You are on page 1of 7

Probing the Interior of Living Cells

with Fluorescence Correlation Spectroscopy


MATTHIAS WEISS
Cellular Biophysics Group (BIOMS), German Cancer Research Center,
Heidelberg, Germany

To a large extent the cellular interior is occupied by two complex fluids, the cytoplasm and the
nucleoplasm, both of which show a considerable degree of macromolecular crowding. While it is
easy to imagine that the chromosomal DNA provides the nucleoplasm with properties similar to
a polymer melt, the material properties of the cytoplasm are also affected by the high amount
of dissolved macromolecules, the cytoskeletal network, and dispersed organelles. By virtue of
the strongly obstructed random motion, reactions in the cytoplasm and nucleoplasm are not
comparable to the aqueous conditions commonly used in biochemical experiments. To overcome
this gap, a thorough understanding of the material properties of intracellular fluids, and hence
transport properties within the cell, is mandatory. Here, we review some recent results on bulk
diffusion in living cells and some generic consequences that arise from these observations.

Key words: fluorescence correlation spectroscopy; anomalous diffusion; subdiffusion; viscoelas-


ticity; macromolecular crowding; diffuse-to-capture

Introduction and may therefore fall short in giving an appropriate


decription of the cell’s behavior on the molecular scale.
The interior of living cells is an amazingly complex To overcome this limitation, a thorough understanding
microscosm in which myriad proteins and lipids act in of the cellular material properties, such as cytoplasmic
concert to sustain vital cellular processes.1 During the viscosity, is mandatory, as these properties lie at the
last decades, molecular biology has boosted our under- very heart of all transport processes, even of the most
standing of cellular dynamics and reaction networks by basic one, namely diffusion.
making virtually any protein accessible to genetic ma-
nipulations, be it by silencing or knocking out specific
genes or by attaching genetically encoded fluorescent A Primer on Diffusion
tags to the protein of interest. Despite these advances
that have shaped our current thinking in terms of well- Diffusion was first characterized in 1827 by Robert
defined cellular pathways and networks, one must not Brown, who observed with his microscope the erratic
forget that cells lack an organizing mastermind but motion of small pollen particles in aqueous solution.
rather have to self-organize on the molecular level due While Brownian motion initially was a mere experi-
to the fundamental laws of physics and chemistry. A mental observation, the foundation of statistical me-
particular aspect of this self-organization is the molec- chanics by Boltzmann provided a basis for a solid un-
ular communication within the cell, that is, proteins derstanding of the curious phenomenon. Indeed, with
have to sample their environment for interaction part- his seminal article on the theory of Brownian motion in
ners to be able to establish a reaction network. Char- 1905, Einstein gave an elegant theoretical basis to dif-
acterizing cellular pathways solely by kinetic equations fusion4 that strongly influenced 20th century physics.
(which indeed is a frequent approach in many systems Since Einstein’s seminal paper, diffusion has been stud-
biology studies2,3 ) neglects this important spatial aspect ied in a variety of contexts that sometimes are far be-
yond Brown’s original observations, for instance, when
considering the erratic motion of stock market indices5
Address for correspondence: Dr. Matthias Weiss, Cellular Biophysics or the diffusive spreading of quantum wave packets in
Group (BIOMS), B085, German Cancer Research Center, BIOQUANT
Center, BQ0019, Im Neuenheimer Feld 267, D-69120 Heidelberg,
nonlinear6 and disordered media.7
Germany. Voice: +49 6221 5451304. A basic fact about diffusion is its random charac-
m.weiss@dkfz.de ter, that is, all diffusing particles perform a random

Ann. N.Y. Acad. Sci. 1130: 21–27 (2008). 


