Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Biophysical Journal Volume 106 March 2014 10331043 1033

Local Viscoelastic Properties of Live Cells Investigated Using Dynamic and


Quasi-Static Atomic Force Microscopy Methods

Alexander Cartagena and Arvind Raman*


School of Mechanical Engineering and Birck Nanotechnology Center, Purdue University, West Lafayette, Indiana

ABSTRACT The measurement of viscoelasticity of cells in physiological environments with high spatio-temporal resolution is a
key goal in cell mechanobiology. Traditionally only the elastic properties have been measured from quasi-static force-distance
curves using the atomic force microscope (AFM). Recently, dynamic AFM-based methods have been proposed to map the local
in vitro viscoelastic properties of living cells with nanoscale resolution. However, the differences in viscoelastic properties esti-
mated from such dynamic and traditional quasi-static techniques are poorly understood. In this work we quantitatively recon-
struct the local force and dissipation gradients (viscoelasticity) on live fibroblast cells in buffer solutions using Lorentz force
excited cantilevers and present a careful comparison between mechanical properties (local stiffness and damping) extracted
using dynamic and quasi-static force spectroscopy methods. The results highlight the dependence of measured viscoelastic
properties on both the frequency at which the chosen technique operates as well as the interactions with subcellular components
beyond certain indentation depth, both of which are responsible for differences between the viscoelasticity property maps
acquired using the dynamic AFM method against the quasi-static measurements.

INTRODUCTION
Many recent efforts in cell mechanobiology (14) aim to model. Another conventional way to evaluate the visco-
quantitatively measure the mechanical properties of living elastic properties of live cells is through the acquisition of
cells and relate them to cell structure and function. Cells force-distance curves by applying a rectangular load-relax-
engage in complex processes changing the viscoelastic ation (stress-relaxation) pattern (21,39). Several groups
response (59) of the cell membrane, cytoskeleton, and have estimated the local mechanical properties using these
cytoplasm. These changes are often heterogeneous in spatial methods on live bacteria (4044), and eukaryotic cells
extent within the cell, and change with time. The measure- (15,21,34,39,4450) in their native liquid environment;
ment of the progressive spatio-temporal variations in visco- however, the methods are low-speed and low-resolution,
elastic properties within living cells in their native limiting their potential for high throughput biomechanical
physiological liquid environments could shed important assay of cells.
insight into cellular processes such as morphogenesis To address some of the speed and spatial resolution chal-
(10,11), mechanotransduction (1214), migration/locomo- lenges in conventional AFM techniques for mapping cell
tion (1518), metastasis (1925), apoptosis (26,27), aging mechanical properties, a method called multi-harmonic
(27), focal adhesion (18,2832), disease progression (33), AFM was recently reported (4) that employed amplitude-
and drug-cell interactions (3436). modulation atomic force microscopy (AM-AFM), a tech-
The atomic-force microscope (AFM) is unique among nique allowing achievement of sub-10-nm resolution
other cell mechanical measurement techniques (1,6,37,38) high-speed mapping of local nanomechanical properties of
in its ability to measure the local force and dissipative gra- live cells in physiological conditions. However, the effective
dients as well as map them across the cell surface with sub- local properties (loss and storage modulus) over the nuclear
10-nm resolution. As a result, the AFM allows researchers region were found to be generally 35 times larger
to develop quantitative methods to map the local mechanical compared to values acquired from quasi-static force-dis-
properties of living cells. The standard force-volume tance curves. Understanding the basic reasons behind the
method is the most widely used imaging method to extract, differences in measured properties using the two methods
simultaneously, the topography and mechanical properties is key to developing quantitative tools to measure cell visco-
of a live cell. It is based on the acquisition of slow-speed, elasticity using dynamic AFM.
quasi-static force-distance curve measurements on a grid To address this fundamental issue, in this article we study
of points defined by the user. The mechanical properties the local force gradient and damping on live rat fibroblast
in force-volume maps are extracted offline for each recorded cells in buffer solution, measured as a function of indenta-
force curve by fitting to an analytical tip-sample contact tion using a soft AFM microcantilever excited by Lorentz
force near its natural frequency (~78 kHz). We find that,
at small indentations, the dynamically measured force
Submitted November 12, 2013, and accepted for publication December 31,
2013. gradient is ~3 times that of the statically measured one,
*Correspondence: raman@purdue.edu showing the frequency-dependence. However, the ratio of
Editor: Daniel Muller. the dynamic to static force gradients begins to increase
2014 by the Biophysical Society
0006-3495/14/03/1033/11 $2.00 http://dx.doi.org/10.1016/j.bpj.2013.12.037
1034 Cartagena and Raman

