You are on page 1of 8

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/270455836

Nanorheology by atomic force microscopy

Article  in  Review of Scientific Instruments · November 2014


DOI: 10.1063/1.4903353

CITATIONS READS
15 105

4 authors, including:

Tai-De Li Hsiang-Chih Chiu


ASRC of CUNY National Taiwan Normal University
34 PUBLICATIONS   932 CITATIONS    36 PUBLICATIONS   546 CITATIONS   

SEE PROFILE SEE PROFILE

Elisa Riedo
New York University
128 PUBLICATIONS   4,646 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Patterning Spin-Wave ReconfIgurable Nanodevices for Logics and Computing (SWING) View project

Pressure induced Diamene phase from Graphene View project

All content following this page was uploaded by Tai-De Li on 22 April 2016.

The user has requested enhancement of the downloaded file.


Nanorheology by atomic force microscopy
Tai-De Li, Hsiang-Chih Chiu, Deborah Ortiz-Young, and Elisa Riedo

Citation: Review of Scientific Instruments 85, 123707 (2014); doi: 10.1063/1.4903353


View online: http://dx.doi.org/10.1063/1.4903353
View Table of Contents: http://scitation.aip.org/content/aip/journal/rsi/85/12?ver=pdfcov
Published by the AIP Publishing

Articles you may be interested in


Accurate formula for dissipative interaction in frequency modulation atomic force microscopy
Appl. Phys. Lett. 105, 233105 (2014); 10.1063/1.4903484

Finite element modeling of atomic force microscopy cantilever dynamics during video rate imaging
J. Appl. Phys. 109, 074309 (2011); 10.1063/1.3567933

Cross-talk compensation in atomic force microscopy


Rev. Sci. Instrum. 79, 103706 (2008); 10.1063/1.3002483

Thermal calibration of photodiode sensitivity for atomic force microscopy


Rev. Sci. Instrum. 77, 116110 (2006); 10.1063/1.2387891

Direct force balance method for atomic force microscopy lateral force calibration
Rev. Sci. Instrum. 77, 043903 (2006); 10.1063/1.2190210

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitationnew.aip.org/termsconditions. Downloaded to IP:
71.202.232.35 On: Tue, 23 Dec 2014 22:06:30
REVIEW OF SCIENTIFIC INSTRUMENTS 85, 123707 (2014)

Nanorheology by atomic force microscopy


Tai-De Li, Hsiang-Chih Chiu, Deborah Ortiz-Young, and Elisa Riedoa)
School of Physics, Georgia Institute of Technology, Atlanta, Georgia 30332, USA
(Received 14 June 2014; accepted 24 November 2014; published online 19 December 2014)
We present an Atomic Force Microscopy (AFM) based method to investigate the rheological prop-
erties of liquids confined within a nanosize gap formed by an AFM tip apex and a solid substrate.
In this method, a conventional AFM cantilever is sheared parallel to a substrate surface by means
of a lock-in amplifier while it is approaching and retracting from the substrate in liquid. The normal
solvation forces and lateral viscoelastic shear forces experienced by the AFM tip in liquid can be
simultaneously measured as a function of the tip-substrate distance with sub-nanometer vertical reso-
lution. A new calibration method is applied to compensate for the linear drift of the piezo transducer
and substrate system, leading to a more precise determination of the tip-substrate distance. By mon-
itoring the phase lag between the driving signal and the cantilever response in liquid, the frequency
dependent viscoelastic properties of the confined liquid can also be derived. Finally, we discuss the
results obtained with this technique from different liquid-solid interfaces. Namely, octamethylcy-
clotetrasiloxane and water on mica and highly oriented pyrolytic graphite. © 2014 AIP Publishing
LLC. [http://dx.doi.org/10.1063/1.4903353]