C 2008 New York Academy of Sciences.
doi: 10.1196/annals.1430.002 21
22 Annals of the New York Academy of Sciences

walk thereby supporting our everyday experience that


diffusion is an irreversible process that aims at flatten-
ing out existing concentration gradients. In this exam-
ple, diffusion acts on behalf of the second law of ther-
modynamics, which demands that molecules spread
over the available volume so that the entire system
achieves a state of maximum entropy. But how can one
quantify the molecules’ erratic motion? Using a sim-
ple one-dimensional model is very instructive at this
point.8 Let us consider the path of a person who tries
to get back home after an extensive pub-crawl. Stag-
gering from street lamp to street lamp, the person has
instantly forgotten from which of the two neighbor-
ing lamps he or she came from and will hence stagger FIGURE 1. Mean square displacement  r(t)2  ∼ t α for
randomly to either one of them. Having a complete ab- normal diffusion (full line, α = 1) and subdiffusion (dashed,
sence of any sense of direction the net movement—the dash-dotted: α = 0.75, 0.5). The change in scaling is best
seen in a double-logarithmic plot (inset) where the power
sum of (positive) steps to the right and (negative) steps
laws appear as straight lines with different slope.
to the left—will tend toward zero. Nevertheless, the
person still makes excursions from the starting point
(the pub) and consequently the mean square displace-
ment (MSD), calculated from the starting position, is Quantifying Diffusion with
non-zero. Going through the details,   one finds that Fluorescence Correlation Spectroscopy
the MSD grows linearly in time, r (t )2 = 2D t , where
the prefactor D is the diffusion coefficient. In fact, the Having noticed that diffusion may be anomalous,
linear growth of the MSD with time is a fingerprint how can one actually assess the diffusional properties
of normal diffusion, and the equation can easily be of individual molecules in living cells? A particularly
expanded to diffusion in any dimension via r (t )2 = useful method in this context is fluorescence corre-
dim ·2D t . lation spectroscopy (FCS). The roots of FCS can be
If we suppose now that the person does not al- traced back to the early 1970s,11 but it was not until
ways move immediately to another lamp after hav- the 1990s and the advent of confocal microscopy that
ing reached one but rather takes a rest at each lamp the technique became more feasible and sufficiently
for a certain time (e.g., admiring the beauty of the sensitive.12 In FCS, a laser beam with a bell-shaped
pavement), the character of the diffusion may change intensity profile is focused on a spot of interest, such
dramatically. While only the diffusion coefficient may as inside a living cell, and a pinhole is then used to
be altered slightly, when the resting times T follow an discriminate the fluorescent light emitted from differ-
exponential distribution (resting is thus a Poisson
 pro- ent focal sections of the sample (FIG. 2A). In this way,
cess), the MSD will assume a scaling r (t )2 ∝ t α when one effectively constrains the collection of photons to a
the resting times T follow a broad power-law distribu- small diffraction-limited volume of about 1 µm3 (called
tion,2 that αis, p (T) ∼ 1/T with α < 1. The scaling
1+α the confocal volume). In contrast to confocal imaging,
r (t ) ∝ t with α < 1 is called anomalous diffusion not the average fluorescence at the illuminated spot
or, to be more precise, subdiffusion, and although but rather the fluctuations around the mean fluores-
we still observe an “unbiased” random walk, the ef- cence is the quantity of interest. Due to the diffusion
ficiency of moving away from the starting point has of fluorescently tagged molecules that enter and leave
decreased qualitatively as it now takes more and the confocal volume, the fluorescence signal rises and
more time to reach the same MSD as compared drops (FIG. 2A). The particles’ Brownian movement is
to normal diffusion (cf. FIG. 1). While we have cho- therefore reflected in the fluctuations f (t) of the flu-
sen here to introduce subdiffusion via a construc- orescence signal F (t ) = F  + f (t ) around the mean
tion called continuous time random walk (CTRW, F , and a noise analysis of the fluctuations can re-
see Ref. 9 for a more thorough introduction), there veal the mean residence time of a fluorescent particle
are several other causes for subdiffusion,10 some of in the confocal volume. In fact, the fluctuations be-
which will be highlighted in the remainder of this come stronger and better visible when only few labeled
article. particles are in the confocal volume. In other words,
Weiss: Probing Cells with FCS 23

FIGURE 2. (A) Sketch of an FCS experiment. Fluorescent molecules diffuse into and out of the confocal
volume, thereby giving rise to fluctuations in the fluorescence time series. (B) The autocorrelation function
C (τ) shows a more shallow decay for α < 1 (dashed, dash-dotted: α = 0.75, 0.5) as compared to normal
diffusion (α = 1, full line).

