ArticlePDF Available

Enhanced Diffusion of Enzymes that Catalyze Exothermic Reactions

Authors:

Abstract and Figures

Enzymes have been recently found to exhibit enhanced diffusion due to their catalytic activities. A recent experiment [C. Riedel et al., Nature 517, 227 (2015)] has found evidence that suggests this phenomenon might be controlled by the degree of exothermicity of the catalytic reaction involved. Four mechanisms that can lead to this effect, namely, self-thermophoresis, boost in kinetic energy, stochastic swimming, and collective heating, are critically discussed, and it is shown that only the last two could be strong enough to account for the observations. The resulting quantitative description is used to examine the biological significance of the effect.
Content may be subject to copyright.
arXiv:1508.03219v1 [q-bio.BM] 13 Aug 2015
Enhanced Diffusion of Enzymes that Catalyze Exothermic Reactions
Ramin Golestanian
Rudolf Peierls Centre for Theoretical Physics, University of Oxford, Oxford OX1 3NP, UK
(Dated: August 14, 2015)
Enzymes have been recently found to exhibit enhanced diffusion due to their catalytic activities.
A recent experiment [C. Riedel et al., Nature 517, 227 (2015)] has found evidence that suggests this
phenomenon might be controlled by the degree of exothermicity of the catalytic reaction involved.
Four mechanisms that can lead to this effect, namely, self-thermophoresis, boost in kinetic energy,
stochastic swimming, and collective heating, are critically discussed, and it is shown that only
the last two could be strong enough to account for the observations. The resulting quantitative
description is used to examine the biological significance of the effect.
PACS numbers: 87.14.ej,87.10.Ca,65.80.-g,87.16.Uv
Introduction.—A most fascinating aspect of the
nonequilibrium processes in living cells is active transport
[1]. The basic units of these processes, which could be in
the form of carrying cargo or sliding actin fibres against
one another, are motor proteins that convert chemical
energy directly into useful mechanical work, amidst dom-
inant thermal fluctuations at the nano-scale [2]. Recent
in vitro studies of mixtures of motors and filaments have
revealed their remarkable ability to self-organize into dy-
namic meso-scale structures that resemble those observed
in living cells [35]. Much less is known about the na-
ture of the nonequilibrium activity of non-cytoskeletal
elements and how they self-organize in living cells.
It has been recently reported that enzymes undergo
enhanced diffusion, i.e. diffusive motion with an effec-
tive diffusion coefficient Deff that is larger than its equi-
librium value D0, as a result of their catalytic activity
[6, 7]. Considering the enzyme as a sphere of radius R
in a medium with viscosity ηand temperature T, the
Stokes-Einstein relation D0=kBT gives us the equilib-
rium diffusion coefficient of the enzyme, where ζ= 6πηR
is its friction coefficient. The additional, nonequilibrium,
contribution to the diffusion coefficient, D=Deff D0,
is found to be proportional to the net rate (or speed) of
the catalytic reaction. The rate has the characteristic
Michaelis-Menten form k=keS/(KM+S), where Sis
the substrate (i.e. reactant) concentration, KMis the
Michaelis constant, and keis the enzyme reaction rate.
Remarkably, Dhas the same order of magnitude as the
equilibrium diffusion coefficient; typically a fraction of it.
It has also been observed that there is a strong correla-
tion between the degree of exothermicity of the catalytic
reaction and the enhancement in the effective diffusion
coefficient of enzymes [8]. In this Letter, I discuss and
critically examine various mechanisms that can lead to
enhanced diffusion for catalytically active enzymes.
Self-phoresis.—Any colloidal particle that actively gen-
erates nonequilibrium phoretic flow in its vicinity exhibits
enhanced effective diffusion at time scales longer than its
orientational persistence time 1/Dr, where Dris the ro-
tational diffusion coefficient [9]. Enzyme activity has the
right ingredients to lead to enhanced diffusion via self-
diffusiophoresis, which takes advantage of gradients in
the concentrations of the chemicals involved in the reac-
tion [10]. However, this contribution is not sensitive to
the degree of exothermicity of the reaction. The appro-
priate mechanism that could account for such an effect is
self-thermophoresis [11, 12]. The heat released from the
chemical reaction during each catalytic cycle, Q, leads to
a temperature difference TkQ/(κR) across the en-
zyme, where κis the thermal conductivity of the medium.
For catalase k= 5 ×104s1,Q= 40 kBT, and R= 4 nm,
which gives D0= 55 µm2s1and Dr= 2.6×106s1.
Using these values and κ0.6 W/(m ·K) for water, we
obtain T106K. The self-propulsion velocity is es-
timated as Vst D0STT/R D0STkQ/(κR2) [11],
where STis the Soret coefficient of the enzyme. Using
ST0.02 K1[13], we find Vst 104µm s1. The
correction to effective diffusion coefficient due to self-
thermophoresis is DV2
st/Dr, for which we find D
1014 µm2s1, and consequently D/D01016. This
is fifteen orders of magnitude too small to account for the
observations.
Boost in Kinetic Energy.—The authors of Ref. [8]
propose a scenario in which the heat released from the
chemical reaction during each catalytic cycle is channeled
into a boost in the translational velocity of the enzyme.
This is not envisaged to be mediated through an effec-
tive temperature increase following the release of heat.
Here I reproduce their analysis using a slightly differ-
ent derivation, to highlight the essence of the proposed
mechanism. Consider the enzyme to be a particle of
mass mwhose stochastic motion satisfies Newton’s equa-
tion md2r
dt +ζdr
dt =f(t), where f(t) is a random force.