I. INTRODUCTION a separation smaller than 1 nm and topography imaging with


high lateral resolution is absolutely necessary. Atomic Force
The behavior of fluids confined within nanometers from
Microscopy (AFM), which permits to acquire high resolution
the solid interface has received a lot of attention in the past
topography images and perform high resolution force mea-
decades due to its importance in tribology, biology, geo-
surements between an AFM tip and a sample surface, provide
physics, and polymer science. Recently, it has been found that
an opportunity to investigate the properties of nano-confined
the behavior of nanoconfined liquids differs significantly from
water with unprecedented advantage. Indeed, a variety of
the bulk. Interesting phenomena occur at the liquid-solid in-
experimental techniques based on AFM have been utilized
terface where the classical continuum laws break down.1 Liq-
to probe the properties of liquid in the vicinity of different
uids confined between two solid surfaces can form ordered
substrates.9, 10 The solvation forces in OMCTS (Octamethyl-
layers, leading to structural oscillatory solvation forces2 and
cyclotetrasiloxane) and n-dodecanol near Highly Oriented
nonlinear viscoelastic dynamics.3 A powerful instrument to
Pyrolytic Graphite (HOPG) are observed using force-distance
probe liquids under nano-confinement is the surface force
spectroscopy with the contact mode AFM, but this oscillatory
apparatus (SFA).4, 5 Some of the disadvantages of SFA in-
behavior is not seen in water.11 The Frequency Modulation
clude the limited materials choice and jump-to-contact insta-
(FM) AFM based technique is employed to measure the sol-
bility, i.e., when the gradient of the force between two sur-
vation forces at the interface between OMCTs and HOPG.12
faces exceeds the spring constant of the apparatus, one surface
The FM-AFM equipped with a carbon nanotube probe is also
snaps into the other one losing force information. The Interfa-
used to study the oscillatory and hydration forces in the vicin-
cial Force Microscope (IFM),6 which uses oxide-terminated
ity of self assembled monolayers on gold substrate13 as well
tungsten tips with radii of about 10 μm,7 was designed to
as on biological membranes.14 Additionally, in FM-AFM,
avoid such instabilities with a self-balancing force feedback
since the oscillating frequency of the cantilever as well as
sensor. The mechanically stable Transverse Dynamic Force
the phase difference with respect to the cantilever driving sig-
Microscope (TDFM), also known as the Shear Force Mi-
nal are both recorded, this technique also enables the study
croscope, is able to probe the viscoelastic behavior of water
of the viscoelasticity of the nano-confined liquids.11, 12 How-
under nano-confinement, but lacks the resolution in vertical
ever, in FM-AFM, the interacting forces and the viscoelas-
separation smaller than 4 molecular layers owing to instru-
tic properties of nano-confined liquid are deduced indirectly
ment limitations.8 Nevertheless, how surface roughness or
from the frequency and phase shifts of the cantilever using
wettability of the confining surfaces influences the properties
complicated mathematical formulas with approximations. On
of the nano-confined water is not clearly known. Obviously
the other hand, optical interferometer based AFM method can
the aforementioned instruments, SFA, IFM, or TDFM, can-
simultaneously measure the normal and shear stiffness using
not image the surface topography with high lateral resolution
the vertical and lateral oscillations with high sensitivity, but
to acquire information about surface properties. Therefore, to
this method is subject to cross-talk of vertical and lateral sig-
comprehensively study the properties of nano-confined liq-
nals, cantilever-optical fiber alignment deviations and it is not
uids, an instrument that can perform force measurements for
compatible with commercial AFMs.15 Therefore, a method
that is easy to use and can simultaneously, directly, and un-
a) Email: elisa.riedo@physics.gatech.edu ambiguously measure the solvation forces and viscoelasticity

0034-6748/2014/85(12)/123707/6/$30.00 85, 123707-1 © 2014 AIP Publishing LLC

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitationnew.aip.org/termsconditions. Downloaded to IP:
71.202.232.35 On: Tue, 23 Dec 2014 22:06:30
123707-2 Li et al. Rev. Sci. Instrum. 85, 123707 (2014)

