You are on page 1of 28

Reports on Progress in Physics

REVIEW
Recent citations
Bacterial growth: a statistical physicist’s guide - Growth-dependent drug susceptibility can
prevent or enhance spatial expansion of a
To cite this article: Rosalind J Allen and Bartlomiej Waclaw 2019 Rep. Prog. Phys. 82 016601 bacterial population
Patrick Sinclair et al

View the article online for updates and enhancements.

This content was downloaded from IP address 190.237.29.125 on 28/07/2020 at 18:04


Reports on Progress in Physics

Rep. Prog. Phys. 82 (2019) 016601 (27pp) https://doi.org/10.1088/1361-6633/aae546

Review

Bacterial growth: a statistical physicist’s


guide
Rosalind J Allen and Bartlomiej Waclaw
School of Physics and Astronomy, The University of Edinburgh, James Clerk Maxwell Building,
Peter Guthrie Tait Road, Edinburgh EH9 3FD, United Kingdom

E-mail: rosalind.allen@ed.ac.uk

Received 13 November 2017, revised 24 August 2018


Accepted for publication 1 October 2018
Published 14 November 2018

Corresponding Editor Professor Mark Geoghegan

Abstract
Bacterial growth presents many beautiful phenomena that pose new theoretical challenges
to statistical physicists, and are also amenable to laboratory experimentation. This review
provides some of the essential biological background, discusses recent applications of
statistical physics in this field, and highlights the potential for future research.

Keywords: statistical physics, bacteria, population dynamics, microbiology

(Some figures may appear in colour only in the online journal)

1. Introduction obviously belong to this class, and statistical physics has a


long history of applications to biological problems. Examples
Consider the following scenario: a small number of patho- include determination of mutation rates by analysis of mutant
genic bacteria (perhaps 10–100) enter the human body and number statistics [9, 10], the totally asymmetric exclusion pro-
cause an infection. An antibiotic is prescribed to fight off the cess [11], which was originally proposed as a model for pro-
infection. Assuming that the bacterial infection is initially tein production from messenger RNA in biological cells [12],
sensitive to the antibiotic, what are the chances of curing models for noise in gene regulation [13, 14], lattice models
the infection, and how likely is it that the infection eventu- of growing populations [15–17], models for collective flock-
ally becomes resistant to the antibiotic? Given the increasing ing and swarming behaviour [18–20] and non-equilibrium
global health issue posed by antibiotic resistant infections, phase transitions in populations of self-propelled ‘swimmers’
this is an important and timely problem [1–5]. Clearly, under- [21–25].
standing the growth of bacterial infections and the potential In this review we argue that the dynamics of growing bac-
for emergence and spread of antibiotic resistance within them terial populations provides another class of systems to which
requires collaboration between scientists from different dis- the methods of statistical physics can naturally be applied. To
ciplines. Do statistical physicists have a role to play here? briefly illustrate this, we notice that the above example of the
Furthermore, thinking more broadly, could our understand- growth of an antibiotic-resistant infection involves stochastic
ing of other processes mediated by bacterial growth, such as phenomena on scales ranging from macroscopic to molecular.
global biogeochemical cycles [6], human gut health [7], and Specifically:
wastewater treatment [8], also profit from a statistical physics- Macroscopic level: population expansion in space-In many
like approach? real-world scenarios, including infections, bacterial popula-
Statistical physicists find inspiration in systems where tions spread in space (e.g. through an infected tissue). This
complex macroscopic behaviour arises from a simple set process could be modelled using the Fisher–Kolmogorov
of underlying microscopic dynamical rules. Living systems equation,

1361-6633/19/016601+27$33.00 1 © 2018 IOP Publishing Ltd  Printed in the UK


Rep. Prog. Phys. 82 (2019) 016601 Review

∂n(x, t) ∂ 2 n(x, t) is coarse-grained enough to provide useful insight, but takes


(1) =D + rn(x, t)(1 − n(x, t)/K), account of the essential biology, allowing it to give useful pre-
∂t ∂x2
dictions. In many cases, the ‘right’ physical or mathematical
where n(x, t) is the population density of bacteria in space and model of a biological system depends on the question that one
time, r is the maximal growth rate, D is a diffusion constant is trying to answer.
that accounts for bacterial motility and K represents a maxi- We begin by introducing the reader to some basic microbi-
mal population density. Section 4 will discuss applications of ology, and to some interesting collective phenomena exhibited
this equation, and other approaches, to bacterial populations by bacterial populations. We do not aim at a comprehensive
growing in heterogeneous environments. introduction, but rather we try to provide just enough informa-
Microscopic level: bacterial replication processes the tion to follow the topics discussed later in the review—more
growth dynamics of a population containing a mixture of detailed background material is available in excellent textbooks
antibiotic-sensitive and antibiotic-resistant bacterial cells, [26–28]. The remainder of this review is devoted to a more
which replicate stochastically with rates rS and rR, could be detailed description of bacterial growth phenomena which
described by the following simple master equation: present interesting challenges for statistical physicists, and
dP(S, R) examples of how statistical physics has been applied to these
= rR (R − 1)P(S, R − 1) + rS (S − 1) problems. This is divided into two parts, which cover growth
dt
× P(S − 1, R) − (rR + rS )P(S, R), (2) phenomena in well-mixed systems and in spatially heteroge-
 neous systems, respectively. Finally, we present our own per-
where S and R are the numbers of sensitive and resistant cells. spectives on the potential of this field, and on the relationship
This type of equation can be solved using methods developed between statistical physics and microbiology. For lack of space,
by statistical physicists to study processes such as random we do not consider in this review the fascinating and important
walks, birth–death processes and coalescence processes, as topic of bacterial evo­lution, where statistical physicists have
we shall discuss in section 3. also made major contrib­utions (for example to understanding
Molecular level: gene expression resistance to an antibi- the structure of fitness landscapes). For this topic, we refer the
otic can be caused by genetic changes in the bacterial DNA reader to the excellent review by De Visser and Krug [29].
(mutations), or by changes in how the bacterium expresses
its genes. To express a gene, the DNA sequence is first tran- 2. Background
scribed, or copied, into an mRNA transcript molecule, which
is then translated into protein (see section 2.1). To model this In this section we give a very brief introduction to the basics
process we can use a set of Langevin equations: of bacteria and their growth. We also introduce the reader to
dx/dt = kx − γx x + ηx (t),
(3) several different types of stochastic collective behaviours that
are exhibited by bacterial populations. Table 1 contains useful
dX/dt = kX x − γX X + ηX (t),
(4) numbers that relate to some of the topics that we discuss in
the text.
where x and X are the concentration of the mRNA and pro-
tein respectively, kx and kX are the transcription and trans-
lation rates, γx and γX are decay rates, and ηx (t) and ηX (t) 2.1.  Basic microbiology for statistical physicists
represent Gaussian noise. These equations are similar to those
From a statistical physicist’s point of view, a bacterium can
encounter­ed in other statistical physics problems.
be viewed as a microscopic particle, or cell, which is bounded
In this review, we discuss how statistical physics models by a pair of membranes with a stiff wall in between them
can be applied to problems in bacterial population dynamics. (specifically, this is the case for a large class of bacteria that
The purpose of this review is to encourage interest in these are known as Gram negatives; Gram positive bacteria have a
problems, and to provide some of the basic biological back- thicker wall and lack the outer membrane). The interior of the
ground that is needed to appreciate the field. Physics and biol- bacterial cell contains a ‘soup’ of DNA (encoding the bacterial
ogy are of course very different in their language, philosophy, genome), RNA, proteins, and other molecules (figure 1). The
background and culture, and full immersion into the world materials that make up the bacterium are generically referred
of bacteria comes with considerable challenges. Nevertheless, to as ‘biomass’. Bacterial cells come in different shapes: from
we hope to show here that bacterial population growth phe- rods, to spheres, to spirals (figure 1), and sizes: from  ∼100 nm
nomena can provide considerable inspiration for the develop- to  ∼100 μm. Escherichia coli, the ‘workhorse’ of the micro-
ment of new and interesting statistical physics models. biology lab, is a spherocylindrical Gram negative bacterium
All the models that are discussed in this review are ide- whose cells are  ∼0.8–1 μm in diameter and  ∼2–4 μm in
alized and abstract descriptions of complex biological pro- length [26, 32].
cesses. It is often necessary to formulate coarse-grained Bacterial growth consists of the conversion of chemical nutri-
models for biological systems, because many of the under- ents into biomass. Nutrients enter the bacterium through pores in
lying details (e.g. interactions or rate constants) are simply its membrane and undergo a series of chemical transformations,
not known. The most difficult aspect of the problem may not converting them into new cellular components; these chemical
be how to solve the model, but how to formulate it so that it transformations are collectively known as metabolism [26, 27].

2
Rep. Prog. Phys. 82 (2019) 016601 Review

(a)

(b) (d)

(c)

Figure 1.  (a) Schematic illustration of the structure of a typical Gram-negative bacterium. Reproduced with permission from Todar’s
Online Textbook of Bacteriology. (b) Scanning electron micrograph of E. coli cells on a silicon surface. Reproduced with permission from
[37]. © 2010 American Society for Microbiology. (c) Scanning electron micrograph of Staphylococcus aureus cells. Reproduced with
permission from [38]. © 1972 Pathological Society of Great Britain and Ireland. Published by the Microbiology Society. (d) Scanning
electron micrograph of Treponema pallidum cells (the causative agent of syphilis), adhering to a human brain epithelial cell. [29] John
Wiley & Sons. © 2016 European Academy of Dermatology and Venereology. In panels (b)–(d), the scale bars represent 2 μm.

The increase in biomass is accompanied by an increase in cell or signals, a bacterial cell can turn on or off the production
size and by replication of the bacterial DNA, possibly with some of particular protein molecules; this is known as gene regu-
errors (mutations). Eventually, the cell divides into two daughter lation. Interestingly, some proteins, known as transcription
cells, in a process called binary fission. The cell size at which factors, turn on or off genes that encode other proteins. This
division occurs is dependent on the growth conditions [40–42] leads to networks of interactions among genes, the proper-
(with cells growing on richer nutrients being larger). However, ties of which (such as modularity [48]) have attracted signif-
the exact mechanism by which cell division is triggered remains icant interest among statistical physicists. Gene regulatory
somewhat mysterious, even after half a century of research networks are especially interesting because the transcrip-
[43–47]. Bacteria are able to reproduce at impressive rates: E. tion factors that control them are often present in only a few
coli can double its population every 20 min, under optimal condi- molecules per cell, leading to stochasticity in the behaviour
tions. This means that very large population sizes can be achieved of the regulatory network (e.g. switching between alterna-
within a few hours in the lab; population densities of  ∼109 cells tive stable states [49, 50]; for theor­etical models see, e.g.
per ml of culture medium are usual in lab experiments. [51–56]).
Protein molecules make up a major component of bio- Many bacterial cells can also engage in self-propelled
mass: typically, ∼55% of the dry mass of a bacterial cell con- motion, which is mediated by various appendages external to
sists of protein [27]. The bacterial DNA sequence contains the cell. For example, bacteria may swim in liquid media by
several thousand genes (∼4000 for E. coli), each of which rotation of whip-like flagella, or crawl on solid surfaces using
encodes a specific protein molecule. Gene expression is the needle-like appendages called pili [27, 28, 57–59]. Bacterial
process by which the DNA-encoded instructions for mak- motility has already attracted much interest among physicists;
ing a particular protein are first transcribed into a messenger topics of particular focus have included the statistics of sus-
RNA molecule, which is then translated, i.e. used to build a pensions in which bacteria stochastically change direction in
chain of amino acid molecules that folds into a protein mol- response to chemical gradients or local density [25, 60] and the
ecule. In response to changes in environmental conditions, hydrodynamics of bacterial swimming motion [61]. Bacterial