when the AFM tip indents the cell sufficiently to interact is prescribed. In this work, we do not focus on extracting constitutive
with the nuclear region of the cell and its associated cyto- material properties like storage or loss modulus from the amplitude and
phase-shift data of the oscillating cantilever. This would require us to use
skeletal structure, nuclear envelope, and additional subcel- well-validated elastic contact mechanics models derived from the literature
lular elements. On the peripheral region, the indentation is such as that of Hertz (4,15,20,34,46), Sneddon (51), or bottom effect cone
much smaller and there is less difference between the correction (BECC) (from Gavara and Chadwick (52)). However, these
measured static and dynamic force gradients. On the other models have only been validated on live cells using quasi-static indentation
hand, the effective contact viscosity is less near the nuclear curves, and it remains an open question as to which models to choose for
dynamic AFM in which the cantilever oscillates in the range of 5
region than the peripheral regions. 10 kHz. Rheological models for live eukaryotic cells (51) generally do
We conclude that, when using AM-AFM with directly not take into account the heterogeneity of nuclear and peripheral regions.
excited probes, the mapped repulsive force gradient appears Although three-element models have been used for cell viscoelasticity mea-
larger on the nuclear region compared to the peripheral re- surements (9,21,53,54), we use here the two-element Kelvin-Voigt model
gion for two different reasons: for the following reasons:
1. The dynamic AFM method we discuss here uses two observables, the
1. Because larger forces need to be applied over the nuclear cantilever harmonic amplitude and phase to determine the sample visco-
region to achieve the same oscillation amplitude as on the elasticity. As a consequence, only two unknown viscoelastic parameters
can be determined at each Z location (during approach curves) or at any
peripheral regions, this requires the AFM tip to be pressed
pixel (during a scan). 2. Thus, the single frequency method presented
sufficiently into the cell membrane to interact primarily here can only treat a viscoelastic model with two unknown parameters
with the nuclear complex and cytoskeleton in the nuclear such as the Kelvin-Voigt element (spring and dashpot in parallel) and the
region. Maxwell element (spring and dashpot in series). 3. Of these two, the
2. The inherent viscoelasticity of the cell leads to a different Kelvin-Voigt model is physically relevant because the oscillation time of
the experiments (8 kHz 125 ms) is faster than the live fibroblast cell relax-
mechanical response at the higher frequency of oscilla-
ation time (~0.1 s) (55).
tion in the dynamic AFM method compared to the In light of this, we convert the amplitude and phase-shift data of the oscil-
quasi-static method. dynamic
lating cantilever into local dynamic repulsive-force gradient ksample and
dynamic
damping csample values. Together these are the parameters of an equiva-
The findings suggest that, at least on live fibroblast cells, lent Kelvin-Voigt element and can be regarded as the local viscoelastic
maps of material properties that were created using both the properties of the cell.
force-volume method and the dynamic AFM method are The theory behind this is described below. Assuming that the Lorentz
not directly comparable because of the indentation- and force cantilever driving frequency is near the first resonance frequency,
the single degree-of-freedom equation of motion governing the tip motion
frequency-dependence.
q(t) when it oscillates far from the sample surface is:
See the Supporting Material for additional details
on the text. q 1 Fmag sinudr t
q_ q ; (1)
u2far ufar Qfar kcant
MATERIALS AND METHODS
where ufar is the cantilever frequency (rad/s), Qfar is the quality factor far
Sample preparation from the sample surface, udr is the cantilever drive frequency (rad/s),
Rat fibroblast cells (ATCC CRL-1213; American Type Culture Collection, Fmag is the magnitude of the magnetic excitation force, kcant is the calibrated
Manassas, VA) were cultured in Dulbeccos modified Eagles medium cantilever spring constant, and q_ is the tip velocity. Solving the steady-state
(Invitrogen, Carlsbad, CA) containing low D-glucose (1000 mg l1), 10% vibration response q(t) A1 sin(udrt f1), it can be easily shown that when
Fetal Bovine Serum (Invitrogen), 1% Penicillin-Streptomycin (Invitrogen), udr is tuned to the peak amplitude of the resonance curve far from the sur-
and 0.1% Amphotericin B (Sigma-Aldrich, St. Louis, MO). The fibroblast face, the following expressions hold:
cells were grown to ~75% confluence in 75 cm2 flasks. The medium was q
removed every three days, and cells were subcultured every seven days at a ra- udr=u 1  1=2Q2 ;
far far
tio of 1:8. For the experiments, the cells grown on plastic flasks were trypsi-
nized with 0.5% Trypsin/EDTA solution (Invitrogen) and the cell Fmag Qfar
suspension was deposited on polystyrene-plasma-treated 60  15 mm petri A1far q (2)
kcant 1  1= 2 ;
dishes (BD Falcon, Franklin Lakes, NJ) precoated with 0.1% gelatin in 4Qfar
water (STEMCELL Technologies, Vancouver, British Columbia, Canada). p
The cells were planted/grown on the dish 12 days before experiments and tan f1far 4Q2far  2 :
kept in an incubator at 37 C in a 5% CO2 atmosphere to ensure complete
spreading.
Here, A1far is the oscillation amplitude and f1far is the phase lag far from the
sample surface. Note that by Eq. 2, when tuning the cantilever far from the
Theory of dynamic AFM material property sample, the phase lag at the frequency of peak amplitude with low Q, say 2,
mapping and spectroscopy in amplitude is not to be set to 90 ; instead, it should be set to f1far 75 . Thus, when the
drive is tuned at the peak amplitude of resonance far from the sample we
modulation AFM on live cells must have
It has been shown in a previous work (4) that harmonic signals (multi-har-
kcant A1far q
monic observables) of cantilever vibration can be combined to extract local Fmag 1  1=4Q2 :
nanomechanical properties once an analytical tip-sample interaction model Qfar far

Biophysical Journal 106(5) 10331043


Viscoelasticity of Live Cells Investigated Using AFM 1035

For soft cantilevers in liquids, their resonance frequencies (measured in damping (viscoelastic properties) at that particular indentation value. More-
dynamic 0 00
rad/s) and Q factors near the sample (i.e., unear and Qnear) are different from over, ksample and cdynamic
sample used here are equivalent to K and K of Mahaffy
the values measured far from the surface (i.e., ufar and Qfar) because of the et al. (58). Substituting Eq. 6 into Eq. 7, substituting the resulting expres-
hydrodynamic coupling between the cantilever and the sample (56). As a sion for Fts into Eq. 4, and taking into account Eqs. 2, 3, and 5, one can
result, the oscillation amplitude and the phase lag near the surface (i.e., match the harmonic terms of each side of the equation of motion (Eq. 4)
A1near and f1near) are different from their values far from the sample (i.e., to yield the following:
A1far and f1far). As shown by Raman et al. (4), the quantities near and far
!
from the sample are related by
kcant A1far kcant A1far
!2 q dynamic
ksample cosf1  cosf1near
A1far 1  1=4Q2 Qfar A1 A1near Qfar
udr far
1 cosf1near ; q
unear A1near Qfar  1  1=4Q2 ;
q (3) far

A1far 1  1=4Q2 !
udr far kcant A1far kcant A1far
sinf1near : cdynamic
sample sinf1  sinf1near
Qnear unear A1near Qfar Qfar A1 udr A1near Qfar udr
q
It is important to note that the hydrodynamic coupling to the surface is  1  1=4Q2 : (8)
far
different on live fibroblast cells that are tall (~24 mm in height) and soft
(11000 kPa), as compared to when the cantilever is located at the same
height above a rigid surface. For example, when the cantilever is excited These equations allow the conversion of the harmonic observables (A0, A1,
at resonance far from the sample so that f1far 74 and with a free and f1) into the local viscoelastic properties, i.e., the dynamic repulsive
dynamic
amplitude A1far ~ 9 nm, we find that when brought close to the sample, force gradient ksample and damping cdynamic
sample , while taking into account
just before the tip-sample interactions begin, A1near /f1near values on the near-surface hydrodynamic corrections.
the gelatin-coated dish and live cell are typically 2.5 nm/95 and
It is important to reiterate the different assumptions under which the
3.5 nm/89 , respectively. This clearly demonstrates that the squeeze-film
above equations are accurate:
hydrodynamic effect is much stronger on the substrate than on the live
cell. This hydrodynamic correction is essential to account for while Assumption 1
measuring the viscoelastic properties of live cells. This correction is rela-
tively straightforward to perform for the case when the cantilever is The cantilever is driven directly, not acoustically or via sample excitation,
excited at very low frequencies (57), where the cantilever inertia is not so that the cantilever has a well-defined transfer function (59,60) with the
important. Here we have presented the correction for the case when the Lorentz force being the only source of excitation without interference
cantilever is excited near its resonance frequency, and accounts for both from fluid-borne excitation (61) that arises when the either the cantilever
the added damping and the added mass of the fluid in the near vicinity or the sample are excited using piezoelectric transducers.
of the sample surface.
When the cantilever is brought closer to the sample so that it interacts Assumption 2
with the soft cell surface, the equation of motion becomes:
The tip is in continuous contact with the sample and that the oscillation
amplitude is small relative to the net indentation of the tip into the sample.