The relative significance of the inertial and the dissipa-
tive terms in the equation of motion is characterized by
the time scale τ=m/ζ . Using m=4π
3R3ρpwith a
typical protein mass density ρp= 1.4×103kg/m3and
η= 1 ×103Pa·s for viscosity of water at room tempera-
ture, we find τ= 5 ps. Invoking an elegant mathematical
trick that Langevin used in his original 1908 paper [14],
2
we can write the equation of motion as
τd
dt + 1Deff (t) = 2
3ζE(t),(1)
where Deff (t) = 1
6
d
dt hr(t)2iis by definition the effective
diffusion coefficient and E(t) = hm
2dr
dt 2iis the average
kinetic energy of the enzyme. Here we have assumed a
separation of time scales between the random thermal
kicks that the enzyme receives from the medium and the
catalytic cycle, and the averaging is performed over the
thermal kicks. We can write E(t) = 3
2kBT+γQh(t) where
γrepresents the fraction of the released thermal energy
that is converted into the translational boost and h(t)
is a series of spikes of width τbthat appear stochasti-
cally at a rate k, through a Poisson process. Here, τbis
the relaxation time of the boost, which depends on the
specific process that generates it. It is reasonable to as-
sume that τbτ. The boost mechanism proposed in
Ref. [8] involves asymmetric excitation of compressional
waves along the enzyme that propagate to the interface
with water and trigger a pressure wave that leads to a
back-reaction on the enzyme itself, giving it a mechani-
cal boost. No evidence is provided in Ref. [8] as to why
the energy is not randomly partitioned between a large
number of possible channels (owing to the large number
of degrees of freedom or normal modes), which would re-
sult in γ1, and subsequently dissipated, as opposed to
being channeled to a small number of modes (correspond-
ing to γ1). Time averaging gives E=3
2kBT+γQkτb,
and consequently
D
D0
Ref.[8]
=2
3
γQ
kBTb,(2)
through Eq. (1). Using the above estimate for τband the
values for kand Qcorresponding to catalase, we obtain
b= 2.5×107and D/D0=γ×105from Eq. (2).
Even with the (unrealistic) maximum value of γ= 1, our
estimate from Eq. (2) is four orders of magnitude too
small to account for the observations.
Stochastic Swimming.—We could examine various hy-
drodynamic effects that might contribute towards such
a behaviour. Substrate binding could change the shape
of the enzyme, and consequently its friction coefficient.
Since this process does not require energy input, however,
it cannot be the cause of a nonequilibrium phenomenon.
Such conformational changes are often relatively small,
and will more likely lead to an increase in size that would
result in a decrease in diffusion coefficient, rather than
the other way around. Moreover, it is not clear why such
an effect could lead to universal trends—as it will depend
on specific cases—and how it can correlate with exother-
micity.
As an alternative scenario, it is possible that the cat-
alytic cycle induces conformational changes in the en-
zyme that lead to stochastic swimming [15]. The am-
plitude of these deformations is typically much smaller
than the size of the enzyme, e.g. when they arise from
mechanochemical coupling of electrostatic nature [16]
(analogous to phosphorylation) or structural changes due
to ligand binding [17]. However, local heat release could
have the possibility to transiently disturb the relatively
more fragile tertiary structure of the folded protein [18]
or the state of oligomerization of the enzyme [19], and
produce an amplitude bthat is a fraction of the size R.
To calculate the contribution of such conformational
changes to effective diffusion coefficient, we use a sim-
ple model in which the conformational change is de-
scribed by one degree of freedom L(t) representing elon-
gation of the structure along an axis defined by a unit
vector ˆ
n(t). To achieve directed swimming, we need
at least two degrees of freedom to incorporate the co-
herence needed for breaking the time-reversal symme-
try at a stochastic level, and we know that realistic con-
formational changes must involve many degrees of free-
dom. The randomization of the orientation, described via
hˆ
n(t)·ˆ
n(t)i=e2Dr|tt|, will turn the directed motion
into enhanced diffusion over the time scales longer than
1/Dr. Since the same can be achieved through recip-
rocal conformational changes described by one compact
degree of freedom, I will adopt this simpler form. The
stochastic motion of the enzyme can be described by the
Langevin equation v(t)αd
dt Lˆ
n(t) + ξ(t) where α
is a numerical pre-factor that depends on the geometry
of the enzyme [20, 21] and ξ(t) is the Gaussian white
noise that will give us the intrinsic translational diffusion
coefficient D0.
We describe the combined mechanochemical cycle us-
ing a two-step process, which takes the enzyme from its
free state to the reaction stage that is followed by the de-
formation with rate k, and a relaxation back to its native
state with rate kr. This is a simplification of a more real-
istic model with three states (free, substrate-bound, and
reacted-deformed) and kis to be understood as the com-
~ l
T
a
Heat Flux
T
real
T
app
FIG. 1. (color online.) Schematic illustration of the essence
of Newton’s law of cooling. The heat generated in the bulk of
a chamber (that is in contact with the environment that has
a fixed ambient temperature Ta), and lost in the form of heat
flux through the boundaries leads to the temperature profile
Treal (solid line), which is approximated by the average value
over the distance ,Tapp (dashed line). This approximation
works best when there is a separation of length scales.
3
0
0.1
Θ
1
kkmax
0
0.1
0.2
Θ
0.5
1
Ce
aCe
0.02
0.04
0.02
0.04
Θ
0.02
0.04
0.1
0.2
0.3
DDD0
0.005
0.01
0.015
0.05
0.1
Θ
(a) (b) (c) (d)
FIG. 2. (color online.) (a) The effective rate of catalytic reaction as a function of the reduced temperature θfor θd= 0.1,
g= 50, and ǫ= 7. The inset shows the fraction of catalytically active enzymes as a function of θin each case. (b) The effective
temperature of the catalytically active medium as a function of the coupling strength δ. (c) The corresponding relative increase
in diffusion coefficient. (d) The effective temperature for ǫ= 30, with the dashed region being unstable.
bined catalytic rate that has the Michaelis-Menten form
as defined above. In stationary state, a master equa-
tion formulation can be used to calculate the elongation
speed autocorrelation function as d
dt L(t)·d
dtL(t)=
2b2kkr
k+kr[δ(tt)1
2(k+kr)e(k+kr)|tt|]. By com-
bining this with the orientation auto-correlation, we can
calculate the effective diffusion coefficient of the enzyme,
which gives the following correction
D=1
3α2b2kkr
k+kr2Dr
2Dr+k+kr
.(3)
Even for the fastest enzymes, we typically have kr
Drk. Using an upper bound of b.R, we can ap-
proximate Eq. (3) as DkR2. For catalase, we ob-
tain D1µm2s1, which gives an upper bound of
D/D0102. This is one order of magnitude smaller
than the observed values.