3
Rep. Prog. Phys. 82 (2019) 016601 Review

Table 1.  Useful numbers for modelling bacterial growth and evolution. All data refers to the bacterium E. coli. If no reference is given, the
values come from in-house experiments for the MG1655 strain of E. coli. We also refer the reader to bionumbers [36]—an excellent source
of biology-related numbers.
Parameter Value
Size Typical width 1 μm, typical length 2–5 μm
Exponential growth rate Maximum:  ∼2 h−1, sub-optimal conditions: 0.3–1.5 h−1
Minimum doubling time ∼20 min
Elongation rate 0.1–0.2 μm min−1 (on rich medium)
Maximum density ∼1–5  ×  109 cells per ml (LB medium, stationary phase)
Mutation rate ∼2 × 10−10 per bp per replication [30]
Glucose molecules consumed to make 1 cell ∼1.8 × 1010 [26, 31]
Weight 280 fg per cell [32]
Protein molecules per cell 2.35 × 106 (1850 distinct protein molecules) [32]
mRNA molecules per cell 1380 [32]
Genome size 4.5 × 106 bp
Genome copy number 1 (slow growth) to 8 (fast growth) [33]
Abundance of RNA polymerase ∼1% of total protein mass [34]
Abundance of ribosomes (growth rate dependent) ∼20%–40% of total mass [32];  ∼7000–70 000 per cell
DNA replication rate 580–1190 bp s−1 [34]
mRNA elongation rate (transcription) 39–56 nucleotides s−1 [34]
Peptide elongation rate (translation) 13–22 amino acids s−1 [34]
Intracellular concentration of ATP (growth in glucose-fed chemostat) 9.6 mM [35]
Intracellular concentration of a typical metabolite 0.1–100 mM [35]
Total intracellular metabolite concentration ∼300 mM [35]
Plasmid size ∼2–500 kbp
Plasmid copy number ∼1–200 per cell
Minimal inhibitory concentration,
– Ampicillin (inhibits cell wall synthesis) ∼8 μg ml−1 (LB medium)
– Rifampicin (RNA synthesis inhibitor) ∼3 μg ml−1 (LB medium)
– Ciprofloxacin (DNA gyrase inhibitor) ∼20 ng ml−1 (LB medium)

motion has also inspired a recent surge of work on the collec- an expanding population into sectors of genetically identical
tive behaviour of self-propelled colloidal particles [24, 62–64]. bacteria [72–74], quasi-nematic ordering [75] and competi-
tion for space between lineages [76]. Mechanical interac-
tions between bacteria and their environment can also lead to
2.2.  Statistical physics of bacterial growth
interesting effects [77], for example a transition from 2d to
Bacterial growth is of interest to statistical physicists for 3d growth as a bacterial colony grows on a semi-solid agar
several reasons. First, the process of division into daughter gel [78].
cells is a branching process with somewhat stochastic timing; Third, because bacteria reproduce rapidly, they also
the time between successive bacterial divisions is a random undergo rapid genetic evolution. The process of evolution
variable with a rather broad distribution [45, 47, 65, 66]. For involves the random generation of cells with mutations in their
E. coli, this causes a loss of synchrony between division events DNA, due to mistakes in DNA replication, and their prolifera-
in sister lineages within a few generations [67, 68]. Stresses tion within the population, starting from initially very small
such as exposure to some antibiotics or to ultraviolet radiation numbers. Bacterial evolution is now widely recognised as an
can interfere with the division process, leading to long, fila- important testbed for evolutionary theory, since it allows lab
mentous cells. Even in the absence of stress, bacterial popula- experiments to be carried out on short timescales (typically
tions can contain small sub-populations of non-growing cells, days-weeks) [79, 80]. Understanding how bacterial popula-
or ‘persisters’, which tend to be resistant to antibiotic treat- tions evolve is also a pre-requisite for our ability to mitigate
ment [69, 70]. against the emergence of antibiotic-resistant infections [81].
Second, growth of bacteria in close proximity to one In the remainder of this review, we highlight in more
another leads to mechanical interactions, which can be detail a number of interesting phenomena that are asso-
thought of as pushing, or excluded volume effects. This is rel- ciated with various modes of bacterial growth, and for
evant when bacteria grow in dense populations such as colo- which statistical physics models have been developed. We
nies on semi-solid surfaces or biofilms on solid surfaces (see divide this discussion into two parts: in section 3 we con-
section 4). Mechanical interactions between bacteria lead to a sider growth in a homogeneous, well-mixed environment,
number of interesting phenomena, including phase separation while in section 4 we discuss growth in spatially structured
of cells with different surface properties [71], segregation of environments.

4
Rep. Prog. Phys. 82 (2019) 016601 Review

2
3.  Bacterial growth in a spatially homogeneous as discrete events which may be synchronous . In some cases,
environment it may be more convenient to use as the dynamical variable
the total biomass of the population rather than the number N
3.1.  Bacterial growth experiments and population of bacteria. The total biomass obeys an identical equation to
dynamic equations equation  (5), but it is a continuous quantity which is unaf-
fected by discrete cell division events [85].
In the laboratory, bacteria are often grown in liquid suspen-
Equation (5) provides a good model for the exponential phase
sion under well-mixed conditions. Here we give a brief over-
of growth of a bacterial population, as we show in figure 2(E).
view of the typical experimental techniques involved and the
However, it does not capture the transition to the stationary
types of equations  used to describe the resulting population
phase, where the population saturates. A simple way to capture
dynamics.
this saturation is to use instead a logistic growth equation [86]
3.1.1. Growth in a batch culture.  Figure 2(A) illustrates a dN
= rN(1 − N/K),
(7)
typical setup for what is known as a ‘batch culture’ growth dt
experiment. A small number of bacteria are inoculated into where K is the maximal population size (or carrying capac-
a well-shaken container filled with liquid nutrient medium ity) and the term (1 − N/K) decreases the effective growth
(figure 2(B) shows a large-volume flask; figure  2(C) shows rate when N becomes large, mimicking the effect of nutri-
a 96-well microplate which can be used to perform multiple ent depletion or toxic waste product buildup. The solution of
simultaneous smaller-volume experiments). Over a period equation  (7), N(t) = N(0)ert /[1 + (N(0)/K)(ert − 1)], does
of  ∼1 day, the density of bacteria n(t) is measured (usually by indeed saturate, as we show in figure 2(F). This model is in
determining the turbidity of the suspension1) and the results quite good agreement with measured growth curves for exper-
are plotted as a function of time t. Typical results are shown iments in simple nutrient media (figure 2(F))3.
in figure  2(D). These ‘growth curves’ have a characteristic Saturating population growth can also be modelled in a
shape: an initial period, known as the lag phase, in which more biologically consistent way by including the dynamics
no growth is detected, followed by a period of exponential of the nutrient explicitly in the equations. The classic equa-
growth (known as the exponential phase), followed by a slow- tion for the nutrient-concentration dependent growth of a bac-
ing down and eventual cessation of net growth, known as the terial population is:
stationary phase. It is generally stated that the lag phase hap-  
dN rmax s
pens because the bacteria need to adjust to the liquid medium (8) = N,
(having typically been stored under different conditions), dt Ks + s
while stationary phase happens when the population exhausts where s is the nutrient concentration, rmax is the maximal per-
its nutrient supply, or builds up waste products. However, the cell growth rate and Ks is the nutrient concentration at which
details of what happens during the lag and stationary phases the growth rate is half-maximal. In equation (8), the per-cell
remains a topic of active research [83, 84]. growth rate is described by a ‘Monod function’ [92]:
Simple equations can be used to describe the results of a
g(s) = rmax s/(Ks + s),
(9)
batch culture growth experiment. Assuming initially that the
nutrients are unlimited, the dynamics of the bacterial popula- which depends linearly on the nutrient concentration s for
tion can be modelled as low nutrient concentrations, but becomes independent of the
nutrient as s → ∞. This captures the fact that for high nutrient
dN(t)
(5) = rN(t), concentration, growth is limited by the bacterium’s capacity
dt to import and use the nutrient, rather than by the availability of
where N(t) is the number of bacteria at time t and r is the the nutrient in the environment. Equation (8) must be coupled
per-bacterium replication rate. Equation (5) predicts that the with a dynamical equation for the nutrient concentration:
population grows exponentially  
ds rmax s
N(t) = ert N(0) = 2t/T N(0),
(6) (10) = −γ N,
dt Ks + s
where T = (ln 2)/r is the doubling time, defined by where γ is a yield coefficient, describing the number of units
N(t + T) = 2N(t). For E. coli under optimal conditions (rich of nutrient that are consumed to produce one bacterium
nutrient broth, 37 °C), T ≈ 20 min which gives r ≈ 2.1 h−1. (divided by the volume).
Equation  (5) is appropriate for relatively large populations
(103 cells). For smaller populations, it may be important to
consider that replication is not a continuous process but occurs 2
In a population starting from a single bacterium, division events in different
cells occur quasi-synchronously for about the first 10 generations [67, 68].
3
The sharp-eyed reader will note that the solution of the logistic equation (7)
1
For a bacterial suspension, turbidity is usually referred to as ‘optical is not in perfect agreement with the MOPS growth curve from figure 2(F), and
density’, or OD. The OD has been shown to correlate well with the biomass cannot replicate the LB growth curve from figure 2(D). The shapes of growth
density in the sample [82]. Other techniques to measure bacterial density curves can in general be more complicated than suggested by these simple
include spreading the suspension on an agar gel of nutrient media, incubat- models, especially where there is more than one growth-limiting nutrient
ing and counting the resulting colonies, or direct counting of cells using a [31, 87]. More complicated models, such as those that use density-dependent
Coulter counter or flow cytometer. growth functions [88–91], have been developed to try to achieve better fits.

5
Rep. Prog. Phys. 82 (2019) 016601 Review

Figure 2.  (A) Sketch of a typical batch culture bacterial growth experiment. (B) and (C) Containers used to grow bacteria: a large-volume
flask (100 ml), and a microtiter plate with 96 individual culture wells of volume  ∼400 μl. (D) Measured growth curves for
E. coli strain MG1655 in simple- (‘rich’ MOPS: glucose, aminoacids, nucleotides, salts (blue curve)) and complex-nutrient medium
(LB broth, yellow curve). ‘OD’ is a measure of the turbidity of the suspension; see footnote to main text. The MOPS medium was created
by mixing 100 ml M2101, 100 ml M2103, 200 ml M2104 (Teknova), 10 ml 0.132M K2HPO4, 1 g glucose, and double-distilled, autoclaved
water to a total volume of 1000 ml. The LB medium consists of 25 g of LB powder (Fisher): tryptone, yeast extract and NaCl, dissolved
in 1000 ml of distilled water and autoclaved. 200 μl of the medium was added to each well of a 96-well plate (panel C), inoculated with 1
μl of PBS-washed overnight culture of E. coli, and incubated at 37C in a BMG FLUOstar plate reader for 24 h. OD was measured every
2 min with shaking for 20 s prior to each measurement. (E) The exponential growth model (equation (5), black curve) fits the experimental
MOPS curve from panel D for low bacterial densities. Fitting to the data in the range t = 1.5 . . . 3.5 h gives an exponential growth rate
r  =  1.94 h−1. (F) Comparison between different models and an experimental growth curve for growth in rich MOPS: the logistic growth
model (equation (7)) is shown by the green line and the Monod growth model (equation (10)) is shown by the red line. The best-fit
maximum growth rate is 2.2 h−1 (logistic growth) and 2.1 h−1 (Monod growth).

Numerical solution of equations (8) and (10) predicts that bacterial growth experiment. An alternative approach is to
the bacterial population size saturates as the nutrient runs out. use a chemostat: a well-mixed vessel in which fresh nutrient
This solution agrees well with experimental data (figure 2(F)), medium is supplied from a reservoir at a constant flow rate,
although it is typically not significantly better than the solu- and the contents of the vessel (bacteria and spent medium)
tion of the logistic growth equation (7)4. are removed at the same rate, so as to keep the volume con-
stant (figures 3(A) and (B)). In the chemostat, one achieves
3.1.2. Growth in a chemostat.  The batch culture setup a steady-state population in which the rate of bacterial repli-
shown in figure 2 is not the only way to perform a well-mixed cation is matched by the rate of removal of bacteria.
The dynamics of bacterial growth in a chemostat can be
4
In addition, the Monod relation (8) has the rather unsatisfactory feature modelled by making minor modifications to equations (8) and
that it is an ad hoc function, rather than being derived from any underlying (10) to account for the inflow of nutrient and the outflow of
model of the cell’s biochemistry. Because of this, attempts have been made bacteria plus medium. The resulting equations are [98]:
over many years to develop more complex nutrient-dependent growth equa-
tions, which take into account features such as population-size dependence
 
dN rmax s
[93], temperature [94], multiple nutrients [95], pH [96], and the thermody- (11) = N − Nd,
namic driving force for the biochemical growth reaction [97]. dt Ks + s