FCS works best at very low expression levels (1 nM A


C (τ) = 
concentration is sufficient). In this sense, FCS is an 1 + τ/(S 2 τD ) · [1 + τ/τD ]
on-average single-molecule technique with which one
can study diffusion in cells almost in their native state Here, S denotes the unavoidable elongation of the con-
without perturbing them by massively overexpressing focal volume along the optical axis while the radius r
a fluorescently tagged protein. of the confocal volume yields the correlation half time
To determine the mean residence time τ D in τ D = r 2 /(4D). Fitting the theoretically derived formula
the confocal volume, one calculates the autocorrela- to experimental data, one can determine the diffu-
tion
 function of the fluorescence fluctuations C (τ) = sion coefficient D and from the offset A = C(τ = 0)
f (t ) · f (t + τ) (with ... denoting a temporal aver- also the mean number N  of particles in the con-
age). Assuming normal diffusion in bulk and a three- focal volume as A ∝ 1/ N .13 In fact, A is not only
dimensional Gaussian shape for the confocal volume, proportional to the inverse of the mean number of
one can derive analytically the expected form for the particles but includes also the photophysics of the flu-
autocorrelation decay to be: orophores. A (small) fraction f T of them will enter the
24 Annals of the New York Academy of Sciences

nonfluorescent triplet state (having a lifetime τ T ) af- Testing the Crowded State
ter excitation and decay without radiation back to the of Intracellular Fluids
ground state. This can be taken into account by setting
A = (1 + f T exp(−t /τT )) / N . It is noteworthy that In contrast to most biochemical assays, the interior
perturbations of the confocal volume, due to local vari- of cells is by no means an aqueous solution. Rather,
ations in the refractive index, for example, may strongly the cytoplasm and the nucleoplasm of living cells are
influence the above autocorrelation function.14,15 crowded solutions with 50–400 g/l of dissolved macro-
For describing anomalous diffusion, the autocorre- molecules (proteins, DNA, RNA, etc.).27,28 Thus, one
lation function can be expanded easily only under a may expect that the material properties differ signif-
few simplifying assumptions. While one would have to icantly from buffer solutions. Indeed, early reports
use the proper description in terms of, say, a fractional on the diffusion of fluorescent proteins in the cyto-
Fokker–Planck equation9 or fractional Brownian mo- plasm of living cells have indicated a three foled to four
tion,16,17 one may also simply argue that a time-varying fold higher viscosity as compared to buffer solutions.29
diffusion coefficient D (t ) ∝ t α−1 is responsible for the Later, a slightly anomalous diffusion of fluorescent pro-
sublinear growth of the MSD. With this mathemat- teins was observed with FCS in the nucleus,30 indicat-
ically incorrect yet helpful assumption, one can use ing a scale-dependent viscosity and the emergence of
again a Gaussian propagator to describe the diffusion an elastic response. In fact, due to its single-molecule
process and one ends up with character, FCS is perfectly suited to determine the ma-