Collective Heating.—In Ref. [8], a calculation similar
to what we have above to estimate the relative change
in temperature across the enzyme, T, is used to ar-
gue that heating of the environment by the enzyme is
negligible. This estimate, however, is only correct for an
isolated enzyme. In practice, an experiment is performed
on a solution with a finite concentration of enzyme, Ce.
For such a sample, the substrate is consumed at the rate
(per unit volume) of keSCa
e/(KM+S), where Ca
eis the
concentration of the catalytically active enzymes. The
exothermic catalytic reaction generates thermal energy
Qper turnover cycle at the location of each active en-
zyme, which then diffuses through the sample container
and escapes via the boundaries. Due to the large number
of heat producing enzymes (of the order of Avogadro’s
number) there will be a significant build-up of thermal
energy in the sample container. To see this, let us define
a length scale that describes the characteristic distance
heat needs to diffuse until it can exit. is typically set by
the smallest length scale in the geometry of the sample
container. We can estimate a characteristic heat diffu-
sion time τh() = 2, using the thermophoretic con-
ductivity χ. For water at room temperature, we have
χ105µm2s1. For = 10 mm, it takes τh= 1000 s for
the heat released from each enzyme during each catalytic
cycle to leave the container. Given Ce= 1 nM [8], during
this time 1020 units of Qwill have been released into the
chamber (assumed to have a volume 3); this is 10 J of
thermal energy.
The heat diffusion equation is written as
χ1tT 2T=Q
κ·keSC a
e
KM+S1
2(TTa).(4)
The right hand side of Eq. (4) contains a source term
that couples the catalytic reaction to the production of
heat, and a sink term that approximates the heat loss
through the boundaries by a bulk term in the form of
Newton’s law of cooling [22], where Tais the ambient
temperature. This term is written in terms of the length
scale that is described above. This approximation has
been widely used in the combustion literature to allow the
temperature-dependent nonlinearities in the source term
to be captured in a manner that does not depend on the
geometric specificities of each experiment [23]. Figure 1
summarizes the essence of this approximation.
Two quantities in the source term of Eq. (4) have sig-
nificant dependence on temperature, which we will rep-
resent using θ= (TTa)/Ta. The first quantity is the
turnover rate, which we can assume to have an Arrhenius
form of ke=k
0eEa/kBT, where Eais the activation en-
ergy. The rate can be rewritten as ke=k0eǫθ/(1+θ),
where ǫ=Ea/kBTaand k0=k
0eǫ. The second quan-
tity is the concentration of active enzymes. Since en-
zymes are proteins, increase in temperature will eventu-
ally denature them upon approaching the denaturation
temperature, which we denote by θd. We can use a simple
two-state model to account for the denaturation, which
yields Ca
e=Ce/eg(θθd)+ 1, where the parameter g
controls the sharpness of the transition. The two effects
enter the heat source term in Eq. (4) via the product
keCa
e, whose temperature dependence is shown in Fig.
4
2(a), for g= 50, θd= 0.1 (that corresponds to Td= 330K
for Ta= 300K), and ǫ= 7, which is a typical value for the
activation energy for enzymes, such as catalase [24]. The
plot shows the two standard regimes of initial increase
in the effective rate due to the Arrhenius temperature
dependence and the sudden decline due to denaturation.
The peak will sharpen with increasing activation energy
ǫ. The inset of Fig. 2(a) shows the fraction of active
enzymes as a function of temperature.
Writing the temperature dependencies in Eq. (4) ex-
plicitly, we find
τhtθ22θ=δeǫθ/(1+θ)
eg(θθd)+ 1 θ, (5)
where
δ=Qℓ2
κTa
·k0SCe
KM+S,(6)
emerges as a single dimensionless parameter that controls
the strength of collective heating. Assuming a uniform
profile, we can find the stationary state temperature of
the system as a function of δby setting the right hand
side of Eq. (5) to zero. The result is shown in Fig. 2(b).
Note that the temperature dependence of the kinetic rate
provides such a sensitive positive feedback mechanism
that could lead to unrealistically high temperatures, eas-
ily above enzyme denaturation and even above boiling
temperature of water (for a case such as catalase with
the large value of Q= 40 kBT). The presence of de-
naturation provides a negative feedback mechanism that
ensures such dramatic temperature increases are cut off.
The temperature increase can be used to calculate the
relative increase in diffusion coefficient D/D0, by tak-
ing into account the corresponding variations in the fric-
tion coefficient. Denaturation will change the hydrody-
namic radius of the protein from its globular (folded)
form, with size RaN 1/3, to its coiled (unfolded) form,
with size RgaN 3/5, where ais the Kuhn length and N
is the polymerization index. This yields RgRN 4/15 .
For catalase, we have a1 nm, N100, Rg14 nm.