of nano-confined liquid with high resolution is apparently in


need.
In this paper, we present a new AFM-based technique for
simultaneous, quasi-statical normal and lateral force measure-
ments in nano-confined liquid as a function of tip-sample sep-
aration distance.2, 3 A lateral shearing signal is applied to the
AFM cantilever by a lock-in amplifier while it is approach-
ing the solid surface. The normal and lateral forces experi-
ence by the AFM tip in liquid can be recorded at the same
time. Thus this technique permits an in situ investigation of
the structural and dynamical properties for nano-confined flu-
ids and can be employed easily with most commercial AFM
without instrument modification. In addition, with a recently
developed correction procedure to account for the linear drift
of the piezo transducer, we are able to precisely determine the
tip-sample separation down to one molecule layer and inves- FIG. 1. Schematic of the experimental setup. With a lock-in amplifier, a lat-
tigate the properties of water confined between a gap smaller eral shearing signal is applied to the AFM tip as it approaches the solid sur-
than 1 nm. To demonstrate the capabilities of this technique, face. The normal and lateral force are recorded simultaneously. Both the tip
and the sample surface are immersed in a liquid environment.
in the following we will present the results obtained from var-
ious liquid-solid interfaces. More specifically, we will discuss
the normal solvation forces of OMCTS in the vicinity of mica dure is highly necessary because if there is a large angle be-
and HOPG, the normal solvation and lateral viscous forces tween the shearing direction and sample surface, there will be
of nano-confined water on mica, as well as the viscoelastic an artifact-like lateral force due to lateral tapping between the
properties of nano-confined water studied with our technique. tip and the surface, leading to unknown experimental errors.
The schematic of experimental setup is shown in
Figure 1. A silicon AFM tip approaches a solid surface (mica
II. EXPERIMENTAL or HOPG) while a lateral oscillation is applied to it by a lock-
in amplifier at the same time. A laser projected onto the back
A. Experimental setup
of the cantilever is reflected into a 4 quadrant photodiodes, as
In this work, we use a commercial AFM (PicoPlus 5500, shown in Figure 1. The normal force FN experienced by the
Agilent Technologies) equipped with a silicon AFM can- cantilever in liquid can be written as
tilever (NSC35, MikroMasch) with a typical tip radius, Rtip
= 40 ± 10 nm to study the physical properties of the nano- FN = δX · kN , (1)
confined water. The normal and torsional spring constants of 3
where kN = E4 wt is the normal spring constant and δX is the
the cantilever are kN ≈ 3–4.5 N/m and kT ≈ 50–120 N/m, re- L3
vertical bending of the cantilever due to tip-surface force. In
spectively. These force constants are determined by imaging
addition, E is the Young’s modulus of silicon16 while L, w,
the cantilevers with a Scanning Electron Microscope (JEOL
and t are the length, width, and thickness of the cantilever,
JSM-5910) to acquire their physical dimensions after each
respectively. The vertical cantilever bending δX = δVN · m,
measurement (More details in the text below). In addition, we
where δVN is the deflection signal measured by the photodi-
also perform inspections on the tip apex. We have found that
odes shown in Figure 1 and m is the optical lever sensitivity of
a rough tip surface with protuberances will prevent the obser-
the cantilever used. During the approach, a lateral oscillation,
vation of solvation forces in liquid.2 An AFM liquid cell is
X0 · sin(ωt), is applied to the piezo-scanner via a lock-in am-
used to contain the sample, liquids, and the AFM cantilever
plifier (Stanford Research Systems, SR830), where X0 and ω
inside an environmental chamber. This chamber is in ambient
are respectively the oscillation amplitude and frequency. The
conditions for experiments with water and is filled with ultra-
application of lock-in amplifier for data acquisition greatly
filtered nitrogen gas for OMCTS. The liquid cell is cleaned
enhances the signal to noise ratio.
by rinsing and sonication first in ethanol followed by Iso-
The lateral force FL experienced by the cantilever in liq-
propyl Alcohol and then blown dry with compressed nitro-
uid can be written as
gen gas. The mica and HOPG surfaces are prepared by the
tape-exfoliation method and then blown with compressed ni- FL = kT · hθ, (2)
trogen gas to eliminate any surface debris. The cleanliness of 2
the surfaces are examined by AFM topography images over where kT = 0.4 · kN · (h+L t )2 is the torsional spring constant
2
areas of 1 μm2 with contact mode AFM imaging. To insure of the cantilever, h represents the tip height, and θ is the
the tip is shearing in parallel to the sample surface, before torsional angle of the cantilever due to the presence of the lat-
each measurement, we tilt the sample stage and obtain the eral viscous force. The torsional angle θ is also monitored
AFM topography image of the surface until the difference in by the laser beam reflected from the backside of the cantilever
height across a surface distance of 1 μm is less than 1 nm. and recorded by the photodiodes as VL as shown in Figure 1.
This height difference corresponds to a tilted angle of less For a rectangular cantilever used in our AFM system, the tor-
than 0.06◦ between the tip and sample surface. This proce- sional angle θ is proportional to VL through the following
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitationnew.aip.org/termsconditions. Downloaded to IP:
71.202.232.35 On: Tue, 23 Dec 2014 22:06:30
123707-3 Li et al. Rev. Sci. Instrum. 85, 123707 (2014)