q 1 _
Fmag sinudr t Fts Z q; q
q q_ ;
2
unear unear Qnear kcant Assumption 3
(4) The eigenmode of the cantilever oscillating away from the sample remains
unchanged when compared to being pressed against the sample. This
where Fts is the tip-sample interaction force, and Z is the distance between
assumption in fact enables us to state that the kcant does not change when
the tip and sample assuming that the cantilever is rigid or unbent. As
coupled to the sample. This assumption is known to be correct only when
observed in Raman et al. (4) and also demonstrated in Results and Discus-
the contact stiffness/repulsive gradient is much smaller than the cantilever
sion, the cantilever motion while interacting with live cells is dominated by
stiffness (nominally 0.09 N/m) (62).
the zeroth and first harmonics, leading to the following form of the tip
motion: Assumption 4
qt A0 A1 sinudr t  f1 : (5) We need to recognize, as stated before, that the hydrodynamic correction is
To convert zeroth and first harmonics observables into local viscoelastic different on the live cell compared to that on the gelatin-coated glass
properties, we first annotate the tip indentation d(t) into the sample as surface.
Typically, A1near and f1near on a gelatin-coated dish are ~1.54 nm/95
dt Z q Z  A0  A1 sinudr t  f1 : (6) 98 , whereas on live cells the values are ~25 nm/8589 on nucleus and
~1.83.5 nm/9295 on periphery. However, Eq. 8 uses a single value of
As described in Results and Discussion and in Raman et al. (4), the tip A1near and f1near. Because we are more interested in the properties of the
oscillation amplitude A1 is much smaller compared to the net average cell than the gelatin surface, we choose to process all the data using
indentation d0 (Z A0) while imaging live cells in AM-AFM. Next, A1near and f1near on top of the cell. This implies that the predicted repulsive
we describe the interaction forces as a Taylor series expansion in (d d0) gradients on a gelatin-coated dish are systematically larger than the actual
and discard higher-order terms, value due to the hydrodynamic effect. However, we expect the correct
values to be mapped on the live cell. All computations and data processing
Fts Fts d0 ksample
dynamic
d  d0 cdynamic _ (7)
sample d; were performed using the software MATLAB (The MathWorks, Natick,
MA). For each set of multi-harmonic observable (A0, A1, and f1) curves
dynamic
where ksample and cdynamic
sample represent the parameters of a Kelvin-Voigt
dynamic
the viscoelastic properties maps (ksample and cdynamic
sample ) were extracted by us-
element, specifically the conservative local force gradient (stiffness) and ing Eq. 8.

Biophysical Journal 106(5) 10331043


1036 Cartagena and Raman

Quasi-static force spectroscopy


In quasi-static force spectroscopy, an unexcited AFM probe approaches and
is retracted from the sample surface and the deflection of the microcantile-
ver tip q is recorded as a function of Z-piezo extension. The resulting
force-distance (F-Z) curve is converted to a force-indentation (F-d) curve
by d Z q. The slope of the F-d curve is the local repulsive force gradient
static
ksample and can be calculated as a function of the indentation d at each point.

AFM experimental protocol


The experimental setup is as follows: Before beginning the AFM imaging,
cells were rinsed thoroughly with 2 mL PBS twice and then 2 mL of fresh ster-
ile PBS was added to simulate near-physiological conditions when imaging.
All live cell AFM imaging and viscoelastic measurements were performed
with a model No. MFP-3D-Bio AFM (Asylum Research, Santa Barbara,
CA) mounted on a model No. IX-71 inverted optical microscope (Olympus,
Melville, NY) which was itself placed on a vibration table inside an acoustic
isolation enclosure. This allows easy positioning and monitoring of cells. We
used soft microcantilevers (model No. BL-TR400PB microcantilever;
Olympus) with a nominal spring constant of 0.09 N/m, and nominal tip radius
of 42 nm (512 nm). The iDrive (Asylum Research) AC mode was used for
the experiment that uses Lorentz force excitation to apply an oscillating
driving force directly to the microcantilever. The importance of using such
directly excited probes for quantitative measurements in liquids has been dis-
cussed before (61,63). During each experimental measurement, we first
localized the cells using the inverted optical microscope, and checked for
viability by means of cell morphology and extracellular matrix spreading/
anchorage.
Before doing experiments on live fibroblast cells to measure quantitative
local viscoelastic properties, certain calibrations need to be performed. The
FIGURE 1 (a) Schematic of a Lorentz force excited microcantilever in-
AFM probe must be engaged on a stiff substrate (mica) and a force-distance
teracting with a live fibroblast cell by dynamic spectroscopy. In conven-
curve obtained to calculate the optical photodiode deflection sensitivity.
tional quasi-static spectroscopy a static cantilever is ramped toward the
Then, the probe is withdrawn from the stiff sample surface and the canti-
soft cell and retracted to record an F-Z curve. However, in dynamic spec-
lever spring constant was calibrated by the thermal noise method (64,65).
troscopy a microcantilever is excited at near the resonant frequency and
Typical estimated values for effective cantilever spring constants were in
moved toward the sample and then retracted to record three channels of in-
the range of 0.030.1 N/m and Q-factors were in the range of 1.52.
formation: mean deflection (A0), first harmonic amplitude, and phase lag
The AFM cantilever probe was directly driven (Lorentz force excitation) at
(A1 and f1). (b) Tune curves performed at 4 mm away and in contact
the peak of the resonance curve of the fundamental flexural mode (typically
with the cell with a force of ~1 nN show that the transfer functions of a fully
7.59 kHz) and then engaged to the live fibroblast cell in its buffer solution.
vibrated cantilever changes when interacting with the cell, resulting in
Phase contrast optical imaging was used to identify a viable cell on which to
changes in amplitude and phase that are closely related to the cell visco-
perform AFM imaging and spectroscopy. We then image the topography of
elastic properties. Therefore, an equivalent spring-dashpot model for the
the cell using AM-AFM while simultaneously acquiring the multi-harmonic
cantilever in contact with the cell can be used to extract the viscoelastic
data. The free amplitudes far from the sample (A1far) are typically in the range
properties. To see this figure in color, go online.
of 316 nm. The phase lag (f1far), from the low quality factor Qfar ~ 1.8, is set
to 74 using the third expression in Eq. 2. When brought closer to the cell, the
amplitude reduces to A1near ~ 25 nm and the phase lag changes to f1near ~85 multi-harmonic theory described above to reconstruct frequency- and inden-
100 just before the probe tip begins to interact with the cell surface. This tation-dependent local force and dissipation gradients on live fibroblast cells.
change in amplitude and phase from far to near the sample is a consequence Here we present data acquired on three fibroblast cells from different cell
of the strong squeeze-film hydrodynamic effect between the cantilever and cultures on different days using different AFM probes (10 repeats). The re-
the cell surface (66). The AM-AFM images are typically acquired at setpoints sults are consistent and repeatable. Many additional cells have been studied,
A1/A1near ~ 7080% at scan speeds of 0.25 Hz. The data were filtered using but we present those for which comprehensive quasi-static and dynamic
low-pass filters with cutoff frequencies of 500 Hz for mean deflection (0th data sets were acquired.
harmonic), and 1 kHz for 1st harmonic data. The images were rendered
with the software IGOR PRO 6.2 (WaveMetrics, Lake Oswego, OR).
After obtaining the AM-AFM topography and multi-harmonic variables
images, three locations of interest (nucleus, periphery, and gelatin-coated RESULTS AND DISCUSSION
dish) are identified. Subsequent approach curves were performed at these Cantilever response while interacting with a live
points to measure their static and dynamic viscoelastic properties. Fig. 1
a shows the schematic diagram of the experiment. Ten quasi-static and dy-
cell
namic approach curves with Z-piezo frequency of 0.25 Hz (linear speed of
Fig. 1 b shows the cantilevers frequency response to Lorentz
750 nm/s) were performed on the cell locations of interest. For quasi-static
curves, the local repulsive force gradient is simply the local slope on the F-d
force excitation off and in contact with the live cell nucleus
curve. For dynamic approach curves, the acquired multi-harmonic and periphery as well as the gelatin-coated dish. This clearly
observables curves (A0, A1, and f1) against Z were used together with the shows that the cell adds stiffness and damping to the