To calculate the diffusion coefficient within our simple
two-state model, we need to use the ensemble average
of the inverse size, namely, use 1/R with the probabil-
ity p= 1/eg(θθd)+ 1and 1/Rgwith the probabil-
ity 1 p. We can use the following expression for the
temperature dependence of the viscosity of water η(θ) =
η0exp n(B/Ta)
1+θ(T0/Ta)owhere B= 579K, T0= 138K, and
η0= 2.41 ×105Pa·s. The result is plotted in Fig. 2(c),
showing a similar pattern of behaviour as seen in θ. The
values obtained for D/D0are of the same order of mag-
nitude as the experimentally observed values, while the
actual temperature increase in the solution is relatively
modest. Moreover, the trend in Fig. 2(c) resembles the
experimental results reported in Refs. [68].
To estimate typical experimental values for δ, we need
a typical length scale from the sample container and the
smallest conductivity involved in the geometry of the sys-
tem. For Fluorescence Correlation Spectroscopy (FCS)
experiments, a droplet of the solution is placed over a
glass slide. Using κ0.02 W/(m ·K) for air, = 5 mm,
and Ce= 1 nM (as used in Ref. [8]), k0= 5 ×104s1,
and Q= 40 kBT, we find δ0.02 at substrate satu-
ration. This estimate appears to provide a rough order
of magnitude agreement with the observations of Ref.
[8]. Note that due to our approximate treatment of the
boundary conditions, order unity differences are to be ex-
pected when comparing to the experimental results. The
experiments in Refs [6, 7] use higher concentrations of
enzymes and similar sample sizes, and thus comfortably
fall in the regime described by the collective heating sce-
nario.
Discussion.—The collective heating model leads to
very specific predictions that could be experimentally
tested. The near linear dependence of the relative en-
hancement in diffusion coefficient on δ[Fig. 2(c)] and
the definition of δ[Eq. (6)] suggest a linear dependence
on the enzyme concentration and a quadratic dependence
on the size of the container. In practice, protein denat-
uration is irreversible, which suggests that recording ex-
perimental data over a long time scale would presumably
lead to a systematic reduction in the magnitude of the en-
hanced diffusion, provided the experiment is done under
the condition that the substrate concentration is main-
tained at a constant level.
The heating mechanism can also lead to interesting
nonlinear phenomena. While at small values of ǫ, we
observe a near-linear dependence of temperature on δ
[Fig. 2(b)], upon increasing ǫthe curve takes an S-shape
that develops an instability at sufficiently large values of
ǫ, as shown in Fig. 2(d). The instability will lead to
the formation of waves, which will dissipate in a sealed
sample container when all the fuel molecules are con-
sumed, following closely the phenomenology of flames in
combustion [25]. Moreover, collective heating could have
synergistic influence on the other mechanisms: while the
increase in temperature could facilitate the emergence of
large conformational changes in the tertiary structure or
the oligomerization state during the enzymatic turnover,
phoretic collective heating can lead to further instabili-
ties [26] that could accentuate the degree of fluctuations
in the system.
Finally, let us examine whether the total heat gen-
erated in a cell could be sufficient to trigger this ef-
fect. Considering a cell of size = 10 µm that is fully
packed with enzymes similar to catalase (that gives us
Ce1 mM), we find δ0.01. While this is an upper
limit, it certainly points to a strong possibility that the
enhanced diffusion via collective heating could be a con-
tributing factor to non-directed intracellular transport in
living cells.
In conclusion, enhanced diffusion of enzymes that cat-
alyze exothermic reactions could be explained by a com-
5
bination of global temperature increase in the sample
container, and possibly enhanced conformational changes
that can lead to a hydrodynamic enhancement of effec-
tive diffusion coefficient. Self-thermophoresis and boost
in kinetic energy as suggested by Ref. [8] are too weak to
account for the experimentally measured values of effec-
tive diffusion. Although the primary focus of this work
has been on enhanced diffusion of enzymes, the theoret-
ical description should be relevant to the study of any
class of thermally activated microswimmers.
I have benefitted from discussions with Carlos Busta-
mante, Krishna Kanti Dey, and Ayusman Sen.
ramin.golestanian@physics.ox.ac.uk
[1] M.C. Marchetti J.F. Joanny, S. Ramaswamy, T.B. Liv-
erpool, J. Prost, M. Rao, and R.A. Simha, Rev. Mod.
Phys. 85, 1143 (2013)
[2] K Svoboda, C.F. Schmidt, B.J. Schnapp, S.M. Block,
Nature 365, 721 (1993).
[3] V. Schaller, C. Weber, C. Semmrich, E. Frey, and A.
Bausch, Nature 467, 73-77 (2010).
[4] T. Sanchez, D.T.N. Chen, S.J. DeCamp, M. Heymann,
and Z. Dogic, Nature 491, 431 (2012)
[5] Y. Sumino, K. Nagai, Y. Shitaka, D. Tanaka, K.
Yoshikawa, H. Chat`e, K. Oiwa, Nature 483, 448 (2012).
[6] H.S. Muddana, S. Sengupta, T.E. Mallouk, A. Sen, and
P.J. Butler, J. Am. Chem. Soc. 132, 2110 (2010).
[7] S. Sengupta, K.K. Dey, H.S. Muddana, T. Tabouillot,
M.E. Ibele, P.J. Butler, and A. Sen, J. Am. Chem. Soc.
135, 1406 (2013).
[8] C. Riedel, R. Gabizon, C.A.M. Wilson, K. Hamadani, K.
Tsekourasl, S. Marqusee, S. Presse, and C. Bustamante,
Nature 517, 227 (2015).
[9] R. Golestanian, Phys. Rev. Lett. 102, 188305 (2009).
[10] R. Golestanian, T.B. Liverpool, and A. Ajdari, Phys.
Rev. Lett. 94, 220801 (2005).
[11] R. Golestanian, T.B. Liverpool, and A. Ajdari, New. J.
Phys. 9, 126 (2007).
[12] H-R. Jiang, N. Yoshinaga, and M. Sano, Phys. Rev. Lett.
105, 268302 (2010).
