. Natl. Acad. Sci. U. S. A.
2011, 108, 16170−16175.
(75) Asay, D. B.; Kim, S. H. Effects of Adsorbed Water Layer
Structure on Adhesion Force of Silicon Oxide Nanoasperity Contact in
Humid Ambient. J. Chem. Phys. 2006, 124, 174712.
(76) Xiao, X. D.; Qian, L. M. Investigation of Humidity-Dependent
Capillary Force. Langmuir 2000, 16, 8153−8158.
(77) Chen, L.; Xiao, C.; Yu, B. J.; Kim, S. H.; Qian, L. M. What
Governs Friction of Silicon Oxide in Humid EnvironmentContact
area between Solids, Water Meniscus Around the Contact, or Water
Layer Structure? Langmuir 2017, 33 (38), 9673−9679.
(78) Chen, L.; He, H. T.; Wang, X. D.; Kim, S. H.; Qian, L. M.
Tribology of Si/SiO2 in Humid Air: Transition from Severe Chemical
Wear to Wearless Behavior at Nanoscale. Langmuir 2015, 31, 149−
156.
(79) Chen, L.; Qi, Y. Q.; Yu, B. J.; Qian, L. M. Sliding Speed-
Dependent Tribochemical Wear of Oxide-Free Silicon. Nanoscale Res.
Lett. 2017, 12 (1), 404.
(80) Alazizi, A.; Barthel, A. J.; Surdyka, N. D.; Luo, J. W.; Kim, S. H.
Vapors in the AmbientA Complication in Tribological Studies or an
Engineering Solution of Tribological Problems? Friction 2015, 3, 85−
114.
(81) Wang, X. D.; Kim, S. H.; Chen, C.; Chen, L.; He, H. T.; Qian, L.
M. Humidity dependence of tribochemical wear of monocrystalline
silicon. ACS Appl. Mater. Interfaces 2015, 7 (27), 14785−14792.

11391 DOI: 10.1021/acs.jpcc.8b01821


J. Phys. Chem. C 2018, 122, 11385−11391

You might also like