6
Rep. Prog. Phys. 82 (2019) 016601 Review

A Pump B C
1.0

0.8

0.6

N t,st
0.4
Bacteria
Waste 0.2

Growth 0.0
Medium 0 1 2 3 4 5
t
Magnetic stirrer
Figure 3.  (A) Schematic illustration of a chemostat. (B) Photograph of a simple chemostat consisting of a glass flask with tubing for media
delivery, aeration, and waste removal, and a magnetic bead for stirring the flask contents. (C) Example curves N(t), s(t) (black and green,
respectively), obtained by numerical simulation of equations (11) and (12) for γ = 1, Ks = 0.1, s0 = 1, rmax = 2 and d  =  1 (solid lines) and
d  =  3 (dashed lines). The simulations are initiated with N(0) = 0.5 and s(0) = 1 (in arbitrary concentration units). dcrit = 1.82 for this set
of parameters. For the solid lines, d < dcrit and a stable bacterial population is maintained in the chemostat (the solid black line reaches a
non-zero steady state); for the dashed lines, d > dcrit and the population is ‘washed out’ (the dashed black line goes to zero).

  solutions (showing, for example, oscillatory dynamics [98,


ds rmax s 103]).
(12)
= −γ N + s0 d − sd,
dt Ks + s
3.1.3. Growth of small populations.  The models which we
where d is the rate of fluid flow into and out of the chemostat
have discussed so far are all deterministic; they represent the
and s0 is the concentration of nutrient in the reservoir. These
dynamics of large bacterial populations, for which fluctua-
equations have the following steady-state solution:
tions in the population size are negligible. Recently, however,
rmax s0 − d(Ks + s0 ) it has become possible to study the dynamics of small bac-
N∗ =
(13) ,
γ(rmax − d) terial populations using microfluidic devices coupled with
microscopy [69, 104, 105]. For example, microfluidic chemo-
dKs stats have been constructed in which the population size is
s∗ =
(14) , 102–104 bacteria [106, 107]. Here, fluctuations in population
rmax − d
size become important and stochastic models are needed. The
for d < dcrit = (rmax s0 )/(Ks + s0 ), and N ∗ = 0, s∗ = 0 if the
birth–death process provides a natural way to model such a
flow rate is larger than dcrit. Therefore, growth in the chemo-
population. If we assume that bacterial reproduction and death
stat is possible only if the flow rate is lower than the maxi-
(removal from the system) are Poisson processes with rates r
mum growth rate of the bacteria5. Figure 3(C) shows example
and d, then we can write the following Master equation for the
plots of the bacterial density and the nutrient as a function of
probability P(N, t) that N bacteria are present at time t:
time, predicted by equations (11) and (12) for two values of d,
above and below dcrit. dP(N, t)
= − (r + d)NP(N, t) + r(N − 1)P(N − 1, t)
The chemostat equations  (11) and (12) can be extended  dt
to predict the dynamics of multiple competing or cooperat- + d(N + 1)P(N + 1, t). (16)
ing populations (see section  3.2), populations preyed on by
The statistical properties of such birth–death processes have
viruses, evolving populations, etc [99–102], providing a well-
been well studied [108–110]. One can think of this process as
founded mathematical model for a host of ecological scenar-
a biased random walk in the space of N, the population size,
ios. Many of these models have mathematically interesting
with the strength of the bias being given by r  −  d. If r  <  d
then the removal rate exceeds the birth rate and one expects
5
The same effect of ‘washing out’ of the population with a high dilution
the population to become extinct within a finite time (i.e. to
rate can also be observed in a simpler, logistic-like model without explicit reach the absorbing state at N  =  0). On the other hand, if
nutrients: r  >  d, then on average the population increases exponentially,
dN N ∼ exp[(r − d)t], but in any given realisation of the dynam-
(15)
= rN(1 − N/K) − dN, ics there is a non-zero probability
dt
which has steady state solution N*  =  K(1  −  d/r) for d  <  r, and N*  =  0 ρN0 = (r/d)N0
otherwise.
(17)

7
Rep. Prog. Phys. 82 (2019) 016601 Review

that the population will become extinct. This probability concentration, the growth rate g(s) ≡ rmax s/ (Ks + s), d is the
decreases exponentially with the initial size N0 of the popula- rate of bacterial removal from the system (by death or dilu-
tion since the lineage arising from each of the initial cells can tion), b is the rate of nutrient supply, R is the rate of nutrient
go extinct with probability ρ1 = r/d [110]. In the critical case removal and Γ ≡ Vγ is the yield coefficient. Equations (18)
where r  =  d the population fluctuates randomly (as an unbi- and (19) have a single (non-trivial) fixed point at
ased random walk in N) and will eventually become extinct, n∗ = (b − RdKs /(rmax − d)) /(Γd) and s∗ = dKs /(rmax − d).
but the average time to extinction is infinite. This solution is independent of the volume of the system because
Branching and birth–death processes similar to that of the model described by equations  (18) and (19) is deter-
equation (16) have been applied to model bacterial evolution. ministic. Linear stability analysis reveals how the system
A classical example is the Luria–Delbrück model [9], or, more approaches this fixed point [112]. If the eigenvalues of the
precisely, its stochastic version by Lea and Coulson [10]. This Jacobian matrix J
model predicts the distribution of the small number of mutant    
bacteria in a large growing population of wild-type (unmu- ∂f /∂n ∂f /∂s g(s) − d n(dg/ds)
J= = ,
tated) bacteria. Comparing the experimentally observed distri- ∂h/∂n ∂h/∂s −Γg(s) −Γn(dg/ds) − R
bution for this quantity with the model prediction is a standard  (20)
method for estimating mutation probability in bacteria (this is evaluated at the fixed point (n∗ , s∗ ), are real and negative, then
known as a ‘fluctuation test’, see [10]). For recent develop- we expect the system to relax monotonically to its fixed point.
ments in this field, see e.g. [111]. In contrast, if the eigenvalues are complex with a negative real
In the next two sections, 3.2 and 3.3, we review several part then we expect exponentially decaying damped oscilla-
pieces of recent work in which the models described above tions as the system approaches the fixed point. The matrix J
are extended to study more complex situations: specifically, evaluated at (n∗ , s∗ ) is given by
noise-driven oscillations in small bacterial populations, and  
populations of bacteria that switch stochastically between dif- 0 β/χ
(21) J∗ = d ,
ferent states. −Γ −β − χ

where we have defined β ≡ (Γn∗ /d)(dg/ds)s=s∗ =


3.2.  Example 1: noise-driven oscillations in bacterial (Γn∗ /d) × [rmax Ks /(Ks + s∗ )2 ], and χ ≡ R/d
. The eigenvalues
*
populations λ of J are given by 2λ/d = −(β + χ) ± (β + χ)2 − 4β .
If χ  1, i.e. the nutrient removal rate R is greater than the
The chemostat, described in section  3.1.2, is designed to
bacterial removal rate d (e.g. because bacteria adhere to a
achieve a steady state of growth for a large bacterial popu-
surface), then λ is real and negative for any value of β, and
lation, by supplying fresh medium at the same rate as spent
we expect the system to approach the fixed point monotoni-
medium (plus bacteria) is removed. In the natural environ­
cally. However, if χ < 1, i.e. bacteria are removed faster than
ment, however, bacteria may experience conditions that are
nutrient (e.g. due to death), then the eigenvalues are com-
very different to those of a chemostat. The population size
plex with negative real part, for β values in a range such that
may be small, as discussed above (e.g. for bacteria inhabiting
(β + χ)2 < 4β . This implies that transient oscillations can
the spaces between soil granules, or growing inside a human
happen as the system approaches the fixed point. The fre-
or animal host cell), nutrient supply may be unpredictable,
quency Ω of the oscillations isgiven by the imaginary part
and bacteria may be removed from the system not just by dilu-
of the eigenvalues λ: 2Ω/d = 4β − (β + χ)2 . Figure 4(A)
tion but also by death due to viral predation or host immune
response. Bacteria may also be retained in the system if they shows that numerical simulations of equations (18) and (19)
adhere to a surface. Extending the chemostat equations (11) indeed predict significant oscillations for a set of parameters
and (12) to include these factors reveals interesting predic- corresponding approximately to E. coli growing on glucose
tions, one of which is that noise-driven stochastic oscillations (see figure caption for details) [26, 31, 113].
may be a common feature of small bacterial populations in the What causes these transient oscillations? Intuitively, they
natural environment [102]. happen because the system builds up surpluses and deficits
To see this, let us start by analysing the deterministic che- of nutrient, relative to the bacterial population density. When
mostat equations (11) and (12), modified to allow for unequal there is a surplus of nutrient, the bacterial population grows
rates of nutrient supply and removal, and for bacterial death. rapidly and overshoots the amount of available nutrient, lead-
These are: ing to a sudden deficit of nutrient, upon which the popula-
tion decreases rapidly and eventually undershoots the nutrient
dn concentration, leading to a nutrient surplus. A transient nutri-
= f (n, s) = (g(s) − d) n,
(18)
dt ent surplus can only happen if excess nutrient is allowed to
accumulate in the system without being washed away; thus
ds
= h(n, s) = −Γg(s)n + b − Rs,
(19) the requirement for χ < 1. One can also explain the require-
dt ment for an intemediate value of β, such that (β + χ)2 < 4β .
where n = N/V is the bacterial density (with N being the The parameter β measures the responsiveness of the bacterial
number of cells and V the chemostat volume), s is the nutrient growth rate to changes in the nutrient concentration. For very

8
Rep. Prog. Phys. 82 (2019) 016601 Review

Figure 4.  Dynamical predictions for the population density of E. coli bacteria growing on glucose, with parameters rmax = 1 h−1,
d  =  0.5 h−1, Ks = 1 μM and Γ = 1.8 × 1010 glucose molecules per bacterium (data courtesy of Bhavin Khatri). The nutrient inflow rate b is
varied, such that we have β = 0.1, χ = 0 (corresponding to b  =  0.1 μM h−1 and R  =  0; such that s*  =  1 μM and n∗ ≈ 107 bacteria per litre)
and β = 0.01, χ = 0 (corresponding to b  =  0.01 μM h−1 and R  =  0; such that s*  =  1 μM and n∗ ≈ 106 bacteria per litre). Panel (A) shows
results for the deterministic model, equations (18) and (19). Panel (B) shows results for the stochastic model, equation (22), for a system
volume of 1 ml; i.e. for approximate absolute bacterial numbers of 104 (red) and 103 (blue). Note the different time axes in the two panels.
In these simulations, the bacterial densities are much lower than in a typical microbiology lab experiment, in fact they are 1–2 orders of
magnitude lower than the bacterial density in seawater, but this is similar to the bacterial density that might be found in drinking water.

small values of β, the growth rate does not respond to changes


in nutrient, so transient nutrient surpluses will not translate dx  1/2
=a+ B η(t).
into bacterial population oscillations. For very large values dt
(22)
of β, the population tracks the nutrient concentration closely, Here, x ≡ (n, s), a ≡ ( f (n, s), g(n, s)) describes the determin-
preventing nutrient surpluses or deficits from building up. istic dynamics, and η(t) is a vector of independent, Gaussian-
This analysis suggests that, in some situations in the natu- distributed random numbers with zero mean and variance
ral environment, bacterial populations whose dynamics is scaling with the inverse of the system volume (thus, the effects
deterministic may undergo transient (damped) oscillations, of noise are more important for small system volume). The
eventually reaching a non-oscillating steady state. But what matrix B is given by
happens for very small populations? It turns out that for small  
populations stochastic fluctuations due to the birth and death/ ng(s) + dn −Γng(s)
(23) B= .
removal of individual bacteria (demographic noise) drive sus- −Γng(s) Γ2 ng(s) + b + Rs
tained oscillations in the population density and the nutrient
concentration. This takes account of the fact that fluctuations in the bacterial
The effects of demographic noise in small bacterial popu- population are coupled to fluctuations in the nutrient concentra-
lations can be modelled in various ways. If the fluctuations tion, and vice versa. Equation (22) can be derived via a Kramers–
due to the noise are expected to be large, then individual birth Moyal expansion [116]; briefly, one expresses the model as a
and death events should be modelled explicitly—typically set of chemical reactions, writes down the corre­sponding mas-
these would be modelled as Poisson processes, and simulat- ter equation, and Taylor expands it under the assumption that
ing using a kinetic Monte Carlo scheme such as the Gillespie changes in the number of molecules/bacteria due to firing of a
algorithm [114, 115]. However, if the fluctuations are expected single reaction are small [102, 116]. The use of Gaussian noise
to be small, they may be approximated by adding stochastic to model demographic stochasticity is particularly convenient in
noise terms to the deterministic equations (18) and (19). This problems like this one, where the nutrient is represented explic-
is the approach taken by Khatri et al [102], leading to a set of itly. This is because the number of molecules of nutrient is typi-
Langevin equations of the form cally far larger than the number of bacteria—thus the nutrient