[13] S.A. Putnam, D.G. Cahill, and G.C.L. Wong, Langmuir
23, 9221 (2007)
[14] P. Langevin, C. R. Acad. Sci. (Paris) 146, 530 (1908).
[15] R. Golestanian and A. Ajdari, Phys. Rev. Lett. 100
038101 (2008); JPCM 21, 204104 (2009).
[16] R. Golestanian, Phys. Rev. Lett. 105, 018103 (2010).
[17] T. Sakaue, R. Kapral, and A.S. Mikhailov, Eur. Phys. J.
B75, 381 (2010).
[18] O. Miyashita, J.N. Onuchic, and P.G. Wolynes, Proc.
Natl. Acad. Sci. USA 100, 12570 (2003).
[19] K. Prakash, S. Prajapati, A. Ahmad, S.K. Jain, and
V. Bhakuni, Protein Sci. 11, 46 (2002); R. Fernandez-
Lafuente, Enzyme Microb. Technol. 45, 405 (2009).
[20] This form is strictly speaking only correct for shapes with
axial symmetry about the axis of deformation. For more
realistic shapes, αwill be a tensor. Introducing this level
of description is an unnecessary complication at this level.
[21] For a simple model of two spheres of radii a1and a2, we
have α= (a1a2)/(a1+a2). In our order of magnitude
estimates, we use an upper limit value of |α|= 1.
[22] I. Newton, Philos. Trans. R. Soc. London 22, 824829
(1701).
[23] Ya.B. Zeldovich, G.I. Barenblatt, V.B. Librovich, G.M.
Makhviladze, The Mathematical Theory of Combustion
and Explosions (Consultants Bureau, Plenum Press, New
York, 1985).
[24] I.W. Sizer, J. Biol. Chem. 154, 461 (1944).
[25] G.N. Mercer, R.O. Weber, B.F. Gray, and S.D. Watt,
Mathl. Comput. Modelling 24, 29-38 (1996).
[26] R. Golestanian, Phys. Rev. Lett. 108, 038303 (2012).
... The idea that some enzymes diffuse faster than expected from Brownian motion in the presence of their substrates has been a topic of debate in recent literature. It has been observed in some studies [1][2][3] and denied by others 4,5 . Most experiments in which this phenomenon, termed enhanced enzyme diffusion (EED), has been observed, were conducted using fluorescence correlation spectroscopy (FCS)an ensemble averaging technique requiring fluorescent labelling. ...
Preprint
Full-text available
The existence of the phenomenon of enhanced enzyme diffusion (EED) has been a topic of debate in recent literature. The majority of experiments confirming the existence of this phenomenon have been conducted using fluorescence correlation spectroscopy (FCS), but an artefact has been found in some of these experiments. There are various proposed mechanisms to explain the origin of EED, such as conformational changes and oligomeric enzyme dissociation. In our study we use mass photometry (MP), a label-free single-molecule interferometric light scattering technique, to investigate the dependence of the oligomeric states of several enzymes on the presence or absence of substrate. The enzymes of interest in this study are catalase, aldolase, alkaline phosphatase and an alcohol oxidase (VAO) the first three of which were previously studied in the context of enhanced enzyme diffusion. We compared the ratios of oligomeric states in the presence and absence of substrate as well as different substrate and inhibitor concentrations. Catalase and aldolase were found to dissociate into smaller oligomers with substrate, while for alkaline phosphatase and VAO, different behaviours were observed. Intriguingly, the changes in oligomeric states in aldolase and catalase are independent of catalysis. Thus, we have identified a possible mechanism which explains the previously observed enhanced diffusion of enzymes by oligomer dissociation through ligand binding.
... Additionally, the catalytic activity of the enzymes may be associated to conformational changes or oscillations in the enzyme shape [15,16]. The effect of such conformational changes on the spatial dynamics and the rheology of enzyme-rich solutions has been a topic of great recent interest [17][18][19][20]. In this context, a new mechanism for synchronization between two enzymes was recently reported, for enzymes that undergo conformational changes during their noise-activated catalytic steps, and which are coupled to each other through a viscous medium [21]. ...
Article
Full-text available
We study the stochastic dynamics of an arbitrary number of noise-activated cyclic processes, or oscillators, that are all coupled to each other via a dissipative coupling. The N coupled oscillators are described by N phase coordinates driven in a tilted washboard potential. At low N and strong coupling, we find synchronization as well as an enhancement in the average speed of the oscillators. In the large N regime, we show that the collective dynamics can be described through a mean-field theory, which predicts a great enhancement in the average speed. In fact, beyond a critical value of the coupling strength, noise activation becomes irrelevant and the dynamics switch to an effectively deterministic ``running'' mode. Finally, we study the stochastic thermodynamics of the coupled oscillators, in particular their performance with regards to the thermodynamic uncertainty relation.
... In the limits of infinitely small swimmer sizes (L → 0) and for the symmetric case (α = 1/2), the model used in the present work reduces to the standard force-dipole model that has been used widely to describe the dynamics of enzymatic molecules [79,81,96]. In this regime, it would be of interest to study the effects of the odd viscosity on various enzymatic transport phenomena, such as diffusion enhancement [96][97][98][99], chemotactic or anti-chemotactic behavior [80,100,101], and synchronization effects [102]. For instance, the non-reciprocal linear response due to odd viscosity can lead to additional correction terms in the diffusion coefficient of passive tracers in solutions and biological membranes [79]. ...