relationship:17 and retracting scanner deformations, respectively. The initial


slopes of the approaching and retracting curves in the contact
θ (rad) = VL (V)/(7.5 × 103 ) (rad/V). (3)
regime, with linear drift present, are Sa 0 and Sr 0 as shown in
Combining Eqs. (2) and (3), we arrive at the lateral force ex- Figure 2(a) For the approaching curve, we have Sa 0 = ba 0 /Za 0 ,
perienced by the cantilever at a given tip-surface separation, where ba 0 and Za 0 are the cantilever bending and piezo defor-
mation when there is linear drift. After applying the new unit,
  the absolute value of the slope of the force curve at the contact
L2 h · V
FL = 0.4 · kN ·  2 · . (4) regime should be 1, that is
h + 2t 7.5 × 103
Ub · ba0 U
= b · Sa0 = 1, (5)
This lateral force, FL (z) = FL sin[ωt+ φ(z)], experienced
0
Ua · Za0 Ua
by the AFM tip in liquid as a function of tip-solid separation
such that
distance z, is then measured by the lock-in amplifier, where
φ(z) is the phase difference between the applied lateral driv- Ua
Ub = . (6)
ing signal and the detected lateral force signal. At φ = 0, the Sa0
tip is in hard contact with the solid surface and the lateral Similarly, for the retracting curve, we will have Ub = Ur /Sr 0.
oscillation amplitude is small enough to guarantee a purely By combining Eqs. (5) and (6), we arrive at
elastic contact deformation without slippage.18
Sr0 · Ua
Ur = Ub · Sr0 = . (7)
Sa0
B. Drift analysis
If we assume that the drift velocity, Vdrift , is in the same direc-
Since the tip approaches the solid surface at a rate of only tion as tip retraction, then the modified approaching (Z[Ua ])
0.2 nm/s (in water) or 0.4 nm/s (in OMCTS), the linear drift and retracting (Z[Ur ]) distances are
of the piezo scanner becomes important and cannot be ne-
Z[Ua ] = (Vscanner + Vdrif t )T [nm], (8)
glected in the data analysis. When no linear drift is present,
the approaching and retracting FN vs. d curves should over-
lap and have a slope of “−1” in the range where the tip is in Z[Ur ] = (Vscanner − Vdrif t )T [nm], (9)
contact with the surface, due to the fact that at hard contact where Vscanner and T are the scanner movement velocity and
the piezo movement and cantilever bending are equivalent. the time to approach or retract the tip, respectively. From
However, because of the linear drift, the contact lines will not Eqs. (5)–(8), Ua , Ub , and Ur can be determined as
overlap as shown in Figure 2(a) This can be understood as pi-
2Sa0
loting a boat downstream and upstream a river to measure the Ua = [nm], (10)
velocity of the current. The measured velocity of the current Sa0 + Sr0
going downstream or upstream will depend on the velocity of 2Sr0
the boat. When no linear drift is present, the total deforma- Ur = [nm], (11)
Sa0+ Sr0
tion for the piezo-scanner, Z as shown in Figure 2, should be
identical for both tip approaching and retracting movements. 2
Ub = [nm]. (12)
The total displacement showed in the raw force curves by the + Sr0
Sa0
AFM for directions are both “Z nm.” However, owing to the A simple home written program has been developed to au-
drift, this “Z nm” for the approaching and retracting curve is tomate this procedure. A typical example of a normal force-
actually different. This drift can be calibrated by introducing distance curve is shown in Figure 3.
new length units (while the old unit is “1 nm”) for the scanner
deformation and cantilever bending such that after correction C. Viscoelasticity of nano-confined water
the slopes of both contact lines shall be “−1.” The new units
are Ub , Ua , and Ur for the cantilever bending, approaching The obtained lateral viscous forces can also be used to
understand the viscoelastic properties of water. When water
is confined between two smooth plates with surface area A,
separated by a distance d, and with one plate sliding in paral-
lel at a velocity Vshear with respect to the other, the viscosity
η of water can be evaluated by η = FL · d/(Vshear · A), where
FL is the lateral force required to maintain one plate moving
at Vshear with respect to the stationary one.3 If a sinusoidal
strain, γ = γ 0 · sin(ωt), is applied to one of the plates while
the other is stationary, the resulting shear stress between two
plates will be σ = σ 0 sin(ω · t + φ). Since the strain and stress
FIG. 2. (a) Typical force distance curves when a constant linear drift is amplitudes can be written as γ 0 = X0 /d and σ 0 = FL /A, re-
present during measurement. The slopes of the approaching and retracting spectively, we arrive at the following relationship:
portions at the hard contact regimes are not the same. This drift can be ac-
FL X
counted for using Eqs. (5)–(12). (b) Schematic shows water molecules con- = |G∗ | 0 , (13)
fined between an AFM tip and a solid surface and form layering structures. A d
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitationnew.aip.org/termsconditions. Downloaded to IP:
71.202.232.35 On: Tue, 23 Dec 2014 22:06:30
123707-4 Li et al. Rev. Sci. Instrum. 85, 123707 (2014)