terial properties of intracellular fluids and the diffu-
A sion of macromolecules without perturbing the cell
C (τ) =   α   too much.
1 + τ/(S 2 τD ) · 1 + (τ/τD )α By microinjection or electroporation, nanometer-
sized fluorescently labeled gold beads or dextran con-
This function is easy to use in fitting procedures structs can be transferred into the cytoplasm and nu-
and essentially gives the same result as a more com- cleus of a variety of mammalian cell lines (from hamster
plicated description by the fractional Fokker–Planck ovary cells to human osteosarcoma cells), and their
equation.18 In FIGURE 2B, we have shown the typical (anomalous) diffusion can be quantified with FCS.
decay of C(τ) for different α while keeping A = 1 and From the autocorrelation curve, one can extract the
τ D = 5 ms fixed. MSD from which one eventually can derive the com-
plex shear modulus G(ω) (see above). With this ap-
proach, one can thus investigate the local viscoelastic-
From Diffusion to Material Properties ity of intracellular fluids on the nanoscale—on length
scales of ∼100 nm—while single-particle tracking of
From the diffusion characteristics, one can deduce larger beads24,25 or the deformation of entire cells in
the viscoelastic properties of a medium in terms of an optical stretcher31 test the mechanical properties of
the complex shear modulus G(ω) = G (ω) + iG (ω).19 cells on larger length scales.
It describes a fluid’s viscous (G ) and elastic (G ) re- In virtually all cell lines approached with FCS, a
sponse when perturbing it with an oscillatory shear strong anomalous diffusion in the cytoplasm and nu-
strain of frequency ω. Water, as a purely viscous fluid, cleoplasm has been observed when using gold parti-
is described by G = 0 and G = ωη, while rubber is cles26,32 and dextrans.18 The anomality 0.5 < α < 0.8
entirely characterized by a nonvanishing elastic mod- for dextran probes was mass dependent, which most
ulus G with no viscous component (G = 0). Measur- likely reflects the peculiar polymeric nature of the
ing the MSD of a diffusing inert, spherical test particle probe. With the more well defined gold beads, cell-
that tests the fluid’s viscoelastic properties, one can cal- to-cell variations were observed, yet the value α ≈ 0.55
culate G(ω) from the Laplace-transformed MSD via for both intracellular fluids did not differ significantly
analytical continuation.19 In fact, this approach has between the tested cell lines.32 The observation and the
been used successfully for determining the viscoelastic- degree of subdiffusion in mammalian cells also com-
ity in F-actin solutions,20–22 colloidal suspensions near pare favorably to reports on the anomalous diffusion
the glass transition,23 and in intracellular fluids.24–26 in the cytoplasm of yeast cells33 and bacteria.34
Thus, tracing the diffusive motion of a spherical parti- Observing a subdiffusive behavior of macro-
cle by means of FCS or single-particle tracking yields molecules in living cells immediately triggered the
an elegant and noninvasive way to study cellular vis- question of its origin. Naively, one may want to trace
coelasticity in the living cell. back the peculiar diffusion to a CTRW (see above).
Weiss: Probing Cells with FCS 25