9
Rep. Prog. Phys. 82 (2019) 016601 Review

essentially behaves deterministically. We note that this Langevin stochastic switching is in general not known (and may dif-
approach can be very convenient: the alternative, using a scheme, fer in different cases) [131]: suggestions have included eva-
would kinetic Monte Carlo scheme would require one either to sion of host immune responses [130, 132, 133], avoidance of
simulate nutrient molecules discretely (which would be highly evo­lutionary fitness valleys [134], division of labour among
inefficient), or to adjust the kinetic Monte Carlo algorithm to cells in the population [135] or ‘bet hedging’ to ensure sur-
take account of a time-varying, continuous, nutrient concentra- vival of the population in an unpredictable environment [136,
tion (similar to [117]). 137]. The challenge of explaining the function of stochastic
Figure 4(B) shows the results of numerical simulations of switching in bacteria has motivated the development of a host
equation  (22), for the same parameter set as in figure  4(A) of theoretical models, of varying degrees of complexity
(representing E. coli growing on glucose), but for a system [133, 138–145]. Here, we review perhaps the simplest of
volume of 1 ml. These parameters represent a very low bacte- these, the case of randomly switching cells in a switching,
rial density (similar to that found in drinking water), so that unresponsive environment [138–143]. An elegant statisti-
the absolute numbers of bacteria present are  ∼104 (blue curve) cal physics model for this case was presented by Thattai and
or  ∼103 (red curve). It is immediately clear that demographic van Oudenaarden [140]; here we follow their approach, even
stochasticity has an important effect: the transient oscillations though it is rather idealistic from a biological point of view.
of the deterministic model (figure 4(A)) have been converted We suppose that individual bacteria in a population can be
into sustained oscillations in the stochastic model. The pres- in either of two states: a fast-growing state which we label
ence of these oscillations is also clearly visible in the power 1 and a slower-growing state, which we label 0. Bacteria
spectrum [102]. This effect may be widespread for very small switch stochastically between the two states, with a rate k0
bacterial populations; for example, it also happens in a model of switching from the fast- to the slow-growing state (1 → 0 )
of a nutrient-cycling bacterial ecosystem with two species and a rate k1 of switching from the slow- to the fast-growing
which feed on each others’ waste products [102]. state (0 → 1). This scenario can be described by the following
These stochastic oscillations are an example of a very gen- equations for the bacterial population dynamics:
eral mechanism that was discovered by McKane and Newman
dN0
in the context of predator-prey models [118], and later found = −k1 N0 + k0 N1 + γ0 N0 ,
(24)
dt
by other statistical physicists in a wide range of models
[119–126]. In this mechanisms, the underlying oscillatory
modes of a deterministic dynamical system are excited by a dN1
= +k1 N0 − k0 N1 + γ1 N1 .
(25)
source of intrinsic noise (in this case demographic noise), lead- dt
ing to sustained oscillatory dynamics in the stochastic version Here, N0 and N1 denote the numbers of bacteria in each of the
of the system, whereas the deterministic system shows only two states, and we have assumed that these are large enough to
damped oscillations. Thus, this example shows how insights be treated as continuous variables. The first two terms in each
from statistical physics can be important in understanding the equation describe switching between states, and the third term
behaviour of bacterial populations. describes growth.
Despite the possible ubiquity of this mechanism, demo- Equations (24) and (25) can be reduced to a single, non-
graphic-noise induced oscillations have not yet been observed linear dynamical equation  by defining the fraction f of the
for bacterial populations. One difficulty is that the effect is bacterial population which is in the faster-growing state 1 as
strong only if number of bacteria is very small (e.g. figure 4 f  =  N1/N, where N = N0 + N1. This leads to [140]
shows predictions for  ∼103–104 bacteria). The predictions are
also for a well-mixed system, while rules out typical exper­ df
(26) = k1 + f [∆γ − k1 − k0 ] − f 2 ∆γ,
imental methods where small populations are grown as micro- dt
colonies on agar plates (see section 4). However, well-mixed where ∆γ = γ1 − γ0 is the difference in growth rate between
conditions for small bacterial populations are starting to be the two states. The variable f is useful because it provides a
achieved using in microfluidic chemostats [106] or microflu- measure of the growth rate of the total population, as we can
idic droplets [127]. These techniques should eventually reveal see by summing equations (24) and (25):
a host of interesting fluctuation-driven dynamical phenomena.
dN
(27) = (γ1 + f ∆γ)N.
dt
3.3.  Example 2: switching bacteria in a switching
environ­ment
Since γ1 and ∆γ are constants, measuring f is equivalent
to measuring the total population growth rate. For this rea-
Up to now, we have mostly assumed that all cells within a son, f has been referred to as a measure of the ‘fitness’ of
bacterial population are identical. However, in many cases, the population [133, 140]. Equation  (26) predicts that the
genetically identical bacteria within a population can show variable f increases in time until it reaches a plateau value
variation in their levels of gene expression (see, e.g. figure 5). which corresponds to the positive root of the quadratic equa-
In the most striking cases, individual bacteria switch stochas- tion  k1 + f [∆γ − k1 − k0 ] − f 2 ∆γ = 0. Thus, although the
tically between very distinct states of gene expression, such total population size increases exponentially in time (equa-
that the population contains subpopulations with very dif- tion (27)), the fraction of the population that is in state 1
ferent behaviours [129, 130]. The biological function of this approaches a steady state.

10
Rep. Prog. Phys. 82 (2019) 016601 Review

Figure 5.  E. coli cells show variability in the expression of a gene encoding a fluorescent protein. A population of E. coli cells (strain
RJA003, created by P1 transduction from strain MRR [13] into MG1655) was grown in a 1 l chemostat with dilution rate 0.5 h−1 on Evans
media [128] supplemented with 50 mM glucose. The bacteria expressed cyan fluorescent protein (CFP) from a constitutive (unregulated)
promoter (the PR promoter from phage λ). After sampling from the chemostat, cells were kept on ice for  ∼1 h prior to being spread on
the surface of a 1% agarose pad and imaged in an epifluorescence microscope. (A) Fluorescence image in the CFP channel showing that
individual cells show different levels of fluorescence. (B) Histogram of fluorescence intensities per area, obtained from analysis of many
such images (units are arbitrary). The width of the histogram, relative to the mean, provides a measure of the population heterogeneity in
gene expression. Data shown courtesy of Joost Teixeira de Mattos, Alex ter Beek, Martijn Bekker and Tanneke den Blaauwen.

Now let us suppose that the bacterial population lives in a For even larger values of k1, switching becomes unfavourable
changing environment: specifically, the environment can flip, even in the periodic environment [140].
such that bacteria in the the slow-growing state become fast- It also turns out that, in regions of parameter space where
growing and vice versa. Thus, the fraction f1 of bacteria that bacterial switching is favoured, the optimal switching rate
are in the fast-growing state undergoes a jump: f1 → 1 − f1. matches the switching rate of the environment [140]. This
The environment can flip either periodically, or stochastically prediction has in fact been tested experimentally by Acar
with a fixed rate [138–143]. The environment is assumed to be et al [146], although using cells of the yeast Saccharomyces
‘unresponsive’, in the sense that its behaviour is not coupled cerevisiae rather than bacteria. In these experiments, yeast
to the state of the bacterial population. cells were engineered to switch stochastically between two
This simple and highly idealized model leads to some inter- states, in which expression of an enzyme for metabolising
esting results. In particular, one can ask what is the optimal the nutrient uracil was either on or off. Importantly, the rate
bacterial switching strategy, i.e. the strategy which maximises of switching could be controlled by addition of a chemical
the total population growth rate. Under what circumstances inducer. The yeast cells were grown in a turbidostat (a setup
should bacteria stochastically switch into a slower-growing similar to a chemostat, but where nutrient is supplied when
state, sacrificing fitness in the current environment, in order to the culture reaches a predefined cell density rather than con-
be prepared for a change in the environmental state? tinuously), in which the environment either contained uracil
For a periodic environment, it is possible to obtain an ana- (such that ON cells were fitter than OFF cells) or a toxic ana-
lytical solution for the average ‘fitness’  f , as a function of logue of uracil (such that OFF cells were fitter than ON cells).
the model parameters [140]. For a stochastically switching Figure 6(B) illustrates this setup: the environment was main-
environment one has to turn to simulations. In either case, it tained in state E1 for time T1 before being switched to state E2
turns out that for a certain a range of parameters (k0 , k1 , γ0 , γ1 ) for time T2. Yeast cells switched between the two phenotypic
bacterial switching out of the fast-growing state is favourable; states at (controllable) rates rON and rOFF and proliferated at
thus bacteria can increase their fitness by entering a slow- rates (here labelled γ) that depended on both the phenotypic
growing state, in readiness for the next environmental change. state and the environment. Figures 6(C) and (D) show results
Figure  6(A) shows the results of numerical simulations of of experiments for a ‘fast’ environment, in which T1 and T2
equation (26), in an environment that switches at rate 1, either are relatively short (figure 6(C)) and a ‘slow’ environment,
stochastically (as a Poisson process) or periodically. When the in which T1 and T2 are long (figure 6(D)). In the fast environ­
rate k1 of switching into the faster growth state is very small, ment, rapidly switching cells, with high rates rON and rOFF
the average fraction  f (k0 ) of slowly-growing cells is peaked (red data points) have, on average, a faster growth rate than
at a non-zero value of k0. This implies that it is favourable slow-switching cells, with lower rates rON and rOFF (blue
for cells to switch into the slower growth state at some non- data points). The situation is reversed, however, for the slow
zero rate, whether the environment is stochastic or periodic. environment.
However for a larger value of k1, switching into the slower This example shows that even the relatively simple case
growth state is favourable only in the periodic environment. in which the switching behaviour of the cells and of the

11
Rep. Prog. Phys. 82 (2019) 016601 Review

Figure 6.  (A) Predicted average value of f as a function of the rate k0 of switching into the less favourable state 0. These results were
obtained by numerical solution of equation (26), for an environment that switches stochastically as a Poisson process (with average
rate 1), or periodically (once per time unit). When the environment switches, we set f → 1 − f in the simulation, keeping all other
parameters fixed. When the rate k1 of switching into the faster growth state is very small, the function  f (k0 ) is peaked at a non-zero
value of k0, implying that it is favourable for cells to switch into the slower growth state at some non-zero rate, whether the environment
is stochastic or periodic. However for the larger value of k1 simulated here, switching into the slower growth state is favourable only in
the periodic environment. For even larger values of k1, switching becomes unfavourable even in the periodic environment [140]. (B)–(D)
Experiments with a stochastically switching strain of yeast cells, performed by Acar et al [146]. [146] (2008)© Springer Nature Limited.
With permission of Springer Nature. (B) Schematic illustration of the experimental setup. Yeast cells can be in either of two states, labelled
ON and OFF. Cells randomly switch between the states at rates rON and rOFF which can be tuned by the experimenter. The environment is
maintained in state E1 for time T1 before being switched to state E2, which is maintained for time T2. The proliferation rates γ of the two
cell types depend on the environment; the ON cell type proliferates faster in E1 while the OFF cell type proliferates faster in E2. (C) and
(D) Growth rates measured as a function of time during such an experiment, for cells that switch fast (red; rON ∼ 0.047 h−1, rOFF ∼ 0.035
h−1) or slow (blue; rON ∼ 0.004 h−1, rOFF ∼ 0.007 h−1). Panel (C) shows results for an experiment with T1  =  20 h, T2  =  37 h; here the fast-
switching cells have on average a higher growth rate. Panel (D) shows results for T1  =  96 h, T2  =  96 h; here the slow-switching cells have a
higher growth rate on average.

environment is uncoupled can produce non-trivial results, which stochastic phenotype switching is advantageous even
which go some way to explaining the possible advantages of in a fixed environment has also been considered, in [134].
stochastic switching. Real infections or environmental scenar-
ios are of course more complex, and other models have been 4.  Spatially structured bacterial populations
developed that reflect different aspects of this complexity.
For example, statistical physicists have considered the case In nature, bacterial populations rarely exist as well-stirred,
of an environmental switch which is triggered when the state homogeneous, liquid cultures. Instead, imperfect mixing,
of the population reaches a threshold (mimicking an immune combined with spatial heterogeneity of the environment (e.g.
response) [133]. With some approximations, this case can be gradients of food, oxygen, temperature), leads to the emer-
treated analytically and reveals a new possible role for stochas- gence of populations which are spatially structured, both
tic switching, in which the population composition is modu- genetically and phenotypically [147, 148]. These structured
lated so as to avoid triggering the environmental response. In populations often take the form of dense conglomerates in
other work, looking at the topic from a different perspective, which bacterial cells interact mechanically with each other.
statistical physics models have been used to investigate the An example is a bacterial biofilm, which is a dense mat of
relative benefits of ‘blind’ stochastic switching compared to cells attached to a surface [149, 150]. This might be a solid
‘responsive’ switching, in which cells detect the state of the surface such as a rock or soil particle, or a semi-solid matrix
environment and respond accordingly [144]. A scenario in such as food, animal or plant tissue. Biofilms are a source of
12
Rep. Prog. Phys. 82 (2019) 016601 Review