Article
Full-text available
We theoretically and computationally study the low-Reynolds-number hydrodynamics of a linear active microswimmer surfing on a compressible thin fluid layer characterized by an odd viscosity. Since the underlying three-dimensional fluid is assumed to be very thin compared to any lateral size of the fluid layer, the model is effectively two-dimensional. In the limit of small odd viscosity compared to the even viscosities of the fluid layer, we obtain analytical expressions for the self-induced flow field, which includes non-reciprocal components due to the odd viscosity. On this basis, we fully analyze the behavior of a single linear swimmer, finding that it follows a circular path, the radius of which is, to leading order, inversely proportional to the magnitude of the odd viscosity. In addition, we show that a pair of swimmers exhibits a wealth of two-body dynamics that depends on the initial relative orientation angles as well as on the propulsion mechanism adopted by each swimmer. In particular, the pusher-pusher and pusher-puller-type swimmer pairs exhibit a generic spiral motion, while the puller-puller pair is found to either co-rotate in the steady state along a circular trajectory or exhibit a more complex chaotic behavior resulting from the interplay between hydrodynamic and steric interactions. Our theoretical predictions may pave the way toward a better understanding of active transport in active chiral fluids with odd viscosity, and may find potential applications in the quantitative microrheological characterization of odd-viscous fluids.
... Phoresis is a movement driven by a gradient of some material property, e.g., of the electric field in the case of electrophoresis, temperature in the case of thermophoresis, and the concentration in the case of diffusiophoresis. Within this mechanism, an enzymatic reaction is the source of such gradient itself, so one refers to the effects as a self-electrophoresis [205], self-thermophoresis [219], and self-diffusiophoresis [196,220], respectively. In turn, conformational changes include size decrease [202,203], active swimming [216,221], damping conformational fluctuations [209], and chemoacoustic effects [208]. ...
Thesis
Full-text available
Whole-cell simulations have been identified as the grand challenge of XXIst-century. Since even in the exascale computing era, first-principle atomistic whole-cell simulations are computationally prohibitive, coarse-grained reaction-diffusion models appear as essential and feasible alternatives. In such models, it is crucial to adequately describe the diffusion and reaction properties of all biologically-relevant molecules. However, as it has been known for decades, the cell interior is crowded with macromolecules occupying 20 to 40% of the cell volume. Clearly, diffusion and reactions under such crowding conditions are not the same as in test tubes. To understand these differences, in this Thesis, we explore the generic effects of how crowding affects macromolecular diffusion, chemical equilibria, and reaction kinetics.
... In the limits of infinitely small swimmer sizes (L → 0) and for the symmetric case (α = 1/2), the model used in the present work reduces to the standard force-dipole model that has been used widely to describe the dynamics of enzymatic molecules 77,79,87 . In this regime, it would be of interest to study the effects of the odd viscosity on various enzymatic transport phenomena, such as diffusion enhancement [87][88][89][90] , chemotactic or anti-chemotactic behavior 78,91,92 , and synchronization effects 93 . For instance, the non-reciprocal linear response due to odd viscosity can lead to additional correction terms in the diffusion coefficient of passive tracers in solutions and biological membranes 77 . ...
Preprint
We theoretically and computationally study the low-Reynolds-number hydrodynamics of a linear active microswimmer surfing on a compressible thin fluid layer characterized by an odd viscosity. In the limit of small odd viscosity compared to the even viscosities of the fluid layer, we obtain analytical expressions for the self-induced flow field, which includes non-reciprocal components due to the odd viscosity. On this basis, we fully analyze the behavior of a single linear swimmer, finding that it follows a circular path, the radius of which is, to leading order, inversely proportional to the magnitude of the odd viscosity. In addition, we show that a pair of swimmers exhibits a wealth of two-body dynamics that depends on the initial relative orientation angles as well as on the propulsion mechanism adopted by each swimmer. In particular, the pusher-pusher and pusher-puller-type swimmer pairs exhibit a generic spiral motion, while the puller-puller pair is found to either co-rotate in the steady state along a circular trajectory or exhibit a more complex chaotic behavior resulting from the interplay between hydrodynamic and steric interactions. Our theoretical predictions may pave the way toward a better understanding of active transport in active chiral fluids with odd viscosity, and may find potential applications in the quantitative microrheological characterization of odd-viscous fluids.
... In cases where the integral heat exchange structure is used to carry out catalytic partial combustion of a fuel followed by complete combustion after the catalyst, the catalyst must burn a portion of the fuel and produce an outlet gas sufficiently hot to induce homogeneous combustion after the catalyst [49,50]. In addition, it is desirable that the catalyst not become too hot since this would shorten the life of the catalyst and limit the advantages to be gained from this approach. ...
Preprint
In modern industrial practice, a variety of highly exothermic reactions are promoted by contacting of the reaction mixture in the gaseous or vapor phase with a heterogeneous catalyst. A need exists for improved catalytic structures employing integral heat exchange which will substantially widen the window or range of operating conditions under which such catalytic structures can be employed in highly exothermic processes like catalytic combustion or partial combustion. The highly exothermic process characteristics of catalytic reactors are investigated with integral heat exchange structures. Ethane mole fraction and gas-phase reaction rate profiles in catalytic reactors are presented, and ethane mole fraction, flow velocity, gas-phase reaction rate, and temperature contour plots are illustrated for catalytically supported thermal combustion systems. The present study aims to provide an improved reaction system and process for combustion of a fuel wherein catalytic combustion using a catalyst structure employing integral heat exchange affords a partially-combusted, gaseous product which is passed to a homogeneous combustion zone where complete combustion is promoted by means of a flame holder. Particular emphasis is placed upon the catalytic reactor configuration that allows the oxidation catalyst to be backside cooled by any fluid passing through the cooling conduits. The results indicate that the percentage of reaction completed in the exothermic catalytic reaction channel depends both upon the flow rate of the fuel-oxidant mixture through the exothermic catalytic reaction channel and upon the physical characteristics of the catalytic reactor. The tortuosity of the catalytic channels is increased by changing their cross-sectional area at a multiplicity of points along their longitudinal axes. The gas flow velocity entering the exothermic catalytic reaction channel should exceed the minimum required to prevent flashback into the fuel-oxidant stream upstream of the reactor if the fuel-oxidant mixture entering the exothermic catalytic reaction channel is within the limits of flammability. Catalytically-supported thermal combustion in the catalytic reactor is achieved by contacting at least a portion of the carbonaceous fuel intimately admixed with air with a solid oxidation catalyst having an operating temperature substantially above the instantaneous auto-ignition temperature of the fuel-air admixture. The film heat transfer coefficient provides useful means of characterizing the different flow geometries provided by the various flow channel configurations which distinguish the catalyst-coated channels from the catalyst-free channels of the catalyst structure. The total residence time in the combustion system should be sufficient to provide essentially complete combustion of the fuel, but not so long as to result in the formation of oxides of nitrogen. Keywords: Catalytic reactors; Physical characteristics; Exothermic reactions; Heterogeneous catalysts; High temperatures; Thermal combustion
... Note that the effect of fluctuations on the diffusion of tracer particles is quite generally weak, also emphasizing that local equilibrium can hold in a broad range of parameters. Our approach can be applied to the diffusion of particles in an active environment, which is different from but related to the problem of the diffusion of active enzymes [43][44][45]. ...