FIG. 3. (a) Typical cantilever bending – Scanner movement curves with constant drift in water. Red is the approaching curve and blue is the retracting curve.
(b) After the correction by Eqs. (10)–(12), the approaching curve overlaps with the retracting curve and the slope of the contact regime is −1. (c) The cantilever
bending – Distance curve before correction. The contact is not steady and drifts linearly. The oscillation period of the normal force does not match water
molecular size, 0.25 nm. (d) The cantilever bending – Distance curve after correction. The curve shows a steady contact and the period of the oscillatory normal
force is now with 0.25 ± 0.05 nm, water molecule size.

where G∗ = G + iG is the viscoelastic modulus19 con- III. RESULTS AND DISCUSSION
taining both the storage (elastic) modulus (G ) and the loss
Figure 4 summarizes the normal forces FN obtained for
(viscous) modulus (G ) of the material. Furthermore, with
OMCTS at the interface of mica and HOPG. Figures 4(a)
Eq. (13), G∗ can be written as
and 4(b) display data acquired without tip shearing dur-
ing approaching the surface, while Figures 4(c) and 4(d) il-
FL d FL d
G = cos φ and G = sin φ. (14) lustrate data obtained with tip shearing with the same sur-
AX0 AX0 faces. Each figure shows the overlap of more than 8 different

FIG. 4. (a) and (c) The approaching normal force-distance curves acquired in OMCTS on Mica with and without tip shearing, respectively. The vertical dashed
lines indicate the layering of water molecules confined between the AFM tip and the surface. These data are obtained with shearing amplitude of 0.4 nm with
different frequencies (34.9 kHz, 107.4 kHz, 212.3 kHz, and 884.7 kHz). (b) and (d) The approaching normal force-distance curves acquired in OMCTS on
HOPG with and without tip shearing, respectively. These data are recorded with a shearing amplitude of 1.18 nm with various frequencies (58.1 kHz, 105.2 kHz,
and 397.7 kHz). Excellent overlapping of multiple force-distance curves in all figures demonstrates the reproducibility of our measurements. The insets show
the same force curves with larger vertical axes to show that all force curves diverge as d approaches zero.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitationnew.aip.org/termsconditions. Downloaded to IP:
71.202.232.35 On: Tue, 23 Dec 2014 22:06:30
123707-5 Li et al. Rev. Sci. Instrum. 85, 123707 (2014)