Biophysical Journal 106(5) 10331043


Viscoelasticity of Live Cells Investigated Using AFM 1037

cantilever oscillation, confirming the validity of modeling the is >~100 times over that of the nucleus, to ~510 times over
interaction as a Kelvin-Voigt viscoelastic element consisting the peripheral region and ~25 times over the gelatin-coated
of a repulsive force gradient and viscous dashpot model. dish. In other words, for the microcantilever to reach a spe-
Before describing the detailed results, it is worth high- cific amplitude reduction, it requires a large net indentation
lighting repeatable key features one can observe from into the living cell. This implies that when imaging, the
dynamic approach curves on the cell: AFM tip is in permanent contact with a large net indentation
dynamic
Our first important finding is that the AFM tip requires allowing the use of linear theory to extract ksample and
dynamic
large net sample indentation to reduce the cantilever ampli- csample , as described in Materials and Methods.
tude to a setpoint value A1 not only on the cell nuclear region The second important finding is that the resonance fre-
but also on the periphery. To demonstrate this we measured quency of the cantilever shifts significantly due to hydrody-
dynamic approach curves at different locations in the image namic squeezing effects from initial measurements taken
and observe the mean indentation, oscillation amplitude while tuning to when it begins to interact with the sample.
reduction, and phase-lag shift (Fig. 2). In Fig. 2 it can be This can be observed in Fig. 2, bd, by noting that the phase
clearly seen that as the cantilever tip is pressed down against lag before interaction begins is ~85100 even though while
the top of the cell, the first harmonic amplitude A1 decreases tuning (at ~4 mm far from the cell surface) the phase lag was
very slowly whereas the zeroth harmonic amplitude A0 in- set to 75 whereas the drive frequency was chosen to match
creases, rapidly becoming >A1. From Fig. 2 b it can be the resonance peak. As mentioned, we use Lorentz force exci-
deduced that on top of the cell nucleus it takes ~400 nm tation to drive the microcantilever near resonance; therefore,
of Z-piezo extension beyond initial interaction with the this phase shift is due to near-surface hydrodynamic effects
cell surface to reduce the AFM probe oscillation amplitude that have been considered in the theory presented before.
to A1 ~ 0.75 A1near. Also, from Fig. 2, c and d, on the cell
periphery region and gelatin-coated dish it takes ~50 and
Depth-dependent viscoelastic properties
25 nm of Z-piezo extension, respectively, to reduce the
amplitude to a value of A1 ~ 0.75 A1near. Thus, at the imaging We then measured the viscoelastic properties of live
setpoint A1 ~ 0.75 A1near, indentation/amplitude ratio (d/A1) fibroblasts cells using the dynamic and quasi-static AFM

FIGURE 2 Dynamic approach curves performed on different locations across the sample. (a) Three-dimensionally rendered topography image of a live
fibroblasts. (Crosses) Measurement locations. (bd) Dynamic-approach curves for multi-harmonic observables (A0, A1, and f1) acquired in the nucleus and
periphery of a live fibroblast cell and the stiffer gelatin-coated petri dish, respectively. (Green vertical lines) Typical imaging setpoint ratios of 85 and 75%.
(Curves) Behavior of the multi-harmonic amplitudes and phase (A0, A1, and f1) as the microcantilever moves toward and interacts with different regions of a
live fibroblast cell in a soft repulsive regime. It is clear that the strongest and most easily accessible harmonic signals that reflect local viscoelastic properties
are those from the 0th and 1st harmonics for live cells. Topography image was acquired using a Lorentz force excited Olympus (Melville, NY) microcanti-
lever (BL-TR400PB: kcant 0.083 N/m, Q 1.9, udr 7.79 kHz, and A1far 8.5 nm). To see this figure in color, go online.