Yet a CTRW-like scenario, due to cooperative or


hierarchical binding, for example, can only provide an
explanation when the system is continuously driven out
of thermal equilibrium.35 The more likely explanation
is thus an obstructed random walk due to (immobile)
obstacles. At a certain density of obstacles, the so-called
percolation threshold, the diffusion characteristics be-
comes anomalous with α ≈ 0.55.10 This phenomenon
has been investigated extensively in many contexts, in-
cluding for obstructed diffusion on surfaces36,37 and
for transport in porous media.38 The drawback of this
explanation is the condition of having immobile obsta-
cles.
An alternative to the model of obstructed diffusion
is the motion in a viscoelastic fluid. Due to the elastic
restoring forces, a purely viscous motion is impossi-
ble, thus rendering the diffusion anomalous. Indeed,
extracting the complex shear modulus from the FCS
curves and the underlying MSD revealed a strongly
viscoelastic behavior for the cytoplasm and nucleo-
plasm of all tested cell lines, that is, G(ω) showed a
power-law dependence on the excitation frequency ω,
|G(ω)| ∼ ωα with 0.5 ≤ α ≤ 0.6. A particular example
for the frequency dependence of the shear modulus FIGURE 3. (Top) Absolute value of the complex shear
G(ω) as extracted from FCS measurements on U2OS modulus, |G(ω)|, measuring the viscoelastic response of the
cells is shown in FIGURE 3. In general, the viscous mod- cytoplasm of a U2OS cell when shearing it with frequency
ulus (G ) was about two to five fold larger than the ω. A clear power-law increase is observed (dashed lines)
elastic modulus (G ) for frequencies ω ≤ 1 Hz, while be- for the cytoplasm (full symbols) and nucleoplasm (open sym-
bols), indicating a stiffening of the cytoplasm when excit-
yond 100 Hz both moduli were about equal. In other
ing it with a higher frequency. (Bottom) The ratio G /G
words, the higher the frequency of the shearing the between elastic (G ) and viscous (G ) modulus highlights
stiffer and more viscous did the fluids behave. The the dominantly viscous behavior at small frequencies, while
phenomenon of a crowding-induced viscoelasticity is both moduli become comparable at higher frequencies.
also observed for other (artificially) crowded solutions,
such as diluted frog egg extracts26 or buffer solution
diameter of some 100 nm also experience organelles
with a high concentration of crowding agents like ficoll
and the cytoskeletal array of semiflexible filaments as
or dextran.18,39 Heuristically, the observed viscoelas-
important obstacles. Single-particle tracking of larger
ticity of crowded fluids shows a similarity to the Zimm
beads consequently reported larger values for G(ω) as
model for dilute polymer solutions at good solvent con-
higher-order structures yield an additional contribu-
ditions. Here, α = 5/9 is predicted in very good agree-
tion to the crowding-induced viscoelasticity.24,25
ment with the experimental observations. Taking this
model seriously, a change in solvent conditions (to the
θ-solvent) would result in a change of the exponent to-
wards α = 2/3. Indeed, applying hypo-osmotic stress Consequences of Anomalous
to living cells by adding 500 mM sucrose to the medium Diffusion in Viscoelastic Fluids
changes the solvent conditions inside the cell and a
change of the anomalous diffusion toward α = 0.65 is Regardless of the underlying microscopic reasons
observed.26 Thus, the Zimm model for dilute polymer for subdiffusion in the living cell—whether obstructed
solutions provides a reasonable heuristic model for the diffusion or immersion in a viscoelastic fluid—the mere
behavior of intracellular fluids. occurrence of subdiffusion has major implications for
It is noteworthy that the size of the probe particle is biological processes. It has been shown by computer
of great importance when determining the viscoelas- simulations that obstructed diffusion at the perco-
ticity. A 5 nm-sized bead adequately probes the envi- lation threshold can dramatically influence the rate
ronment on the nanoscale, while latex beads with a with which binary40 and enzymatic reactions41 take
26 Annals of the New York Academy of Sciences

(i.e., anomalous diffusion) has a trail with dimension


2/α. Since α can be as small as 0.5, a more than
space-filling path emerges for searching the binding
partners which enhances the probability of reliably
finding a reaction partner. Thus, cells indeed benefit
from their crowded state by a facilitated search algo-
rithm for reaction partners in the cytoplasm and/or
nucleoplasm.

Acknowledgment

This work was supported by the Institute for Mod-


eling and Simulation in the Biosciences (BIOMS) in
FIGURE 4. The probability P (R) of finding a target with Heidelberg. I would like to thank G. Guigas for helpful
(anomalous) diffusion when starting off at a distance R (cf.
comments on the manuscript.
inserted sketch) depends on the anomality α. For normal
diffusion, P (R) ∼ 1/R (full line), while subdiffusion yields a
strongly increased probability (α = 0.8, 0.7, 0.5 dotted, Conflict of Interest
dashed, dash-dotted). Inset: When increasing the available
search time, subdiffusion also yields a higher P(R) when The author declares no conflicts of interest.
starting in a larger distance R (α = 0.5, 1 dash-dotted, full).