Article
Full-text available
Chemically active systems such as living cells are maintained out of thermal equilibrium due to chemical events which generate heat and lead to active fluctuations. A key question is to understand on which time and length scales active fluctuations dominate thermal fluctuations. Here, we formulate a stochastic field theory with Poisson white noise to describe the heat fluctuations which are generated by stochastic chemical events and lead to active temperature fluctuations. We find that on large length- and timescales, active fluctuations always dominate thermal fluctuations. However, at intermediate length- and timescales, multiple crossovers exist which highlight the different characteristics of active and thermal fluctuations. Our work provides a framework to characterize fluctuations in active systems and reveals that local equilibrium holds at certain length- and timescales.
Article
Macromolecular crowding affects the activity of proteins and functional macromolecular complexes in all cells, including bacteria. Crowding, together with physicochemical parameters such as pH, ionic strength, and the energy status, influences the structure of the cytoplasm and thereby indirectly macromolecular function. Notably, crowding also promotes the formation of biomolecular condensates by phase separation, initially identified in eukaryotic cells but more recently discovered to play key functions in bacteria. Bacterial cells require a variety of mechanisms to maintain physicochemical homeostasis, in particular in environments with fluctuating conditions, and the formation of biomolecular condensates is emerging as one such mechanism. In this work, we connect physicochemical homeostasis and macromolecular crowding with the formation and function of biomolecular condensates in the bacterial cell and compare the supramolecular structures found in bacteria with those of eukaryotic cells. We focus on the effects of crowding and phase separation on the control of bacterial chromosome replication, segregation, and cell division, and we discuss the contribution of biomolecular condensates to bacterial cell fitness and adaptation to environmental stress.
Article
Recent experiments have shown that the diffusion of reagent molecules is inconsistent with what the Stokes-Einstein equation predicts during a chemical reaction. Here, we used single-molecule tracking to observe the diffusion of reactive reagent molecules during click and Diels-Alder (DA) reactions. We found that the diffusion coefficient of the reagents remained unchanged within the experimental uncertainty upon the DA reaction. Yet, diffusion of reagent molecules is faster than predicted during the click reaction when the reagent concentration and catalyst concentration exceed a threshold. A stepwise analysis suggested that the fast diffusion scenario is due to the reaction but not the involvement of the tracer with the reaction itself. The present results provide experimental evidence on the faster-than-expected reagent diffusion during a CuAAC reaction in specific conditions and propose new insights into understanding this unexpected behavior.
Article
Full-text available
Recent studies have shown that the diffusivity of enzymes increases in a substrate-dependent manner during catalysis. Although this observation has been reported and characterized for several different systems, the precise origin of this phenomenon is unknown. Calorimetric methods are often used to determine enthalpies from enzyme-catalysed reactions and can therefore provide important insight into their reaction mechanisms. The ensemble averages involved in traditional bulk calorimetry cannot probe the transient effects that the energy exchanged in a reaction may have on the catalyst. Here we obtain single-molecule fluorescence correlation spectroscopy data and analyse them within the framework of a stochastic theory to demonstrate a mechanistic link between the enhanced diffusion of a single enzyme molecule and the heat released in the reaction. We propose that the heat released during catalysis generates an asymmetric pressure wave that results in a differential stress at the protein-solvent interface that transiently displaces the centre-of-mass of the enzyme (chemoacoustic effect). This novel perspective on how enzymes respond to the energy released during catalysis suggests a possible effect of the heat of reaction on the structural integrity and internal degrees of freedom of the enzyme.
Article
Full-text available
This review summarizes theoretical progress in the field of active matter, placing it in the context of recent experiments. This approach offers a unified framework for the mechanical and statistical properties of living matter: biofilaments and molecular motors in vitro or in vivo, collections of motile microorganisms, animal flocks, and chemical or mechanical imitations. A major goal of this review is to integrate several approaches proposed in the literature, from semimicroscopic to phenomenological. In particular, first considered are “dry” systems, defined as those where momentum is not conserved due to friction with a substrate or an embedding porous medium. The differences and similarities between two types of orientationally ordered states, the nematic and the polar, are clarified. Next, the active hydrodynamics of suspensions or “wet” systems is discussed and the relation with and difference from the dry case, as well as various large-scale instabilities of these nonequilibrium states of matter, are highlighted. Further highlighted are various large-scale instabilities of these nonequilibrium states of matter. Various semimicroscopic derivations of the continuum theory are discussed and connected, highlighting the unifying and generic nature of the continuum model. Throughout the review, the experimental relevance of these theories for describing bacterial swarms and suspensions, the cytoskeleton of living cells, and vibrated granular material is discussed. Promising extensions toward greater realism in specific contexts from cell biology to animal behavior are suggested, and remarks are given on some exotic active-matter analogs. Last, the outlook for a quantitative understanding of active matter, through the interplay of detailed theory with controlled experiments on simplified systems, with living or artificial constituents, is summarized.