measurements demonstrating data reproducibility. When the a frequency ω = 995 Hz, respectively. The dashed vertical
tip is not laterally sheared with respect to the surface, the lines highlight the molecular order in the repulsive oscilla-
structural layering of OMCTS molecule is clearly observed tory solvation force. The average diameter of water molecule
in the FN vs. d data beginning approximately at d ≈ 4 nm is found to be 0.27 nm, consistent with results reported
from both surfaces. In Figures 4(a) and 4(b), the vertical in the literature.4, 7, 13–15, 23–25 Furthermore, the lateral force
dashed lines indicate the force maxima which correspond to FL increases drastically when the tip-sample distance de-
the first through fourth molecular layers. These oscillatory creases, especially for d < 1 nm. Such increase of FL of liq-
forces have a period of 0.85–1 nm which is consistent with uid at small d has been demonstrated by other experimen-
the diameter of the OMCTS molecules.20 The oscillations in tal techniques.7, 23, 26, 27 The increase of FL is related to the
the force curve occur when the tip-sample separation distance increasing viscosity of water, which can be estimated with
decreases and squeezes out a molecular layer of OMCTS, Eq. (13), assuming water is confined between two smooth
as already demonstrated by SFA5 and several AFM-based plates. In the case of a spherical AFM tip with radius RTip
measurements.11, 12, 21, 22 Additionally, the oscillatory behav- ≈ 40 nm moving in parallel to a smooth surface with a sep-
ior is more clearly shown on Mica than on HOPG probability aration distance d < 1.5 nm and a shearing velocity Vshear , a
due to different surface wettability with OMCTS. The first more rigorous treatment2 leads to
layer of OMCTS molecules on the surface might not stick
well on hydrophobic HOPG, leading to difficulties in forming FL
η=    , (15)
layered structures of OMCTS molecules. 2π Vshear (R + d) ln 1 + h
d
− h
Figures 4(c) and 4(d) show the approaching normal force
for OMCTS molecules at the mica and HOPG interface with for water confined between 0 < z < d + h (See Figure 2(b)).
the tip is laterally sheared with respect to the sample surface For h ∼ 0.25 nm (size of water molecule) and h = 0.5 nm,
(corresponding FL not shown), respectively. The oscillatory the viscosity of the confined water can be estimated to be ∼3
behavior of FN vs. d is clearly shown again for OMCTS on × 102 P, almost 4 order in magnitude larger than that of the
Mica but not on HOPG, due to different surface wettability. bulk water, which is about 10−2 P. This unusual behavior sug-
The force profiles of OMCTS confined between a silicon tip gests the possibility of lateral order in water molecules, which
and HOPG are attractive when d < 2 nm, similar to previ- is usually observed only for spherical non-polar molecules.28
ous results obtained with other AFM based techniques.12, 21, 22 Recent AFM based measurements have also observed later-
The insets of Figures 4(c) and 4(d) display the same data with ally organized water on Mica and deduced a molecule di-
larger vertical scale showing that all forces are well converged ameter of 0.2–0.3 nm.29, 30 Next, by using Eq. (14), the vis-
when d approaches 0. coelastic response of the nano-confined water can be derived
The comparison between the normal forces in presence directly from the lateral force, FL , and phase change, , as
and absence of shearing indicates that shearing the tip mildly shown in Figure 5(b). The obtained elastic component, G ,
disrupts the layering process, giving rise to less pronounced and viscous G component, as a function of tip-sample dis-
oscillations. This effect is particular evident on HOPG. tance d is displayed in Figures 5(c) and 5(d), respectively. At
Figures 5(a) and 5(b) shows the normal and lateral forces this experimental conditions, X0 = 0.4 nm and ω = 955 Hz,
simultaneously captured as a function of the tip-sample sep- the elastic component is larger than the corresponding viscous
aration in deionized ultra filtered (DIUF) water on Mica component for d < 1 nm, indicating the glassy state of nano-
with the technique discussed in Sec. II. The data were ac- confined water, consistent with previous measurements with
quired with a lateral shearing amplitude X0 = 0.4 nm and TDFM8 and other AFM based techniques.30, 31

FIG. 5. (a) and (b) The normal and lateral force curves obtained simultaneously as a function of tip-sample distance in water on a Mica surface, respectively.
All curves displayed here are acquired with a shearing amplitude of 0.4 nm and a frequency of 995 Hz. (c) and (d) Elastic and viscous moduli G and G as a
function of tip-sample distance determined from one of the lateral force curves (open square) shown in (b).