Figure 7.  (A) A colony of E. coli growing on the surface of an agarose gel in a Petri dish. The colony is about 7 mm wide and less than
1 mm thick. (B)–(D) Successive close-ups of a fragment of the colony’s rough border. In (D), individual bacteria may be seen.

concern in both medicine and industry because they can cause From a physicist’s point of view, the growth of biofilms
chronic infections when they form on medical implants, and and bacterial colonies are beautiful examples of self-assembly
biofouling when they form on industrial devices [151, 152]. processes, in which the structural properties of the popula-
In the microbiological lab, dense, spatially structured bac- tion are closely coupled with local gradients of nutrient (or,
terial populations are often encountered in the form of ‘colo- potentially, of signalling molecules or toxic substances such
nies’. These colonies arise when individual bacterial cells are as antibiotics) [81, 155–157]. As with many other statistical
dispersed across the surface of a layer of nutrient-containing physics models, the inclusion of space in models for bacterial
semi-solid agar gel and allowed to proliferate in an incubator population growth leads to many interesting new phenomena.
for a day or so (figure 7(A))6. The colonies are visible by eye Biofilm and colony self-assembly are particularly inter-
as small spots of size  ∼0.5–5 mm on the agar surface; each one esting from a statistical physics perspective because of their
contains  ∼108 or more bacteria, all of which are progeny of a connection with the well-established field of interface growth
single founder cell (figures 7(B)–(D)). While figure 7 shows models [158]. Interface growth models fall into a small num-
colonies formed by E. coli, figure 8 shows those formed by ber of universality classes, with well-defined scaling expo-
several genetic variants of the bacterium Pseudomonas aer- nents for the interface roughness as a function of time and
uginosa (PAO1 strain). These variants show strong differences system size [159]. Frustratingly, though, there are few exper­
in colony appearance due to differences in their production of imental systems [160–163] for which these theoretically-
extracellular polymers which affect cell–cell and cell-surface predicted scaling exponents can be measured. The edge of an
interactions. More generally, the shape and size of a bacterial expanding bacterial colony or the surface of a growing biofilm
colony depend on factors such as the nutrient concentration could provide an excellent system to measure such exponents
and the agar gel stiffness as well as the bacterial strain that is and may stimulate research into models that involve non-local
used [153, 154]. interactions between remote regions of the interface. Such
long-range interactions can happen in bacterial populations
6
Agarose is a polymer of agarobiose monomers, whereas agar is a natural due to the interplay between growth and nutrient/waste dif-
product produced by algae that contains a mixture of agarose and agaro-
pectin. Agar is cheaper and is used in plating experiments; agarose is more fusion, and the dynamics behind the front caused by physical
expensive but is often preferred for microscopy. interactions between growing cells [78, 164].

13
Rep. Prog. Phys. 82 (2019) 016601 Review

A B C

Figure 8.  (A)–(C) Colonies formed by genetic variants of Pseudomonas aeruginosa strain PA01 on the surface of a nutrient agar pad.
(A) A colony formed by the standard (non-mutated) version of this strain. (B): A‘mucoid’ colony formed by a strain that overproduces
extracellular polymeric substances. (C) A ‘rugose’ colony formed by a ‘rugose small colony variant’ (RSCV) strain. RSCV strains often
show increased levels of the intracellular signalling molecule cyclic di-GMP. Scale bars are the same in panels A, B and C. Image courtesy
of Yasuhiko Irie, University of Dayton. Used with permission.

Figure 9.  Top: schematic illustration of a model in which a spatially structured bacterial population is represented as a chain of well-
mixed sub-populations connected by migration. In this model, bacterial growth in each compartment is assumed to follow the logistic
model (equation (7)), and the migration rate per bacterium between neighbouring compartments is assumed to be a constant, m. Bottom:
the model predicts the emergence of travelling waves of bacteria. The plot shows the results of numerical simulations of equation (28), for
M = 24, r = 1, m = 0.1 and any K  >  0, in which bacteria are initially present only in compartment i  =  1. The curves correspond to times
t = 0, 7.5, 15, 22.5, 30 (from left to right).

4.1.  Modelling spatially structured bacterial populations each compartment can be described by the logistic model
(equation (7)), and that the migration rate per bacterium
Many different approaches can be used to model spatially
between neighbouring compartments is constant. This leads
structured bacterial populations, depending on the system
to the following set of differential equations:
being studied and the desired level of physical and biological
realism. dNi /dt = rNi (1 − Ni /K) + m(Ni+1 + Ni−1 − 2Ni ), for 1 < i < M
dN1 /dt = rN1 (1 − N1 /K) + m(N2 − N1 ),
4.1.1. Connected habitats and Fisher-KPP waves.  Perhaps dNM /dt = rNM (1 − NM /K) + m(NM−1 − NM ), (28)
the simplest approach is to construct a model that consists of 
connected well-mixed compartments, between which bacte- where Ni denotes the number of bacteria in compartment
ria can migrate (mimicking motility, diffusion or flow). The i = 1, . . . , M , r is the replication rate, K is the carrying capacity
dynamics of the bacterial population in each compartment and m is the migration rate. Let us suppose that bacteria are ini-
can be described using the same type of equations as in sec- tially present only in compartment i  =  1. In this case, the popu-
tion  3.1, with the additional of coupling terms to describe lation spreads in a wave-like manner to the other compartments,
migration of bacteria between compartments. This kind of as we show by numerical simulation in figure  9 (bottom). In
approach is appropriate for situations where the local environ­ the limit of many compartments, assuming a small distance ∆x
ment of the bacterial population is liquid-like, and the length- between compartments, and setting m/(∆x)2 → D , we can
scale over which the environment varies is large. It is often rewrite equation (28) as a partial differential equation:
used, for example, in large-scale models of ocean plankton 
dynamics [165], and in macro-scale models for the treatment ∂n ∂2n n
(29)
= D 2 + rn 1 −  ,
of infections with antibiotics [166]. ∂t ∂x K
As a specific example, let us consider a population of bac- where x denotes spatial position, n(x, t) is the local density of
teria growing in a chain of M connected compartments (figure bacteria and K  = K/V (with units of bacterial density; here
9, top). We assume that the bacterial growth dynamics within V is the volume of one compartment). This is an example of

14
Rep. Prog. Phys. 82 (2019) 016601 Review

Figure 10.  (A) Confocal laser scanning microscope image of a biofilm formed by P. aeruginosa PAO1 grown for 24 h in a flow cell.
Reproduced from [173] Copyright © 2016 American Society for Microbiology. (B) Illustration of a simple model of a growing biofilm with
a flat boundary. The position of the boundary is given by z = h(t) and the nutrient profile is s(z).

a Fisher–Kolmogorov–Petrovsky–Piscounov (FKPP) equa- where g(s) is the growth rate, which depends on the local
tion  [86]; its solutions are travelling waves similar to those nutrient concentration s, for example via a Monod function
observed for the discrete case (figure 9). Stochastic versions (equation (9)). The dynamics of the nutrient concentration
of the FKKP equation have also been studied [167, 168], and s(z, t) is governed by diffusion into the biofilm and consump-
these may provide a good description of bacterial population tion by the bacteria:
dynamics in some circumstances [169]. Related approaches
have also been used to model more complex situations, ∂s(z, t) ∂ 2 s(z, t)
(31) =D − Γng(s(z, t))Θ(h(t) − z).
including the spatial expansion of several interacting bacterial ∂t ∂z2
populations [170] and the evolution of resistant bacteria in a Here, D represents the diffusion constant of the nutrient
drug gradient [171, 172]. (assumed to be the same inside and outside the biofilm, for
simplicity), Γ is a yield coefficient (nutrient consumed per
4.1.2. Continuum models for dense populations.  The situa- unit of biomass created) and Θ is the Heaviside step func-
tion is different when bacteria grow in a densely packed assem- tion. Depending on the boundary conditions for the nutrient
bly such as a colony or a biofilm, in which individual cells do field s(z, t), the choice of growth function g(s) and the param­
not migrate freely. Here, physical interactions between bacte- eters D and Γ, this model can predict linear √ growth: h(t) ∝ t,
ria are likely to be important, and there are also likely to be growth that slows down in time: h(t) ∼ t , or exponential
steep local gradients of nutrient or other chemicals, making it growth: ln[h(t)] ∼ t [155, 164, 175].
necessary to model chemical concentration fields explicitly. In In the above example, we have assumed a flat biofilm, which
such cases, one can still use a continuum approach, in which allows us to reduce the problem to one dimension. In real-
both chemical concentrations and the bacterial population ity, however, most biofilms have rough surfaces when grown
density are represented as continuous fields (as in the FKPP in the typical laboratory flow cell setup (e.g. figure  10(A))
equation), but the equations  must be formulated differently [176]; some even have ‘mushroom’-like protrusions [177].
[164, 174]. More realistic continuum models take surface roughness into
As an example, let us consider the growth of a biofilm on account by representing the biofilm in two or three dimen-
a solid surface, as shown in figure 10(A). If we suppose that sions, and also account for spatially varying bacterial density
the bacterial cells are densely packed then it is reasonable to and local pressure within the biofilm [155, 175, 178]. Such
assume the bacterial density n is constant within the biofilm. If models can show interesting phenomena including a ‘finger-
we also assume the biofilm is flat (even though this is clearly ing instability’ [155, 175, 179], as we discuss in more detail
not a good assumption for the biofilm of figure 10(A)!), then in section 4.2. ‘Active nematics’ models that take into account
the problem becomes one-dimensional and the bacterial den- orientation of non-spherical cells inside the biofilm can also
sity as a function of the vertical coordinate is a step function of explain features of bacterial colonies such as the existence of
height h(t) (figure 10(B)). As the biofilm grows, h(t) increases nematic-like defects and micro-domains of locally-aligned
to accommodate the increase in biomass. The dynamics of cells [180, 181] .
h(t) can be written as
 h(t) 4.1.3. Individual-based models, on and off-lattice.  In some
∂h(t) situations, it is not appropriate to treat a spatially structured
(30) = ng(s(z, t))dz,
∂t 0 bacterial population as a continuous field; one requires instead

15
Rep. Prog. Phys. 82 (2019) 016601 Review

A B C

Figure 11.  (A) The Eden model on a 2D lattice. Cells that are completely surrounded and cannot replicate are shown in yellow. The bright
green cell illustrates the replication rules: it can replicate to a neighbouring empty lattice site (including the diagonal ones in this variant
of the model) but not to an occupied one. (B) A snapshot of a simulation of the 2D Eden model in which a cluster of cells ( N = 65 536)
has grown from a single initial cell. (C) Simulation snapshot for a 2D Eden model in a domain of width L  =  250 that is semi-infinite in the
vertical dimension, with periodic boundaries in the horizontal dimension.