Article
Full-text available
With remarkable precision and reproducibility, cells orchestrate the cooperative action of thousands of nanometre-sized molecular motors to carry out mechanical tasks at much larger length scales, such as cell motility, division and replication. Besides their biological importance, such inherently non-equilibrium processes suggest approaches for developing biomimetic active materials from microscopic components that consume energy to generate continuous motion. Being actively driven, these materials are not constrained by the laws of equilibrium statistical mechanics and can thus exhibit sought-after properties such as autonomous motility, internally generated flows and self-organized beating. Here, starting from extensile microtubule bundles, we hierarchically assemble far-from-equilibrium analogues of conventional polymer gels, liquid crystals and emulsions. At high enough concentration, the microtubules form a percolating active network characterized by internally driven chaotic flows, hydrodynamic instabilities, enhanced transport and fluid mixing. When confined to emulsion droplets, three-dimensional networks spontaneously adsorb onto the droplet surfaces to produce highly active two-dimensional nematic liquid crystals whose streaming flows are controlled by internally generated fractures and self-healing, as well as unbinding and annihilation of oppositely charged disclination defects. The resulting active emulsions exhibit unexpected properties, such as autonomous motility, which are not observed in their passive analogues. Taken together, these observations exemplify how assemblages of animate microscopic objects exhibit collective biomimetic properties that are very different from those found in materials assembled from inanimate building blocks, challenging us to develop a theoretical framework that would allow for a systematic engineering of their far-from-equilibrium material properties.
Book
This monograph contains 7 chapters. Topics discussed include basic physical chemistry and thermodynamics of combustion; chemical reactions and critical phenomena, and laminar flames; time-independent theory of thermal explosions; initiation of chemical reaction waves in fuel mixtures; time-dependency and ignition; laminar flames: equations, steady-state solutions, flame propagation, velocities, fronts, stability, and self-ignition; complex and chain reactions in flames: multi-step, unbranched, branched reactions, and cool flames; gas dynamics of combustion: stationary and closed-vessel combustion, flame acceleration and detonation in tubes, and hydrodynamic instability of flames; and diffusional combustion of gases. Refs.
Article
The moderate stability of enzymes is one of the main drawbacks that hinder general implementation of these interesting biocatalysts at industrial scale. An especially complex problem is the stabilization of multimeric proteins, where dissociation of the subunits produces enzyme inactivation and even product contamination. In this review, different strategies to stabilize multimeric enzymes at different levels are revised. First, the use of proper experimental conditions may facilitate the handling of the enzymes (ions, polymers, etc.). Second, genetic tools may be used to crosslink (via disulfide bonds) or just to reinforce the subunit–subunit interactions. The physical or chemical crosslinking of the enzyme subunits will be also discussed. Finally, the use of immobilization strategies (with or without pre-existing supports) will be discussed. Special emphasis will be put on the new immobilization strategies specifically designed to involve the maximum amount of enzyme subunits in the immobilization (and thus, in the further multipoint covalent attachment).
Article
Using fluorescence correlation spectroscopy, we show that the diffusive movement of catalase enzyme molecules increase in the presence of the substrate, hydrogen peroxide, in a concentration-dependent manner. Employing a microfluidic device to generate a substrate concentration gradient, we show that both catalase and urease enzyme molecules spread towards areas of higher substrate concentration, a form of chemotaxis at the molecular scale. Using glucose oxidase and glucose to generate a hydrogen peroxide gradient, we induce the migration of catalase towards glucose oxidase thereby showing that chemically interconnected enzymes can be drawn together.
Article
A model for the combustion of a reactant with heat loss in an infinite two-dimensional layer or an infinitely long circular cylinder is derived. Using center manifold techniques the model is shown to reduce to a one-dimensional model on an infinite region for both geometries. For an exothermic first-order reaction with Arrhenius temperature dependence, a numerical method is used to calculate solutions and comparisons are made with the no-reactant consumption case. Traveling pseudo-waves are shown to exist and their speeds determined. A first-order estimate of the reaction zone thickness, obtained by perturbation methods, is shown to be in excellent agreement with the full numerical solution.
Article
Spontaneous collective motion, as in some flocks of bird and schools of fish, is an example of an emergent phenomenon. Such phenomena are at present of great interest and physicists have put forward a number of theoretical results that so far lack experimental verification. In animal behaviour studies, large-scale data collection is now technologically possible, but data are still scarce and arise from observations rather than controlled experiments. Multicellular biological systems, such as bacterial colonies or tissues, allow more control, but may have many hidden variables and interactions, hindering proper tests of theoretical ideas. However, in systems on the subcellular scale such tests may be possible, particularly in in vitro experiments with only few purified components. Motility assays, in which protein filaments are driven by molecular motors grafted to a substrate in the presence of ATP, can show collective motion for high densities of motors and attached filaments. This was demonstrated recently for the actomyosin system, but a complete understanding of the mechanisms at work is still lacking. Here we report experiments in which microtubules are propelled by surface-bound dyneins. In this system it is possible to study the local interaction: we find that colliding microtubules align with each other with high probability. At high densities, this alignment results in self-organization of the microtubules, which are on average 15 µm long, into vortices with diameters of around 400 µm. Inside the vortices, the microtubules circulate both clockwise and anticlockwise. On longer timescales, the vortices form a lattice structure. The emergence of these structures, as verified by a mathematical model, is the result of the smooth, reptation-like motion of single microtubules in combination with local interactions (the nematic alignment due to collisions)--there is no need for long-range interactions. Apart from its potential relevance to cortical arrays in plant cells and other biological situations, our study provides evidence for the existence of previously unsuspected universality classes of collective motion phenomena.