References
place. Also, subdiffusion can stabilize the formation of
spatiotemporal Turing patterns in activator–inhibitor
1. ALBERTS, B. 2002. Molecular biology of the cell. Garland
scenarios42 by mimicking a slow diffusion of the ac- Science. New York.
tivator. Indeed, an anomality α = 0.9 is sufficient to 2. KOLCH, W., M. CALDER & D. GILBERT. 2005. When ki-
stabilize a spatial pattern in the presence of strong par- nases meet mathematics: the systems biology of MAPK
ticle number fluctuations. signalling. FEBS Lett. 579: 1891–1895.
In more general terms, one may ask if subdiffu- 3. BRUGGEMAN, F. & H. WESTERHOFF. 2007. The nature of
sion should have a major impact for biology given systems biology. Trends Microbiol. 15: 45–50.
4. EINSTEIN, A. 1905. Über die von der molekularkinetischen
that it seems to simply reduce the spreading of reac-
Theorie der Wärme geforderte Bewegung von in ruhen-
tion partners. It is worthwhile to approach this prob- den Flüssigkeiten suspendierten Teilchen. Annalen der
lem in terms of a diffuse-to-capture scenario (FIG. 4). Physik. 17: 123.
Here, a point-like test particle is started in a distance 5. BONESS, A.J. & F.C. JEN. 1970. A model of information
R from a spherical binding partner of radius a. Via diffusion, stock market behavior, and equilibrium price.
(anomalous) diffusion, the test particle tries to find the Journal of Financial and Quantitative Analysis 5: 279–
binding partner within a given time t max . Perform- 296.
6. KOTTOS, T. & M. WEISS. 2004. Current relaxation
ing this search for many particles, one can record in nonlinear random media. Phys. Rev. Lett. 93:
the capture probability P(R). For normal bulk diffu- 190604.
sion, a simple argument shows that P(R) ∼ 1/R.43 In 7. WEISS, M., T. KOTTOS & T. GEISEL. 2001. Spreading and
contrast, subdiffusion shows a strongly enhanced, α- localization of wavepackets in disordered wires in a mag-
dependent probability to find the target provided that netic field. Phys. Rev. B 63: 081306.
t max is large enough to bridge the initial distance R via 8. WEISS, M. & T. NILSSON. 2004. In a mirror dimly: tracing
the movements of molecules in living cells. Trends in Cell
anomalous diffusion. For increasing t max , subdiffusion
Biology 14: 267–273.
turns out to be even more efficient, that is, allowing for 9. METZLER, R. & J. KLAFTER. 2000. The random walk’s guide
extremely large search times considerably enhances to anomalous diffusion: a fractional dynamics approach.
the probability of finding a target via subdiffusion. In Physics Reports-Review Section of Physics Letters 339:
other words, subdiffusion may be the slower way to 1–77.
find reaction partners but it will be the more reliable 10. BOUCHAUD, J.P. & A. GEORGES. 1990. Anomalous diffusion
search algorithm.44 Essentially, this feature arises as a in disordered media – statistical mechanisms, models and
physical applications. Physics Reports-Review Section of
consequence of the fractal properties of the searcher’s
Physics Letters 195: 127–293.
random walk: While normal Brownian motion has a 11. MAGDE, D., W.W. WEBB & E. ELSON. 1972. Thermody-
plane-filling search path, fractional Brownian motion namic fluctuations in a reacting system – measurement by
Weiss: Probing Cells with FCS 27