detailed spatial resolution at the level of individual cells. This been measured under conditions where the bacteria are non-
is the case, for example, if one is interested in small popula- motile [186, 187]. The results are somewhat mixed: some
tions, heterogeneous populations (e.g. stochastically switch- experiments have produced exponents α and β that are dif-
ing cells, as in section  3.3), or population-level processes ferent from those of the KPZ universality class (with various
that are triggered by single-cell events (as we shall see in suggested explanations [158, 186, 187]), while other experi-
section 4.3). Models in which the position and state of each ments have produced exponents consistent with KPZ [188]. It
bacterial cell is tracked in time are known as individual-based seems that the jury is still out on whether at least some bacte-
models or agent-based models. rial colonies fall into the KPZ universality class. For colonies
The simplest form of an individual-based model of a bacte- of motile bacteria, or under conditions of low nutrient concen-
rial population is a lattice-based one, in which bacteria occupy tration, more complicated, fractal-like colony structures can
sites on a lattice and reproduce into neighbouring lattice sites arise [153, 189–191] and the interfacial growth exponents α
according to certain rules. A classic example is the Eden and β are very different to that of the KPZ universality class
model [15] (figure 11(A)). In the Eden model, lattice sites are (and correspondingly the Eden model).
either empty or occupied, and an occupied site, or ‘cell’, can The most serious limitation of the Eden model and similar
reproduce if empty sites are available in the neighbourhood lattice models is that growth is restricted to cells that are at
(different variants of the Eden model make different assump- the boundary of the cluster, and hence the centre of the cluster
tions about how the replication rate depends on the number of is static. This is not a good representation of most bacterial
empty neighbours [182–184]). Starting from a single occupied colonies. While bacteria in the centre of a colony do become
lattice site, the Eden model produces a cluster of occupied starved due to insufficient nutrient penetration as the colony
sites (figure 11(B)) whose interfacial properties fall into the becomes large, growth typically occurs in the outer parts of
Kardar–Parisi–Zhang (KPZ) universality class [185]. A par­ the colony in a layer of considerable thickness (tens of cellular
ticularly important descriptor of an interface is its roughness, diameters), and within this growing layer elongation and pro-
defined as the standard deviation of its height fluctuations. If liferation of bacteria behind the colony edge lead to pushing
we consider for simplicity a system with the geometry shown forces on bacteria that are closer to the edge. A similar picture
in figure 11(C) (an infinitely long column of width L sites), holds for bacterial biofilms. One can devise lattice models that
then the interface roughness W is given by are somewhat more realistic, by allowing cells in the centre
 to replicate and push away surrounding cells [17]. However,
2
(32) i (hi − h) off-lattice individual-based models offer a much greater level
W= ,
L of realism. Such models can account for physical interactions
where hi is the vertical height of the cluster of cells at horizon- between neighbouring bacterial cells and between bacteria
tal position i. The roughness scales as W ∼ tβ for short times and their environment, as well as the dynamics of nutrients,
and W ∼ Lα for long times, with the two critical exponents intra-cellular chemical signals, toxins, etc.
being α = 1/2 and β = 1/3 for the 2D Eden model. The same In off-lattice individual-based models, individual bacteria
critical exponents are obtained in an off-lattice version of the (usually modelled as disks in 2D, or spheres in 3D, although
Eden model [74]. rod-shaped cells can also be modelled) move in continuous
How well does the Eden model capture the behaviour space and interact via physical mechanisms (e.g. elastic repul-
of real bacterial populations? To our knowledge, interfacial sion, friction etc). These simulations are somewhat analogous
growth exponents have not been measured for flow-cell bio- to molecular dynamics or Brownian dynamics simulations
films like that of figure 10(A). For bacteria growing as colo- in condensed matter physics. For example, the dynam-
nies on agar gel surfaces, however, surface roughness W has ics of 2D rod-shaped bacteria whose motion is opposed by

16
Rep. Prog. Phys. 82 (2019) 016601 Review

viscous-like friction can be described by the following equations  Let us imagine a 2D population of bacteria which expands
[164, 192] in the z direction and is confined between two walls in the
x-direction (figure 12(E)). This could represent a 2D bacterial
dri /dt = F i /(ζli ),
(33) colony or a 2D section of a biofilm; here we will refer to it
as a biofilm. The interface of the growing biofilm has pro-
dφi /dt = 12τi /(ζli3 ),
(34) file z = h(x, t) at time t. We assume that bacteria grow only
where li is the length of bacterium i, ri is the position of its in a narrow zone of width b, close to the interface, because
centre of mass, φi is its angular orientation, F i and τi are nutrient does not penetrate far into the biofilm. This assump-
the total force and torque acting on it, and ζ is the friction tion is based on simulations like that shown in figure 12(D),
(damping) coefficient. The dependence on li comes from the in which both bacteria and nutrients are modelled explicitly;
assumption that every infinitesimally thin section  of the rod here the bacteria shown in bright green, which are close to
experiences a friction force proportional to the local veloc- the nutrient, are able to replicate, while the bacteria shown
ity. The model must also account for bacterial growth (here, in dark green, which are far from the nutrient, are not able
increase in li with time) and division. The rate of growth typi- to replicate. Although we do not model the nutrient concen-
cally depends on a local nutrient concentration field which is tration field explicitly here, we will assume that parts of the
represented on a grid and is updated at each timestep accord- interface that protrude in the z direction experience a higher
ing to a reaction-diffusion equation  accounting for diffusive nutrient concentration because they are closer to a nutrient
transport and bacterial consumption: source (this would be the case in a typical biofilm flow setup
 [176]). Thus, the local growth rate depends on the height of
∂s 2
(35) = D∇ s − γ g(s(r i )). the interface: g(x, t) = f (h(x, t) − h̄(t)), where h̄(t) is the aver-
∂t i age height at time t (figure 12(E)). We also suppose that the
A variety of models of this type have been developed and used interface has a ‘stiffness’, or a tendency to be flat. This is an
to simulate bacterial colonies and biofilms [71, 77, 78, 164, ad hoc assumption, but it mimics, to some extent, adhesion
193–195]; they differ in their choices of which physical inter- between the bacteria. The dynamics of the interface can then
actions to include (and how to include them), as well as in how be described approximately as
they account for biological details such as bacterial shape and  2 
∂h ∂ h
metabolism. = bf (h(x, t) − h̄(t)) ζ 2 + 1 .
∂t ∂x
In the next three sections, 4.2–4.4, we discuss three exam- (36)
ples of interesting phenomena produced by bacterial growth In equation (36), the term ζ∂ 2 h/∂x2 accounts for the surface
in spatially structured environments: fingering instabilities at stiffness by favouring growth in concave regions of the inter-
the edges of colonies and biofilms, the transition from 2D to face (troughs) and disfavouring growth in convex regions
3D colony growth and the emergence of genetically segre- (peaks).
gated sectors during expansion of a colony. In choosing these To see how this model can produce interesting behaviour,
examples, we focus on phenomena that are unconnected with let us make a small perturbation around an initially flat inter-
bacterial motility, since motility-induced collective phenom- facial profile:
ena in bacterial populations, and in general, have been exten-
(37) h(x, t) = h̄(t) + (t)eikx .
sively described elsewhere (see, e.g. [18–20, 23, 25, 60, 196]).
Inserting this into equation (36) and expanding to first order
in the magnitude of the perturbation,   1, we obtain the fol-
4.2.  Example: fingering instabilities at the interfaces
of bacterial colonies and biofilms
lowing equation for (t):
d(t)  
As we have already mentioned, the expanding edge of a (38) = b(t) f  (0) − ζk2 f (0) ,
growing bacterial colony, and the surface of a growing bio- dt
film, are examples of growing interfaces. Depending on the where f (0) is the derivative of the growth function f (h − h̄),


growth conditions, the interface can be smooth, rough or fea- evaluated for h = h̄ . Thus, ε is predicted to grow in time for
ture long finger-like shapes (for colonies) or ‘mushrooms’ perturbations whose wavenumber k obeys
(for biofilms) [153, 154, 158, 177, 186, 187, 189, 191]; see
k2 < f  (0)/(ζf (0)).
(39)
figures 12(A)–(C).
Various theoretical approaches have been used to model This condition is equivalent to stating that interfacial
the shape of these interfaces, ranging from continuum equa- ‘bumps’ of dimension Λ = 2π/k will tend to grow if

tions  [175, 178] to individual-based models [74, 164, 192]. Λ > 2π ζf (0)/f  (0). This means that the interface will be
Rather than describing these in detail here, we will instead unstable to the growth of finger-like protrusions, provided the
use a simple toy model to illustrate some basic factors that can width of the system L is large enough to allow such protru-

affect the shape of the growing edge of a bacterial population. sions to develop, i.e. for systems of size L > 2π ζf (0)/f  (0)
Although the model that we will present here is unrealistic in (figure 12(F)). Thus, a transition from a smooth to a fingered
many ways, it has the advantage of allowing a simple math- front is predicted to occur for a growing bacterial population if
ematical analysis. (i) the spatial extent L of the population is big enough, (ii) the

17
Rep. Prog. Phys. 82 (2019) 016601 Review

Figure 12.  (A)–(C) Examples of E. coli colonies (on 2 mm-thick, 2% agarose infused with LB) with different roughness of the colony
boundary: smooth (A), rough (B), and branched (C). These different shapes have been obtained by using different strains (MG1655 for
B, MG1655ΔfimAΔfliF for A [211]), or incubating colonies of MG1655 for different amounts of time at 37C (B = short time, C = long
time). The colony size is 2-3 mm. (D) Boundary of a simulated colony (simulation details as in [192]). The nutrient concentration is shown
as different shades of red (brightest colour = highest concentration). Replicating cells are shown in bright green, whereas stationary cells
with no access to nutrients are shown in dark green. (E) Schematic illustration of the model from equation (36). (F) Interface profiles h(x, t)
obtained by numerically solving equation (36) for L = 4, ζ = 0.02, b = 1, f (h) = 1/(1 + e−h ), and t = 0, . . . , 10
. The initial condition is
a superposition of two sine functions with periods 4 and 1. The oscillation with period 1 is damped (1 < Λ = 2π ζf (0)/f  (0) ≈ 1.256),
whereas the one with period 4 grows in time.

stiffness ζ of the interface is small enough and (iii) the growth Experimental work on E. coli colonies growing on agarose
function f depends strongly enough on the height—e.g. due to suggests that mechanical forces are likely to play an impor-
rapid nutrient consumption or slow nutrient diffusion [164]. tant role in the 2D to 3D transition [77, 78, 199]. In a setup
This toy model is of course highly simplistic—among where bacteria are sandwiched between the agar and a glass
other deficiencies, it does not account for the dynamics of coverslip, microscopic tracking of the growth of colonies
the nutrient, or for changes in the thickness of the growing from single cells reveals a well-defined ‘buckling transition’
layer as the colony expands. Nevertheless it illustrates how at which bacteria start to invade the agarose, leading eventu-
instabilities can arise from the coupling between the shape ally to 3D growth (figures 13(A)–(C)) [78]. In this transition,
of the growing colony/biofilm interface and the local avail- the first cells to invade the agarose are usually located close
ability of nutrient. Similar phenomena, driven by the same to the centre of the 2D colony. Moreover, the average size of
interface-nutrient coupling, also arise in more realistic models the 2D colony at the moment when this transition happens
[155, 157, 173, 175, 178]. depends non-monotonically on the concentration (and hence
the stiffness) of the agarose gel: it happens later (i.e. at larger
colony area) for intermediate agarose stiffness. Using indi-
4.3.  Example: 2D to 3D transition in bacterial colony growth
vidual-based simulations, Grant et al [78] could match these
Another interesting feature of bacterial colony or biofilm experimental results, under the assumption that the friction
growth is the transition from 2D to 3D growth. Starting from a coefficient between the bacteria and the agarose has a par­
single cell seeded on an agarose gel surface, a colony initially ticular non-linear dependence on the agarose stiffness.
spreads as a 2D layer of cells on the surface, but it later devel- While Grant et  al’s simulations were quite complex, the
ops into a 3D structure. If the agarose gel surface is covered basic physics that may control the invasion transition can be
by a glass coverslip (with the bacteria sandwiched between illustrated with a much simpler model (figure 13(D)). Let us
the agarose and the coverslip), then the colony becomes 3D by imagine a 1D chain of bacteria, extending from x  =  −L/2
growing into the agarose layer. However, if there is no bound- to x  =  L/2. The bacteria elongate at rate g and so the chain
ing coverslip, the colony instead expands into the space on top length L(t) = L0 exp(gt) increases with time. As the bacteria
of the agarose layer. Biofilm growth on a solid surface also grow, they exert outward pushing forces on each other and
often starts with the proliferation of flat microcolonies, which experience inward forces due to friction with the surrounding
later expand vertically. This 2D to 3D transition has parallels medium (here assumed to be agarose). This produces a local
in the growth of some cancer tumours [197] and in embryonic stress σ(x, t) within the chain. Because the frictional forces
development [198]. are transmitted along the chain of bacteria, we expect σ(x, t)

18
Rep. Prog. Phys. 82 (2019) 016601 Review

Figure 13. (A)–(C) 2D to 3D transition in bacterial colonies. The arrows indicate locations where the colony has invaded the agarose and a
second layer of cells has begun to form. (A) An image taken just before the transition. (B) An image of the same colony taken just after the
transition. (C) An image taken when the second layer of cells is already well-developed. (D) A simple model of the ‘buckling’ transition,
for a 1D chain of bacteria (see main text). (E) Length of the bacterial chain at the onset of the buckling transition, as a function of the
friction coefficient κ. The parameters are g  =  2 h−1, L0  =  1 μm, b  =  107 Pa−1, w0 = 10−7 h−1, p  =  105 Pa.