Biophysical Journal 106(5) 10331043


1038 Cartagena and Raman

spectroscopy methods as described in Materials and Sample dynamic damping cdynamic sample also strongly depends
Methods, and obtained the following key results: on indentation depth becoming larger as the tip is pressed
When comparing the static and dynamic force gradients down on the cell. Dynamic damping was also extracted for
(at 0*udr and udr, respectively) on the cell nucleus as a func- the two regions (Fig. 3 c) showing that there is a variation
tion of indentation (Fig. 3 a), the dynamic values are three in damping across the cell. The plots in Fig. 3 c are the aver-
times larger than those measured statically until a critical ages of 10 repeats at the same location on the cell. The data
depth of ~100 nm, beyond which the ratio of dynamic force are highly repeatable with a typical variance of 10% over the
dynamic static
gradient ksample to ksample becomes larger. However, in the 10 repeats for all the cells discussed in this article. The damp-
dynamic
case of the periphery shown in Fig. 3 b, ksample and ing behavior for the two regions shows an intrinsic critical
static
ksample are comparable and vary in a similar manner with indentation depth where it deviates from zero value and in-
average indentation. These two plots are the averages of creases monotonically. Finally, we note that the damping is
10 repeats at the same location on the cell. The data are larger on the periphery than the nucleus.
highly repeatable with a typical variance of 10% over the The viscoelastic properties of live fibroblast cells
10 repeats for all the cells discussed in this article. From exhibited a weak dependence on the operating oscillation
dynamic static
Fig. 3, a and b, we note the ratio of ksample to ksample strongly amplitude. Dynamic force-indentation curves at different
depends on indentation depth on the nuclear region but not cantilever free oscillation amplitudes from 5 to 25 nm for
significantly on the cell periphery. a maximum loading force of ~2.5 nN were performed and
no significant difference in dynamic viscoelastic properties
was observed. This suggests that dynamic viscoelastic
properties measurements are independent of the free oscilla-
tion amplitude.
Fig. 4 shows the extracted viscoelasticity properties of
three different fibroblasts cells from different cell cultures.
In the nuclear region of all cells the indentation depth
dynamic
dependency is observed, where the ratio of ksample
static
to ksample increases beyond a certain indentation while on
dynamic static
the periphery ksample and ksample follow a similar depth
dynamic
dependence. Also, csample on the nucleus and periphery
of all cells demonstrates the indentation-depth dependency.
Our observation that the ratio of dynamic and static stiff-
ness increases beyond a certain indentation level on the
nuclear region, but not on the periphery, could arise from
the following possibilities:
As the AFM probe approaches and is being pressed
down into the cell, the AFM tip initially interacts with
brushes and the brush-type structures layer (combination
of glycocalyx, microvilli, and microridges) on the cell
membrane surface (67,68), and it is possible that the visco-
elastic properties of these extramembrane components are
very different from that of the underlying subcellular
components leading to the observed depth-dependent
divergence of the static and dynamic stiffness. However,
if this were the reason for the observations, the depth-
dependent divergence of static and dynamic stiffness
FIGURE 3 Frequency- and indentation-depth dependence of the visco- should be detected not only in the nuclear region but across
elastic properties measured on a live fibroblast cell. The dynamic visco- the entire cell, which is not the case.
elastic properties have been measured at a fixed high excitation
dynamic Another explanation for the observed divergence between
frequency of udr 7.79 kHz. (a and b) Dynamic ksample (blue) and static
static
ksample (red) force gradients at different mean indentations d0 on the nuclear the dynamic and static force gradients on the nuclear region
dynamic static
and peripheral region of a live fibroblast cell. Increase of ksample /ksample could be that the increase in dynamic and static force
force gradients after a critical indentation depth is observed in the cell nu- gradient ratio begins when the tip indents sufficiently to
cleus but not in the periphery. (c) Damping cdynamic
sample at different mean in- interact with the nuclear complex. The nuclear complex is
dentations d0 on the nucleus and periphery region of the cell. The plots
expected to have quite different viscoelastic properties
are the averages of 10 dynamic and static approach curves performed in
each location on the cell. (Insets) Images are the topography of the cell (69) compared to the cytoskeleton that lays just beneath
probed. (Crosses) Area where the measurements were made. To see this the cell membrane. This hypothesis is consistent with
figure in color, go online. the observation that the indentation-depth-dependent

Biophysical Journal 106(5) 10331043


Viscoelasticity of Live Cells Investigated Using AFM 1039

dynamic static
FIGURE 4 Viscoelastic properties response for different fibroblast cells. (a) Dynamic ksample (solid lines) and static ksample (dashed lines) force gradients
and (b) damping cdynamic
sample as a function of mean indentations d 0 on the nuclear region. (c and d) Same measurements are collected on the peripheral region. The
dynamic static
plots represent the averages of 10 dynamic and static approach curves performed in each specified location on each cell. Increase of ksample /ksample force
gradients in the nucleus is observed for every cell, which confirm our major finding that there are difference in viscoelastic properties in the nucleus and
is initially frequency-dependent, but after a critical indentation depth is frequency- and indentation-dependent. The dynamic viscoelastic properties of
the cells have been measured using three different cantilevers at a fixed high excitation frequency udr and with kcant and Q. Cell No. 1: 7.79 kHz,
0.083 N/m, and 1.9. Cell No. 2: 7.69 kHz, 0.076 N/m, and 1.8. Cell No. 3: 8.18 kHz, 0.076 N/m, and 1.8. (Insets) Topography images of the cells. (Crosses)
Locations where the measurements were made. To see this figure in color, go online.

dynamic
divergence of static and dynamic stiffness is observed only The map in Fig. 5 f suggests that ksample over the
over the nuclear region. nuclear region is greater than in the peripheral part of
the cell and the gelatin-coated dish. Moreover, from
Fig. 5 g, the cdynamic
sample over the nuclear region is lower in
Implications on mapping in vitro local magnitude than the periphery and dish. This result was
viscoelastic properties also observed in Raman et al. (4). On the other hand,
Now that we have studied the viscoelastic properties as a many prior works have shown that the gelatin-coated
function of indentation depth at high excitation frequency dish has the largest elastic modulus, followed by the pe-
on different cell locations, we turn our attention to visco- ripheral region, and lastly the nucleus (45,47,48). Also,
elastic property maps that can be easily extracted in AM- it was expected that the cell nucleus should be the surface
AFM scan over live cells. In Fig. 5 ,we show a series of with higher damping compared to the peripheral region
AM-AFM images acquired over the live fibroblast cell in and the dish. This apparent contradiction between the
physiological conditions. The multi-harmonic observables maps in Fig. 5 and known properties can be explained
(A0 and f1) are converted to detailed local material prop- in terms of the force spectroscopy results presented earlier.
erty maps of effective dynamic repulsive force gradient Fig. 6 shows a repulsive gradient with respect to applied
dynamic
ksample and damping cdynamic as described earlier. We force plot on the nucleus of a live fibroblast cell. It
sample dynamic
also convert these maps into loss tangent tand (Fig. 5 h). can be clearly seen that, at forces >200 pN, ksample in-
The loss tangent is the ratio between the energy dissipated creases rapidly. We actually apply forces >750 pN when
and the energy stored in one cycle of oscillation in contact imaging in AM-AFM over the nuclear region, and as a
dynamic
with the sample (70). In this case, the loss tangent is result we observe large ksample . However, on the periph-
defined as eral region (Fig. 6), we apply very small forces <150
pN while imaging in AM-AFM yielding relatively small
  dynamic
tan d cdynamic  ksample , lesser than the nucleus. At such low forces, the
sample u dr =k dynamic :
sample force gradient on the gelatin is also very small in