fluorescence correlation spectroscopy. Phys. Rev. Lett. 29: 29. SWAMINATHAN, R., C. HOANG & A. VERKMAN 1997.
705–708. Photobleaching recovery and anisotropy decay of green
12. RIGLER, R. & E.S. ELSON. 2001. Fluorescence correlation fluorescent protein GFP-S65T in solution and cells: cy-
spectroscopy theory and applications. Springer. Berlin. toplasmic viscosity probed by green fluorescent protein
13. SCHWILLE, P. 2001. Fluorescence correlation spectroscopy translational and rotational diffusion. Biophys. J. 72:
and its potential for intracellular applications. Cell 1900–1907.
Biochem. Biophys. 34: 383–408. 30. WACHSMUTH, M., W. WALDECK & J. LANGOWSKI. 2000.
14. HESS, S. & W. WEBB. 2002. Focal volume optics and ex- Anomalous diffusion of fluorescent probes inside living
perimental artifacts in confocal fluorescence correlation cell nuclei investigated by spatially-resolved fluorescence
spectroscopy. Biophys J. 83: 2300–2317. correlation spectroscopy. J. Mol Biol. 298: 677–689.
15. ENDERLEIN, J. et al. 2004. Art and artifacts of fluorescence 31. GUCK, J. et al. 2005. Optical deformability as an in-
correlation spectroscopy. Curr. Pharm. Biotechnol. 5: herent cell marker for testing malignant transforma-
155–161. tion and metastatic competence. Biophys. J. 88: 3689–
16. MANDELBROT, B.B. 1985. Self-affine fractals and fractal di- 3698.
mension. Physica Scripta 32: 257–260. 32. GUIGAS, G., C. KALLA, & M. WEISS, 2007. The degree of
17. SEBASTIAN, K.L. 1995. Path integral representation for frac- macromolecular crowding in the cytoplasm and nucleo-
tional Brownian motion. J. Phys. A: Math. Gen. 28: 4305– plasm of mammalian cells is conserved. FEBS Lett. 581:
4311. 5094–5098.
18. WEISS, M. et al. 2004. Anomalous subdiffusion is a mea- 33. TOLIC-NORRELYKKE, I. et al. 2004. Anomalous diffusion in
sure for cytoplasmic crowding in living cells. Biophysical living yeast cells. Phys. Rev. Lett. 93: 13.
Journal 87: 3518–3524. 34. GOLDING, I. & E. COX, 2006. Physical nature of bacterial
19. MASON, T. & D. WEITZ. 1995. Optical measurements of cytoplasm. Phys Rev Lett. 96: 10.
frequency-dependent linear viscoelastic moduli of com- 35. SAXTON, M. 2007. A biological interpretation of transient
plex fluids. Phys. Rev. Lett. 74: 1250–1253. anomalous subdiffusion. I. Qualitative model. Biophys. J.
20. MASON, T. et al. 2000. Rheology of F-actin solutions deter- 92: 1178–1191.
mined from thermally driven tracer motion. J. Rheol. 44: 36. SAXTON, M. 1993. Lateral diffusion in an archipelago.
917–928. Single-particle diffusion. Biophys. J. 64: 1766–1780.
21. GARDEL, M. et al. 2003. Microrheology of entangled F-actin 37. SAXTON, M. 1993. Lateral diffusion in an archipelago. De-
solutions. Phys. Rev. Lett. 91: 7. pendence on tracer size. Biophys. J. 64: 1053–1062.
22. THARMANN, R., M. CLAESSENS & A. BAUSCH. 2007. Vis- 38. MAKSE, H. et al. 1996. Long-range correlations in perme-
coelasticity of isotropically cross-linked actin networks. ability fluctuations in porous rock. Phys. Rev. E. 54: 3129–
Phys. Rev. Lett. 98: 21. 3134.
23. MASON, T. & D. WEITZ. 1995. Linear viscoelasticity of col- 39. BANKS, D. & C. FRADIN. 2005. Anomalous diffusion of pro-
loidal hard sphere suspensions near the glass transition. teins due to molecular crowding. Biophys. J. 89: 2960–
Phys. Rev. Lett. 75: 2770–2773. 2971.
24. TSENG, Y., T. KOLE & D. WIRTZ. 2002. Micromechani- 40. SAXTON, M.J. 2002. Chemically limited reactions on a per-
cal mapping of live cells by multiple-particle-tracking mi- colation cluster. Journal of Chemical Physics 116: 203–
crorheology. Biophys. J. 83: 3162–3176. 208.
25. TSENG, Y. et al. 2004. Micro-organization and visco-elasticity 41. BERRY, H. 2002. Monte Carlo simulations of enzyme reac-
of the interphase nucleus revealed by particle nanotrack- tions in two dimensions: fractal kinetics and spatial segre-
ing. J. Cell Sci. 117: 2159–2167. gation. Biophys. J. 83: 1891–1901.
26. GUIGAS, G., C. KALLA & M. WEISS. 2007. Probing the 42. WEISS, M. 2003. Stabilizing Turing patterns with subdiffu-
nanoscale viscoelasticity of intracellular fluids in living sion in systems with low particle numbers. Phys. Rev. E.
cells. Biophys. J. 93: 316–323. 68: 036213.
27. ELLIS, R. & A. MINTON. 2003. Cell biology: join the crowd. 43. BERG, H.C. 1993. Random Walks in Biology. Princeton
Nature 425: 27–28. University Press.
28. MINTON, A. 2006. How can biochemical reactions within 44. GUIGAS, G. & M. WEISS. 2008. Sampling the cell
cells differ from those in test tubes? J. Cell Sci. 119: 2863– with anomalous diffusion – the discovery of slowness.
2869. Biophys. J. 94: 90–94.

You might also like