to be largest at the centre of the chain, x  =  0, and to increase being related to the friction coefficient. This form of σ might
in time as the chain elongates. This stress may cause the be expected if the frictional force generated by bacterial
chain to buckle (figure 13(D)); we denote the stress-depend- motion is equal for all bacteria, and the contributions of each
ent rate at which a bacterium buckles as w(σ). We suppose bacterium sum up along the chain.
that w is small for small stress σ, but increases strongly for Changing variables to y  =  L/2  −  x, using the symmetry of
large σ. One form of w(σ) consistent with this expectation is σ(x, t) about x  =  0 and substituting in the chosen form of w,
w(σ) = w0 exp [bσ], where b is some constant; for illustrative we obtain
purposes we will assume this form here (although it is not     
t L(t )/2
motivated by any mechanistic understanding of the buckling P(t) = 1 − exp −2 w(ky)dydt 
process; a beautiful mechanistic description of buckling has 0 0
   
recently been presented by Beroz et al [200]). 2w0 t
Focusing on a particular position x along the chain, the = 1 − exp − (exp [bkL(t )/2] − 1) dt . (41)
bk 0
probability that the chain has not buckled at this position
   We now use the fact that L = L0 exp (gt) to replace time t by
t
by time t is exp − 0 w(σ(x, t ))dt , and the probability
the chain length L. The probability Q(L) that the chain has not
that the chain has not buckled at any position by time t is buckled by the time it reaches length L is
    
t L(t )/2
exp − 0 −L(t )/2 w(σ(x, t ))dxdt . Therefore, the probabil-   
2w0 L exp (bkl/2) − 1
 
Q(L) = 1 − exp − dl
ity P(t) that a buckling event has happened by time t is 
bkg 0
 
l
    
     2w0 bkL bkL bkL
t L(t )/2 = 1 − exp − −γ + Shi + Chi − log ,
bkg 2 2 2
(40)P(t) = 1 − exp − w(σ(x, t ))dxdt .  (42)
0 −L(t )/2
where γ is the Euler–Mascheroni constant and Shi and Chi are
Now let us assume a particular form for the stress function: the sinh and cosh integral functions. The dependence of the
σ(x, t) = k[L(t)/2 − x] for x  >  0 and σ(x, t) = k[L(t)/2 + x] probability Q(L) on the length of the bacterial chain L arises
for x  <  0. This simply describes a linear decrease in stress only from terms in the combination bkL. This leads to our first
from the centre to the edge of the chain, with the constant k important observation: the critical size at which the buckling

19
Rep. Prog. Phys. 82 (2019) 016601 Review

transition happens is expected to scale approximately as of red and green sectors is visible. This implies genetic seg-
1/(bk): i.e. it decreases with increasing friction/adhesion k regation: the descendants of different cells within the founder
and increasing growth rate b. The most likely length L of the population occupy different regions of space [72]. The same
chain at which the transition happens can be determined from phenom­enon occurs for other microorganisms [73, 201], and
       for colony growth in different geometries [73, 202].
2w0 bkL bkL bkL
−γ + Shi + Chi − log ≈ 1. The emergence of sectors is closely connected with the
bkg 2 2 2 fact that (after an initial period of exponential growth) only
(43)
bacteria that are close to the expanding edge of the colony
To relate the predictions of this simple model to exper­
are able to replicate; deeper in the colony nutrient becomes
imental results such as those of Grant et al, the coefficient k in
depleted and waste products may accumulate. Demographic
the model must be related to the friction coefficient κ between
fluctuations at the growing colony front can cause a bacterial
bacteria and the agarose/glass surfaces, and the stress p that
lineage to become ‘trapped’ behind the front, in which case it
pushes a bacterium against the glass surface, due to elastic
cannot proliferate further. Thus, a stochastic process is at play,
compression of the agarose (figure 13(D)). Dimensional anal-
in which some lineages come to dominate the growing front
ysis suggest that k = κp/L0 . Figure 13(E) shows the resulting
(i.e. form sectors) while others are buried behind the front.
predictions of this simple model for the chain length at the
To better understand this process, we follow [73] and
onset of buckling, as a function of κ, for g  =  2 h−1, L0  =  1
imagine that the growing layer is infinitely thin and circular
μm, and with the parameters b  =  107 Pa−1 and w0 = 10−7
symmetric, such that the proliferating bacteria are located
h−1 chosen so that the result is comparable with the exper­
on the perimeter of a circle of radius R = vt expanding with
imentally observed buckling diameter (≈50 μm) of a 2D constant velocity v . Let us suppose that the initial radius
colony for κ = 0.7 and p  =  105 Pa [78]. The model predicts of the circle (i.e. the radius of the drop of bacteria that is
that the size of the colony upon buckling decreases with the deposited on the agar) is R0. We divide the perimeter into
friction coefficient κ which characterises the strength of cell- sectors, and we track the positions of the sector boundaries
agarose interactions. One can also use similar arguments to on the perimeter of the circle7. As time goes on, the bac-
show that the average position at which the chain buckles is
teria within each sector proliferate, or become lost behind
very close to its centre, in agreement with the experimental
the growing front, in a stochastic process. While the size of
results [78].
a sector will increase on average as the colony expands, at
This model, while it is undoubtedly simplistic, provides
any given moment in time it may fluctuate either upwards
some insight into the effects of friction on the 2D to 3D trans­
or downwards. A sector may even contract so much that it
ition. More generally, understanding the 2D to 3D transition
vanishes altogether, representing loss of the lineages of the
in bacterial colonies presents a host of interesting challenges.
bacteria within that sector. This stochastic dynamics can be
These include analysing simple statistical physics models like
modelled by the following Langevin equation  for the arc
the one described here, developing individual-based simula-
length w occupied by a sector:
tions that explicitly include interactions with the surrounding
elastic medium, and carrying out experimental measurements dw w √
= + 4Dη(R),
(44)
of the frictional and adhesion forces between bacteria and aga- dR R
rose and glass surfaces. It is also important to note that the with the initial condition w(R0 ) = w0. Note that because the
buckling transition that we have discussed here is only the first colony radius R increases linearly with t, tracking the dynam-
stage the development of a 3D bacterial colony. Once buckling ics as a function of R is equivalent to tracking it as a function
has happened, the subsequent development from a two-layered of time t. In equation  (44), the first term on the right hand
structure to a larger 3D colony also presents beautiful and side accounts for the radial expansion of the colony, which
interesting phenomena which remain to be explained [199]. ‘stretches’ the sector. The second term accounts for stochas-
ticity in the replication events and local movements of bac-
4.4.  Example: formation of clonal sectors in growing bacterial
teria at the front; here D is an effective diffusion constant
colonies and η(R) represents uncorrelated Gaussian noise with zero
mean and unit variance. The lack of scaling of η(R) with R
A very interesting feature of the growth of bacterial colonies is because intersector competition is assumed to occur only
and biofilms is the spatial distribution of lineages within the at sector boundaries of constant width. We stress that equa-
population—or in other words, the locations of the descen- tion  (44) defines an idealized mathematical model of a cir-
dants of a particular ‘founder cell’. Figure  14(A) shows the cular-symmetric, infinitesimally-thin edged colony; sector
outcome of a simple experiment in which a bacterial colony is dynamics in real colonies may deviate from it due to edge
initiated not from a single cell, but from a droplet containing roughness (see sections 4.1.3, 4.2, and [169]). In the absence
a mixture of two strains of E. coli which are identical except of the noise term, equation (44) predicts that w increases deter-
that they produce different-coloured fluorescent proteins (here ministically as w(R) = w0 R/R0. For D  >  0, however, the arc
shown red and green). The area covered by the initial droplet length w follows a biased random walk with a time-dependent
appears yellow, indicating a mixture of red and green cells.
In the surrounding regions, however, where the population 7
In this calculation we do not specify the number of bacteria in each sector
has expanded out from the initial droplet, a striking pattern as this turns out not to be important.

20
Rep. Prog. Phys. 82 (2019) 016601 Review

A B
3
2
1
0

y
1
2
3
3 2 1 0 1 2 3
100µm
x
Figure 14.  (A) An expanding population of fluorescently-labelled E. coli cells, growing on the surface of a nutrient agar pad. The
population was initiated from a drop containing a 50:50 mixture of cells labelled in two different colours (here shown red and green). The
population is mixed (appears yellow) in the region of the initial drop, but has segregated into clonal sectors at its expanding edge. Image
courtesy of Diarmuid Lloyd. Used with permission. (B) Results of a simulation in which sector boundaries are modelled as annihilating
random walks, as described by equation (44) with D  =  0.02, R0  =  1, R  =  3. The simulation starts with 50 random walkers.

diffusion constant. This is illustrated in figure 14(B) in which From an experimental point of view, one can easily measure
we plot trajectories of sector boundaries, simulated using the sizes of sectors in relatively large colonies (e.g. R ∼ a
equation (44). few mm), but it is much harder to measure sectors in very
To proceed further, we introduce the angular size of the small colonies. Therefore we would like to use equation (48)
sector, φ = w/R. In this coordinate, equation (44) becomes to predict the distribution of sizes of surviving sectors, in the
√ large colony limit R → ∞ . We first normalize equation (48)
dφ 4D
(45) = η(R). to obtain the distribution Psurv (φ, R|φ0 , R0 ) of sector sizes,
dR R conditioned on sector survival:
Thus, we see that the magnitude of the angular fluctuations  
decreases as the colony radius increases. Equation (45) can be ∼ φ φ2
(49) Psurv (φ, R|φ0 , R0 ) = 2 exp − 2 .
translated into a Fokker–Planck equation [203]: σ 2σ

∂P(φ, R|φ0 , R0 ) 2D ∂ 2 P(φ, R|φ0 , R0 ) The mean angular sector size in the limit R → ∞ is thus
(46) = 2 ,  ∞
∂R R ∂φ2
φ(R → ∞) = φPsurv (φ, R → ∞|φ0 , R0 )dφ
where P(φ, R|φ0 , R0 ) is the probability that a sector has angu- 0

lar size φ when the colony radius is R, given that its size was πσ 2 (R → ∞) 2πD
φ0 at R0. If a sector shrinks to angular size φ = 0 then we = = , (50)
2 R0
assume it cannot recover (since the lineage becomes lost 
behind the growing layer): this implies the boundary condition where we have used σ 2 (R → ∞) = 4D/R0 . The average
P(0, R|φ0 , R0 ) = 0 . We also set P(∞, R|φ0 , R0 ) = 0 because number of sectors is thus
sectors cannot become arbitrarily large8. With these boundary 
conditions the solution of equation (46) is 2π 2πR0
(51) Nsectors (R → ∞) = = .
(φ+φ0 )2  2φφ0  φ(R → ∞) D
1
(47) P(φ, R|φ0 , R0 ) = √ e− 2σ2 e σ2 − 1 , Interestingly, this theory predicts that the number of sec-
2πσ 2
tors in the large colony limit is finite, showing that coex-
where σ 2 (R) = 4D(R−1 istence between different lineages is possible. Note that
0 − R ). This result can be obtained
−1

using the method of images and taking a Fourier transform of Nsectors (R → ∞) is independent of the initial number of sec-
equation (46) [204]. For small initial sector size φ0  2π we tors, as long as this initial number is large (small φ0 ), but it
can expand equation (47) to first order in φ0 , depends on the initial radius R0 of the colony. Individual-based
  simulations of colony growth and experiments with bacteria

 φφ0 φ2 growing on agar plates confirm this prediction [74, 188, 192,
(48) P(φ, R|φ0 , R0 ) = 2/π 3 exp − 2 .
σ 2σ 193] and show that it remains qualitatively true if the growing
layer has finite thickness. Simulations and extensions of the
8 theory can also be used to predict what happens when the col-
Actually, φ cannot be larger than 2π but we expect most sectors to be
much smaller than this if the initial number of sectors is large. Assuming an ony contains mixtures of bacteria with different growth rates
absorbing boundary at φ → ∞ simplifies the calculations. [73, 188], when the bacteria are able to undergo horizontal