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitationnew.aip.org/termsconditions. Downloaded to IP:
71.202.232.35 On: Tue, 23 Dec 2014 22:06:30
123707-6 Li et al. Rev. Sci. Instrum. 85, 123707 (2014)

IV. CONCLUSION 8 M. Antognozzi, A. D. L. Humphris, and M. J. Miles, “Observation of


molecular layering in a confined water film and study of the layers vis-
In the foregoing we have presented a technique with coelastic properties,” Appl. Phys. Lett. 78, 300–302 (2001).
9 M. Abdelhamid and B. Bharat “Nanorheology and boundary slip in con-
AFM which permits simultaneous measurements of the nor-
fined liquids using atomic force microscopy,” J. Phys.: Condens. Matter
mal solvation force and lateral viscous forces of nano-
20, 315201 (2008).
confined liquid as well as its viscoelastic properties. With 10 E. Bonaccurso, M. Kappl, and H.-J. Butt, “Thin liquid films studied by
proper cantilever calibrations and the successful correction of atomic force microscopy,” Curr. Opin. Colloid Interface Sci. 13, 107–119
the linear drift of the piezo scanner, we are able to capture the (2008).
11 S. J. O’Shea, M. E. Welland, and T. Rayment, “Solvation forces near a
structural and viscoelastic properties of liquid at the solid in- graphite surface measured with an atomic force microscope,” Appl. Phys.
terface with a confining gap down to ∼0.3 nm, which are dif- Lett. 60, 2356–2358 (1992).
ficult to achieve with other techniques summarized in Sec. I. 12 S. J. O’Shea and M. E. Welland, “Atomic force microscopy at solid−liquid

We demonstrate the capability of this experimental technique interfaces,” Langmuir 14, 4186–4197 (1998).
13 S. P. Jarvis, T. Uchihashi, T. Ishida, H. Tokumoto, and Y. Nakayama, “Local
by studying the oscillatory normal force of OMCTS on Mica solvation shell measurement in water using a carbon nanotube probe,” J.
surfaces, showing a oscillation period of 0.85–1 nm, which is Phys. Chem. B 104, 6091–6094 (2000).
consistent with diameter, 0.7–0.85 nm, of OMCTS.20 Simi- 14 M. J. Higgins et al., “Structured water layers adjacent to biological mem-

lar experiments with DIUF water on mica surfaces also show branes,” Biophys. J. 91, 2532–2542 (2006).
15 S. Patil et al., “A highly sensitive atomic force microscope for linear mea-
the layered structure of nanoconfined water, finding an oscil- surements of molecular forces in liquids,” Rev. Sci. Instrum. 76, 103705
lation period of ∼0.27 nm, consistent with the diameter of a (2005).
water molecule and with previously reported values.4, 15, 24, 25 16 E. Meyer, H. J. Hug, and R. Bennewitz, Scanning Probe Microscopy: The

Furthermore, the viscoelastic properties, namely viscous and lab on a Tip (Springer, 2004).
17 T.-D. Li, Ph. D. thesis, Georgia Institute of Technology, 2008.
elastic modulus, as well as relaxation time of confined liq- 18 R. W. Carpick, D. F. Ogletree, and M. Salmeron, “Lateral stiffness: A new
uids can be directly obtained from the acquired lateral force nanomechanical measurement for the determination of shear strengths with
data. This novel technique provides an approach to compre- friction force microscopy,” Appl. Phys. Lett. 70, 1548–1550 (1997).
19 J. D. Ferry, Viscoelastic Properties of Polymers (Wiley, 1980).
hensively study and understand the physical properties of 20 D. W. Scott “Equilibria between linear and cyclic polymers in
nano-confined liquids.32 methylpolysiloxanes,” J. Am. Chem. Soc. 68, 2294–2298 (1946).
21 W. Han and S. M. Lindsay, “Probing molecular ordering at a liquid-solid

interface with a magnetically oscillated atomic force microscope,” Appl.