Biophysical Journal 106(5) 10331043


1040 Cartagena and Raman

FIGURE 5 Multi-harmonic AFM images acquired on a live fibroblast cell in physiological conditions. (a) Optical phase contrast micrograph using a 10
objective; (b) topographic image; (c) 0th harmonic amplitude map A0; (d) first harmonic phase map f1; (e) applied mean normal force map F0ts; and (fh)
dynamic
maps of local dynamic force gradient ksample , damping cdynamic th
sample , and loss tangent tand, respectively, extracted from the measured 0 and first harmonic data
using the theory described in the text. Surprisingly, the loss tangent values of the cell periphery and dish are larger than the cell nucleus. Images were acquired
using a Lorentz force excited Olympus (Melville, NY) microcantilever (BL-TR400PB: kcant 0.083 N/m, Q 1.9, and udr 7.79 kHz). The scale bar on
images represents 10 mm (size; 60  60 mm2, pixels; 256  256). To see this figure in color, go online.

comparison to the force gradient on the cell at much 2. Cell microrheological models have not been validated for
higher forces. high frequencies; and
3. A549 cells lack the cytoskeletal element concentration
and organization that characterize fibroblasts.
Comparison with microrheological models of live
cells The viscoelastic properties behavior dependence on inden-
tation depth over more limited frequency ranges (20
It is interesting to compare the results in this work to prior
400 Hz) of polymer gels and NIH3T3 fibroblast cells has
works on the rheology and viscoelasticity of live cells. Pre-
been shown before by Mahaffy et al. (57,58). However,
vious work by Alcaraz et al. (51) has used a power-law fre-
the divergence between low and high frequency as a func-
quency-dependent structural damping model to estimate the
tion of sample indentation has not been presented before.
complex shear modulus G*(u) of a live cell. This model has dynamic
Moreover, to the best of our knowledge, ksample data on
been validated systematically at low frequencies (~0.1 the nucleus and periphery regions reported in this work
100 Hz), although not at the high frequencies (510 kHz)
are the first AFM high-frequency- and indentation-depen-
as presented in this work. A benchmark calculation using
dent (~8 kHz) nanorheological measurements in live cells.
values reported by Alcaraz et al. (51) for A549 human
This finding has a major impact when comparing quantita-
lung epithelial cells, suggests Gstorage (8 kHz)/Gstorage
dynamic tive results obtained from quasi-static and high-frequency
(0.25 Hz) ~ 9.8. However, the ratio of ksample udr to
static dynamic methods.
ksample is smaller than the one predicted by the rheological
power-law model ~3.56.5. There could be several reasons
for this: CONCLUSIONS
1. The rheological models do not take into account the We have demonstrated the ability of dynamic AFM to
indentation depth dependence; quantify the nanorheological properties of live cells and

Biophysical Journal 106(5) 10331043


Viscoelasticity of Live Cells Investigated Using AFM 1041

500 nm) over the nuclear region and much lesser on the pe-
riphery (~50 nm). The combination of two effects (i.e., high
frequency vibration and variation of indentation depth) per-
formed while imaging the cell leads to the dynamic repul-
sive force gradient on the nucleus to be generally greater
than those mapped using standard quasi-static methods.
These results confirm that dynamic AFM methods can in
fact be used for the quantitative mapping of viscoelastic
properties of subcellular components of nuclear and periph-
eral regions of live cells. However, the interpretation of
these properties and comparison with quasi-static AFM
measurements requires careful consideration of the fre-
quency- and indentation-dependence. We have demon-
strated the frequency- and indentation-dependence of local
viscoelastic properties of living cells by comparing the force
gradients determined from dynamic and quasi-static force
spectroscopy methods, reporting, for the first time to our
knowledge, measurements for high frequencies with an
elucidation of the biomechanical role of subcellular compo-
nents. This has significant relevance not only for cell mecha-
FIGURE 6 Plots are the averages of 10 dynamic force gradients of (a) nobiology but also for AFM-based imaging and force
dynamic
ksample and (b) damping cdynamic
sample against the applied force performed on
spectroscopy of live cells.
the nucleus (blue), periphery (red) of the live fibroblast cell, and the
gelatin-coated petri dish (green). At forces higher than ~800 pN on the
cell nucleus the dynamic force gradient and damping becomes higher SUPPORTING MATERIAL
than the periphery and dish, increasing rapidly as the cell indentation pro-
gressively increases. This explains the stiffness and damping contrast One table, seven figures, References (7173), and Supplemental informa-
between the peripheral and nuclear regions observed in Fig. 5s maps. To tion are available at http://www.biophysj.org/biophysj/supplemental/
see this figure in color, go online. S0006-3495(14)00011-3.

The authors thank Prof. E. Nauman (School of Mechanical Engineering and


Weldon School of Biomedical Engineering, Purdue University) for
compared the dynamic and quasi-static methods to under- providing the rat fibroblast cells.
stand the viscoelastic response of live fibroblast cells at This research was supported in part by the National Science Foundation
very low frequencies (quasi-static) and at the cantilever Materials World Network under grant No. DMR 1008189.
resonance frequency (dynamic).
We have also found interesting differences between
viscoelastic measurements made using quasi-static and REFERENCES
dynamic AFM modes. On the fibroblasts nuclear region
1. Suresh, S. 2007. Biomechanics and biophysics of cancer cells. Acta
the local dynamic force gradient is larger than the statically Biomater. 3:413438.
measured one for small indentations (0250 nm), however, 2. Guck, J., F. Lautenschlager, ., M. Beil. 2010. Critical review: cellular
beyond a certain critical indentation depth (say >250 nm) mechanobiology and amoeboid migration. Integr. Biol. (Camb). 2:
the ratio of force gradients derived from the dynamic and 575583.
quasi-static methods increases rapidly due to interactions 3. Azeloglu, E. U., and K. D. Costa. 2011. Atomic force microscopy in
with subcellular components such as the nuclear complex. mechanobiology: measuring microelastic heterogeneity of living cells.
Methods Mol. Biol. 736:303329.
This depth dependency is also seen in the viscous damping.
4. Raman, A., S. Trigueros, ., S. A. Contera. 2011. Mapping nanome-
However, on the peripheral parts of the fibroblasts the ratio chanical properties of live cells using multi-harmonic atomic force
of dynamic and static force gradients does not change appre- microscopy. Nat. Nanotechnol. 6:809814.
ciably with indentation. 5. Putman, C. A. J., K. O. van der Werf, ., J. Greve. 1994. Viscoelasticity
Consequently, it is difficult to compare maps of visco- of living cells allows high resolution imaging by tapping mode atomic
elastic properties acquired using the quasi-static and force microscopy. Biophys. J. 67:17491753.
dynamic AFM modes. Specifically, when these properties 6. Bausch, A. R., W. Moller, and E. Sackmann. 1999. Measurement of
local viscoelasticity and forces in living cells by magnetic tweezers.
are mapped over live cells in buffer media using AM- Biophys. J. 76:573579.
AFM, the indentation required to maintain constant oscilla- 7. Fabry, B., G. N. Maksym, ., J. J. Fredberg. 2001. Scaling the micro-
tion amplitude changes from pixel to pixel because the rheology of living cells. Phys. Rev. Lett. 87:148102.
nanomechanical properties on the cell are heterogeneous. 8. Trepat, X., L. Deng, ., J. J. Fredberg. 2007. Universal physical
Specifically, the tip indents the cell much more (300 responses to stretch in the living cell. Nature. 447:592595.

Biophysical Journal 106(5) 10331043


1042 Cartagena and Raman

9. Vadillo-Rodriguez, V., and J. R. Dutcher. 2009. Dynamic viscoelastic 32. Matthews, B. D., D. R. Overby, ., D. E. Ingber. 2004. Mechanical
behavior of individual Gram-negative bacterial cells. Soft Matter. properties of individual focal adhesions probed with a magnetic micro-
5:50125019. needle. Biochem. Biophys. Res. Commun. 313:758764.
10. Ingber, D. E. 1993. The riddle of morphogenesis: a question of solution 33. Maciaszek, J. L., and G. Lykotrafitis. 2011. Sickle cell trait human
chemistry or molecular cell engineering? Cell. 75:12491252. erythrocytes are significantly stiffer than normal. J. Biomech. 44:
657661.
11. Montell, D. J. 2008. Morphogenetic cell movements: diversity from
modular mechanical properties. Science. 322:15021505. 34. Rotsch, C., and M. Radmacher. 2000. Drug-induced changes of cyto-
skeletal structure and mechanics in fibroblasts: an atomic force micro-
12. Wang, N., J. P. Butler, and D. E. Ingber. 1993. Mechanotransduction scopy study. Biophys. J. 78:520535.
across the cell surface and through the cytoskeleton. Science. 260:
11241127. 35. Bhadriraju, K., and L. K. Hansen. 2002. Extracellular matrix- and cyto-
skeleton-dependent changes in cell shape and stiffness. Exp. Cell Res.
13. Ingber, D. E. 1997. Tensegrity: the architectural basis of cellular me- 278:92100.
chanotransduction. Annu. Rev. Physiol. 59:575599.
36. Brandt, R. 2001. Cytoskeletal mechanisms of neuronal degeneration.
14. Ingber, D. E. 2003. Tensegrity I. Cell structure and hierarchical systems Cell Tissue Res. 305:255265.
biology. J. Cell Sci. 116:11571173.
37. van Vliet, K. J., G. Bao, and S. Suresh. 2003. The biomechanics
15. Rotsch, C., K. Jacobson, and M. Radmacher. 1999. Dimensional and toolbox: experimental approaches for living cells and biomolecules.
mechanical dynamics of active and stable edges in motile fibroblasts Acta Biomater. 51:58815905.
investigated by using atomic force microscopy. Proc. Natl. Acad. Sci.
38. Bao, G., and S. Suresh. 2003. Cell and molecular mechanics of biolog-
USA. 96:921926.
ical materials. Nat. Mater. 2:715725.
16. Yamazaki, D., S. Kurisu, and T. Takenawa. 2005. Regulation of cancer 39. Benoit, M., and H. E. Gaub. 2002. Measuring cell adhesion forces with
cell motility through actin reorganization. Cancer Sci. 96:379386. the atomic force microscope at the molecular level. Cells Tissues
17. Discher, D. E., P. Janmey, and Y. L. Wang. 2005. Tissue cells feel and Organs (Print). 172:174189.
respond to the stiffness of their substrate. Science. 310:11391143. 40. Arnoldi, M., M. Fritz, ., A. Boulbitch. 2000. Bacterial turgor pressure
18. Lo, C. M., H. B. Wang, ., Y. L. Wang. 2000. Cell movement is guided can be measured by atomic force microscopy. Phys. Rev. E Stat. Phys.
by the rigidity of the substrate. Biophys. J. 79:144152. Plasmas Fluids Relat. Interdiscip. Topics. 62 (1 Pt B):10341044.
19. Guck, J., S. Schinkinger, ., C. Bilby. 2005. Optical deformability as 41. Touhami, A., B. Nysten, and Y. F. Dufrene. 2003. Nanoscale mapping
an inherent cell marker for testing malignant transformation and met- of the elasticity of microbial cells by atomic force microcopy. Lang-
astatic competence. Biophys. J. 88:36893698. muir. 19:45394543.
20. Cross, S. E., Y. S. Jin, ., J. K. Gimzewski. 2007. Nanomechanical 42. Dupres, V., F. D. Menozzi, ., Y. F. Dufrene. 2005. Nanoscale mapping
analysis of cells from cancer patients. Nat. Nanotechnol. 2:780783. and functional analysis of individual adhesins on living bacteria. Nat.
Methods. 2:515520.
21. Darling, E. M., S. Zauscher, ., F. Guilak. 2007. A thin-layer model for
43. Longo, G., L. M. Rio, ., S. Kasas. 2012. Force volume and stiffness
viscoelastic, stress-relaxation testing of cells using atomic force micro-
tomography investigation on the dynamics of stiff material under bac-
scopy: do cell properties reflect metastatic potential? Biophys. J. 92:
terial membranes. J. Mol. Recognit. 25:278284.
17841791.
44. Dufrene, Y. F., D. Martnez-Martn, ., D. J. Muller. 2013. Multipara-
22. Cross, S. E., Y. S. Jin, ., J. K. Gimzewski. 2008. AFM-based analysis metric imaging of biological systems by force-distance curve-based
of human metastatic cancer cells. Nanotechnology. 19:384003. AFM. Nat. Methods. 10:847854.
23. Faria, E. C., N. Ma, ., R. D. Snook. 2008. Measurements of elastic 45. Radmacher, M., M. Fritz, ., P. K. Hansma. 1996. Measuring the visco-
properties of prostate cancer cells using AFM. Analyst (Lond.). elastic properties of human platelets with the atomic force microscope.
133:14981500. Biophys. J. 70:556567.
24. Kumar, S., and V. M. Weaver. 2009. Mechanics, malignancy, and 46. A-Hassan, E., W. F. Heinz, ., J. H. Hoh. 1998. Relative microelastic
metastasis: the force journey of a tumor cell. Cancer Metastasis Rev. mapping of living cells by atomic force microscopy. Biophys. J. 74:
28:113127. 15641578.
25. Goetz, J. G., S. Minguet, ., M. A. Del Pozo. 2011. Biomechanical re- 47. Haga, H., S. Sasaki, ., T. Sambongi. 2000. Elasticity mapping of
modeling of the microenvironment by stromal caveolin-1 favors tumor living fibroblasts by AFM and immunofluorescence observation of
invasion and metastasis. Cell. 146:148163. the cytoskeleton. Ultramicroscopy. 82:253258.
26. Pelling, A. E., F. S. Veraitch, ., M. A. Horton. 2009. Mechanical dy- 48. Solon, J., I. Levental, ., P. A. Janmey. 2007. Fibroblast adaptation and
namics of single cells during early apoptosis. Cell Motil. Cytoskeleton. stiffness matching to soft elastic substrates. Biophys. J. 93:44534461.
66:409422.
49. Darling, E. M. 2011. Force scanning: a rapid, high-resolution approach
27. Gourlay, C. W., and K. R. Ayscough. 2005. The actin cytoskeleton: a for spatial mechanical property mapping. Nanotechnology. 22:175707.
key regulator of apoptosis and aging? Nat. Rev. Mol. Cell Biol. 50. Prabhune, M., G. Belge, ., M. Radmacher. 2012. Comparison of
6:583589. mechanical properties of normal and malignant thyroid cells. Micron.
28. Ilic, D., Y. Furuta, ., T. Yamamoto. 1995. Reduced cell motility and 43:12671272.
enhanced focal adhesion contact formation in cells from FAK-deficient 51. Alcaraz, J., L. Buscemi, ., D. Navajas. 2003. Microrheology of
mice. Nature. 377:539544. human lung epithelial cells measured by atomic force microscopy.
29. Pelham, Jr., R. J., and Y. Wang. 1997. Cell locomotion and focal adhe- Biophys. J. 84:20712079.
sions are regulated by substrate flexibility. Proc. Natl. Acad. Sci. USA. 52. Gavara, N., and R. S. Chadwick. 2012. Determination of the elastic
94:1366113665. moduli of thin samples and adherent cells using conical atomic force
30. Balaban, N. Q., U. S. Schwarz, ., B. Geiger. 2001. Force and focal microscope tips. Nat. Nanotechnol. 7:733736.
adhesion assembly: a close relationship studied using elastic micropat- 53. Darling, E. M., S. Zauscher, and F. Guilak. 2006. Viscoelastic proper-
terned substrates. Nat. Cell Biol. 3:466472. ties of zonal articular chondrocytes measured by atomic force micro-
31. Riveline, D., E. Zamir, ., A. D. Bershadsky. 2001. Focal contacts as scopy. Osteoarthritis Cartilage. 14:571579.
mechanosensors: externally applied local mechanical force induces 54. Vadillo-Rodriguez, V., T. J. Beveridge, and J. R. Dutcher. 2008. Surface
growth of focal contacts by an mDia1-dependent and ROCK-indepen- viscoelasticity of individual Gram-negative bacterial cells measured
dent mechanism. J. Cell Biol. 153:11751186. using atomic force microscopy. J. Bacteriol. 190:42254232.