ACKNOWLEDGMENTS Phys. Lett. 72, 1656–1658 (1998).
22 R. Lim, S. F. Y. Li, and S. J. O’Shea, “Solvation forces using sample-

T.-D.L., H.-C.C., D.O.-Y., and E.R. acknowledge the modulation atomic force microscopy,” Langmuir 18, 6116–6124 (2002).
23 R. C. Major, J. E. Houston, M. J. McGrath, J. I. Siepmann, and X. Y.
financial support of the Office of Basic Energy Sciences
Zhu, “Viscous water meniscus under nanoconfinement,” Phys. Rev. Lett.
DOE (DE-FG02-06ER46293). E.R. acknowledges the Na- 96, 177803 (2006).
tional Science Foundation NSF (CMMI-1100290) for partial 24 J. R. Bonander and B. I. Kim, “Cantilever based optical interfacial force

support. We thank Suenne Kim for extensive discussions. microscope,” Appl. Phys. Lett. 92, 103124 (2008).
25 S. Jeffery et al., “Direct measurement of molecular stiffness and damping

in confined water layers,” Phys. Rev. B 70, 054114 (2004).


1 J. 26 Y. Zhu and S. Granick, “Viscosity of interfacial water,” Phys. Rev. Lett. 87,
N. Israelachvili, Intermolecular and Surface Forces (Academic Press,
1992). 096104 (2001).
2 T. D. Li, J. P. Gao, R. Szoszkiewicz, U. Landman, and E. Riedo, “Structured 27 U. Raviv, P. Laurat, and J. Klein, “Fluidity of water confined to subnanome-

and viscous water in subnanometer gaps,” Phys. Rev. B 75, 115415 (2007). tre films,” Nature (London) 413, 51 (2001).
3 T. D. Li and E. Riedo, “Nonlinear viscoelastic dynamics of nanoconfined 28 J. Gao, W. D. Luedtke, and U. Landman, “Origins of solvation forces in

wetting liquids,” Phys. Rev. Lett. 100, 106102 (2008). confined films,” J. Phys. Chem. B 101, 4013–4023 (1997).
4 J. N. Israelachvili and R.M.Pashley, “Molecular layering of water at sur- 29 K. Kimura et al., “Visualizing water molecule distribution by atomic force

faces and origin of repulsive hydration forces,” Nature (London) 306, 249 microscopy,” J. Chem. Phys. 132, 194705 (2010).
30 T. Fukuma, Y. Ueda, S. Yoshioka, and H. Asakawa, “Atomic-scale distri-
(1983).
5 R. G. Horn and J. N. Israelachvili, “Direct measurement of structural forces bution of water molecules at the mica-water interface visualized by three-
between two surfaces in a nonpolar liquid,” J. Chem. Phys. 75, 1400–1411 dimensional scanning force microscopy,” Phys. Rev. Lett. 104, 016101
(1981). (2010).
6 S. A. Joyce and J. E. Houston, “A new force sensor incorporating force- 31 S. H. Khan, G. Matei, S. Patil, and P. M. Hoffmann, “Dynamic solidifica-

feedback control for interfacial force microscopy,” Rev. Sci. Instrum. 62, tion in nanoconfined water films,” Phys. Rev. Lett. 105, 106101 (2010).
32 D. Ortiz-Young, H. C. Chiu, S. Kim, K. Voitchovsky, and E. Riedo, “The
710–715 (1991).
7 M. P. Goertz, J. E. Houston, and X. Y. Zhu, “Hydrophilicity and the viscos- interplay between apparent viscosity and wettability in nanoconfined wa-
ity of interfacial water,” Langmuir 23, 5491–5497 (2007). ter,” Nat. Commun. 4, 2482 (2013).

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitationnew.aip.org/termsconditions. Downloaded to IP:
71.202.232.35 On: Tue, 23 Dec 2014 22:06:30
View publication stats

You might also like