Biophysical Journal 106(5) 10331043


Viscoelasticity of Live Cells Investigated Using AFM 1043

55. Bausch, A. R., F. Ziemann, ., E. Sackmann. 1998. Local measure- 64. Hutter, J. L., and J. Bechhoefer. 1993. Calibration of atomic-force
ments of viscoelastic parameters of adherent cell surfaces by magnetic microscope tips. Rev. Sci. Instrum. 64:18681873.
bead microrheometry. Biophys. J. 75:20382049.
65. Butt, H.-J., and M. Jaschke. 1995. Calculation of thermal noise in
56. Tung, R. C., A. Jana, and A. Raman. 2008. Hydrodynamic loading of atomic force microscopy. Nanotechnology. 6:17.
microcantilevers oscillating near rigid walls. J. Appl. Phys. 104: 66. Xu, X., C. Carrasco, ., A. Raman. 2008. Unmasking imaging forces
114905. on soft biological samples in liquids when using dynamic atomic force
57. Mahaffy, R. E., C. K. Shih, ., J. Kas. 2000. Scanning probe-based fre- microscopy: a case study on viral capsids. Biophys. J. 95:25202528.
quency-dependent microrheology of polymer gels and biological cells. 67. Sokolov, I., S. Lyer, ., C. D. Woodworth. 2007. Detection of surface
Phys. Rev. Lett. 85:880883. brush on biological cells in vitro with atomic force microscopy. Appl.
58. Mahaffy, R. E., S. Park, ., C. K. Shih. 2004. Quantitative analysis of Phys. Lett. 91:023902.
the viscoelastic properties of thin regions of fibroblasts using atomic 68. Iyer, S., R. M. Gaikwad, ., I. Sokolov. 2009. Atomic force micro-
force microscopy. Biophys. J. 86:17771793. scopy detects differences in the surface brush of normal and cancerous
59. Xu, X., and A. Raman. 2007. Comparative dynamics of magnetically, cells. Nat. Nanotechnol. 4:389393.
acoustically, and Brownian motion driven microcantilevers in liquids. 69. Jamali, Y., M. Azimi, and M. R. K. Mofrad. 2010. A sub-cellular visco-
J. Appl. Phys. 102:034303. elastic model for cell population mechanics. PLoS ONE. 5:e12097.
60. Xu, X., M. Koslowski, and A. Raman. 2012. Dynamics of surface- 70. Proksch, R., and D. G. Yablon. 2012. Loss tangent imaging: theory and
coupled microcantilevers in force modulation atomic force micro- simulations of repulsive-mode tapping mode atomic force microscopy.
scopymagnetic vs. dither piezo excitation. J. Appl. Phys. 111: Appl. Phys. Lett. 100:073106.
054303.
71. Cartagena, A., M. Hernando-Perez, ., A. Raman. 2013. Mapping
61. Kiracofe, D., and A. Raman. 2011. Quantitative force and dissipation in vitro local material properties of intact and disrupted virions at
measurements in liquids using piezo-excited atomic force microscopy: high resolution using multi-harmonic atomic force microscopy. Nano-
a unifying theory. Nanotechnology. 22:485502. scale. 5:47294736.
62. Rabe, U., K. Janser, and W. Arnold. 1996. Vibrations of free and sur- 72. Dulebo, A., J. Preiner, ., P. Hinterdorfer. 2009. Second harmonic
face-coupled atomic force microscope cantilevers: theory and experi- atomic force microscopy imaging of live and fixed mammalian cells.
ment. Rev. Sci. Instrum. 67:32813293. Ultramicroscopy. 109:10561060.
63. de Beer, S., D. van den Ende, and F. Mugele. 2008. Atomic force mi- 73. Kiracofe, D., J. Melcher, ., S. Hu. 2013. VEDA: virtual environment
croscopy cantilever dynamics in liquid in the presence of tip sample for dynamic AFM. https://nanohub.org/resources/veda. http://dx.doi.
interaction. Appl. Phys. Lett. 93:253106. org/10.4231/D3KW57J1T.

Biophysical Journal 106(5) 10331043

You might also like