21
Rep. Prog. Phys. 82 (2019) 016601 Review

gene transfer between neighbouring cells [205, 206], or when From a biological point of view, the growing threat of
the growing population encounters obstacles [207]. antimicrobial resistance, increasing awareness of the role of
This theoretical analysis provides an example of the use biofilms in infection, and the growing understanding of the
of statistical physics to understand a complex biological importance of microbes in gut health have raised the profile
phenom­enon. However it does not explain what features of of microbiology in recent years. There has also been a resur-
the growth process control the diffusion constant D, which gence in interest among microbiologists in both fundamental
plays a critical role in determining the number of sectors. growth phenomena and in the use of mathematical models
Indeed, the number of sectors has been observed to differ to explain them. Thus, statistical physics models of bacterial
between different organisms: fewer sectors are observed growth have the potential to make a significant impact.
for the yeast S. cerevisiae than for E. coli, and also, intrigu- We would also like to highlight here the importance of
ingly, fewer sectors are observed for a spherical mutant of experiments. Many (although not all) of the bacterial growth
E. coli than for the usual rod-shaped E. coli cells [72, 73]. phenomena discussed in this review arise in rather simple micro-
Individual-based simulations have an important role to play biological experiments. Our own experience is that even a brief
in explaining these observations; these simulations have immersion into experimental work with bacteria can greatly
already pointed to mechanical interactions between cells and improve one’s ability to develop relevant, realistic and interest-
the surface on which they grow as major players in determin- ing statistical physics models. Moreover this is often a fun expe-
ing D [192] . rience! We therefore advocate spending some time in the lab to
Genetic segregation within an expanding bacterial popula- even the most hardened theoretician, if it is at all possible.
tion, as described in this example, has important evolutionary
implications, since it significantly affects the ‘surfing prob-
Acknowledgments
ability’, or the probability that a mutant arising at the front of
an expanding population forms a macroscopic sector [188].
The authors thank Lucas Black, Richard Blythe, Rebecca
This is relevant, for example, to the evolution of antibiotic
Brouwers, Timothy Bush, Martin Carballo Pacheco, Mike
resistance in bacterial biofilms. A similar problem arises in
Cates, Luca Ciandrini, Pietro Cicuta, Steven Court, Susana
the evolution of drug resistance in cancer tumours [17].
Direito, Martin Evans, Andrew Free, Philip Greulich, Oskar
Hallatschek, Bhavin Khatri, Elin Lilja, Diarmuid Lloyd, Cat
5.  Conclusions and outlook Mills, Nikola Ojkic, Eulyn Pagaling, Jakub Pastuszak, Wilson
Poon, Patrick Sinclair, Sharareh Tavaddod, Daniel Taylor,
The primary purpose of this review has been to illustrate the Simon Titmuss, Juan Venegas-Ortiz, Paolo Visco, Patrick
rich array of beautiful and interesting phenomena displayed by Warren and Ellen Young for many discussions. Special thanks
growing bacterial populations. These phenomena are intrinsi- are due to Bhavin Khatri, Joost Teixeira de Mattos, Alex ter
cally non-equilibrium and many of them lend themselves nat- Beek, Martijn Bekker, Tanneke den Blaauwen, Yasuhiko Irie
urally to analysis using the tools of statistical physics. and Diarmuid Lloyd, who provided original data or images
Research at the interface between microbiology and statis- for the figures  shown here. RJA was supported by a Royal
tical physics can have great benefits for both fields. Statistical Society University Research Fellowship, by the Human
physics models can cut through biological detail and pro- Frontier Science Program under grant RGY0081/2012, by the
vide insight into basic biological mechanisms, when they US Army Research Office under grant 64052-MA, by EPSRC
are properly constructed with knowledge of the underlying under grant EP/J007404/1, and by ERC Consolidator Grant
biology. The application of statistical physics to biological 682237-‘EVOSTRUC’. BW was funded by a Leverhulme
problems can also generate new non-equilibrium models that Trust Early Career Research Fellowship and by a Royal
drive further development in statistical physics. Fruitful inter- Society of Edinburgh Research Fellowship.
play between statistical physics and biology is nothing new:
examples include the totally asymmetric exclusion process
[11], which was introduced as a model for cellular protein ORCID iDs
production [12] and has since become a paradigm for non-
equilibrium transport processes. What we aim to highlight Rosalind J Allen https://orcid.org/0000-0003-4110-2962
Bartlomiej Waclaw https://orcid.org/0000-0001-5355-7994
here is the attractiveness of bacterial populations, specifically,
as subjects for statistical physics models. We believe that this
is a timely topic, from the point of view of both physics and References
biology. From a physics point of view, new non-equilibrium
physics is emerging from the study of active systems, which [1] World Health Organization 2014 Antimicrobial Resistance:
have up to now been mainly focused on motile particles Global Report on Surveillance (Geneva: World Health
(‘swimmers’) [25, 60, 208–210]. Yet bacteria are also active Organization)
in many other ways: they grow, divide, secrete signals and [2] World Health Organization 2001 WHO Global Strategy for
macromolecules, and interact chemically and mechanically Containment of Antimicrobial Resistance (Geneva: World
Health Organization)
in a complex way. The statistical physics of these behaviours, [3] Centers for Disease Control and Prevention (US) 2013
especially growth in dense assemblies of bacteria, has just Antibiotic Resistance Threats in the United States (Atlanta,
started to be explored. GA: Centers for Disease Control and Prevention)

22
Rep. Prog. Phys. 82 (2019) 016601 Review

[4] Davies S C and Gibbens N 2013 UK Five Year Antimicrobial [27] Schaechter M, Ingraham J L, Neidhardt F C, Schaechter M,
Resistance Strategy 2013–8 (London: Department of Health Ingraham J L and Neidhardt F C 2006 Microbe
and Social Care) (Washington, DC: ASM)
[5] O’Neill J 2014 Review on antimicrobial resistance. [28] Madigan M T, Martinko J M, Dunlap P V and Clark D P 2009
Antimicrobial resistance: tackling a crisis for the future Brock Biology of Microorganisms 12th edn (San Francisco,
health and wealth of nations CA: Pearson)
[6] Fenchel T, King G M and Blackburn T H 1998 Bacterial [29] de Visser J A G M and Krug J 2014 Empirical fitness
Biogeochemistry: the Ecophysiology of Mineral Cycling landscapes and the predictability of evolution Nat. Rev.
(Amsterdam: Elsevier) Genet. 15 480–90
[7] Flint H J, Scott K P, Louis P and Duncan S H 2012 The role [30] Lee H, Popodi E, Tang H and Foster P L 2012 Rate and
of the gut microbiota in nutrition and health Nat. Rev. molecular spectrum of spontaneous mutations in the
Gastroenterol. Hepatol. 9 577–89 bacterium Escherichia coli as determined by whole-genome
[8] Sevious P and Halkjaer Nielsen P 2010 Microbial Ecology of sequencing Proc. Natl Acad. Sci. USA 109 E2774–83
Activated Sludge (London: IWA) [31] Kovarova-Kovar K and Egli T 1998 Growth kinetics of
[9] Luria S E and Delbrück M 1943 Mutations of bacteria from suspended microbial cells: from single-substrate-controlled
virus sensitivity to virus resistance Genetics 28 491–511 growth to mixed-substrate kinetics Microbiol. Mol. Biol.
[10] Lea D E and Coulson C A 1949 The distribution of the Rev. 62 646
numbers of mutants in bacterial populations J. Genet. [32] Neidhardt F C (ed) 1987 Escherichia coli and Salmonella
49 264 typhimurium: Cellular and Molecular Biology (Washington,
[11] Blythe R A and Evans M R 2007 Nonequilibrium steady states DC: ASM)
of matrix-product form: a solver’s guide J. Phys. A: Math. [33] Nordström K and Dasgupta S 2006 Copy-number control of
Theor. 40 R333–441 the Escherichia coli chromosome: a plasmidologist’s view
[12] MacDonald C, Gibbs J and Pipkin A 1968 Kinetics of EMBO Rep. 7 484–9
biopolymerization on nucleic acid templates Biopolymers [34] Bremer H and Dennis P 1996 Modulation of Chemical
6 1 Composition and Other Parameters of the Cell by Growth
[13] Elowitz M B, Levine A J, Siggia E D and Swain P S 2002 Rate in Escherichia coli and Salmonella typhimurium:
Stochastic gene expression in a single cell Science Cellular and Molecular Biology, F C Neidhardt (Ed)
297 1183–6 (Washington DC: ASM)
[14] Thattai M and van Oudenaarden A 2001 Intrinsic noise [35] Bennett B D, Kimball E H, Gao M, Osterhout R, van
in gene regulatory networks Proc. Natl Acad. Sci. USA Dien S J and Rabinowitz J D 2009 Absolute metabolite
98 8614–9 concentrations and implied enzyme active site occupancy in
[15] Eden M 1961 A two-dimensional growth process Proc. 4th Escherichia coli Nat. Chem. Biol. 5 593–9
Berkeley Symp. on Mathematical Statistics and Probability [36] BioNumbers—the database of useful biological numbers
vol 4 pp 223–39 (bionumbers.hms.harvard.edu/)
[16] Court S J, Blythe R A and Allen R J 2013 Parasites on [37] Hartmann M, Berditsch M, Hawecker J, Fotouhi Ardakanai M,
parasites: coupled fluctuations in stacked contact processes Gerthsen D and Ulrich A S 2010 Damage of the bacterial
Europhys. Lett. 101 50001 cell envelope by antimicrobial peptides gramicidin S
[17] Waclaw B, Bozic I, Pittman M E, Hruban R H, Vogelstein B and PGLa as revealed by transmission and scanning
and Nowak M A 2015 A spatial model predicts that electron microscopy Antimicrobial Agents Chemotherapy
dispersal and cell turnover limit intratumour heterogeneity 54 3132–42
Nature 525 261–4 [38] Greenwood D and O’Grady F 1972 Scanning electron microscopy
[18] Vicsek T, Czirók A, Ben-Jacob E, Cohen I and Shochet O of Staphylococcus aureus exposed to some common anti-
1995 Novel type of phase transition in a system of self- staphylococcal agents J. Gen. Microbiol. 70 263–70
driven particles Phys. Rev. Lett. 75 1226–9 [39] Wu F, Zhang J-P and Wang Q-Q 2017 Scanning electron
[19] Toner J and Tu Y 1995 Long-range order in a two-dimensional microscopy of the adhesion of Treponema pallidum
dynamical XY model: how birds fly together Phys. Rev. subspecies pallidum (Nichol strain) to human brain
Lett. 75 4326–9 microvascular endothelial cells in vitro JEADV 31 e222
[20] Grégoire G and Chaté H 2004 Onset of collective and cohesive [40] Grossman N, Ron E Z and Woldringh C L 1982 Changes in
motion Phys. Rev. Lett. 92 025702 cell dimensions during amino acid starvation of Escherichia
[21] Fily Y and Marchetti M C 2012 Athermal phase separation of coli J. Bacteriol. 152 35–41
self-propelled particles with no alignment Phys. Rev. Lett. [41] Nelson D E and Young K D 2000 Penicillin binding protein
108 235702 5 affects cell diameter, contour, and morphology of
[22] Buttinoni I, Bialke J, Kummel F, Löwen H, Bechinger C and Escherichia coli J. Bacteriol. 182 1714–21
Speck T 2013 Dynamical clustering and phase separation in [42] Donachie W D, Begg K J and Vicente M 1976 Cell length, cell
suspensions of self-propelled colloidal particles Phys. Rev. growth and cell division Nature 264 328–33
Lett. 110 238301 [43] Donachie W D and Blakely G W 2003 Coupling the initiation
[23] Marchetti M C, Joanny J F, Ramaswamy S, Liverpool T B, of chromosome replication to cell size in Escherichia coli
Prost J, Rao M and Simha R A 2013 Hydrodynamics of soft Curr. Opin. Microbiol. 6 146–50
active matter Rev. Mod. Phys. 85 1143 [44] Osella M, Tans S J and Cosentino Lagomarsino M 2017 Step
[24] Stenhammar J, Tiribocchi A, Allen R J, Marenduzzo D and by step, cell by cell: quantification of the bacterial cell cycle
Cates M E 2013 Continuum theory of phase separation Trends Microbiol. 25 250–6
kinetics for active Brownian particles Phys. Rev. Lett. [45] Jun S and Taheri-Araghi S 2015 Cell-size maintenance:
111 145702 universal strategy revealed Trends Microbiol. 23 4–6
[25] Cates M E and Tailleur J 2015 Motility-induced phase [46] Harris L K and Theriot J A 2016 Relative rates of surface and
separation Annu. Rev. Condens. Matter Phys. volume synthesis set bacterial cell size Cell 165 1479–92
6 219–44 [47] Wallden M, Fange D, Gregorsson Lundius E, Baltekin Ö and
[26] Ingraham J L, Maaloe O and Neidhardt F C 1983 Growth of Elf J 2016 The synchronization of replication and division
the Bacterial Cell (Sunderland, MA: Sinauer Associates) cycles in individual E. coli cells Cell 166 729–39

23
Rep. Prog. Phys. 82 (2019) 016601 Review

[48] Alon U 2007 An Introduction to Systems Biology: Design [72] Hallatschek O, Hersen P, Ramanathan S and Nelson D R
Principles of Biological Circuits (London: Chapman and 2007 Genetic drift at expanding frontiers promotes gene
Hall) segregation Proc. Natl Acad. Sci. USA 104 19926–30
[49] Gardner T S, Cantor C R and Collins J J 2000 Construction [73] Hallatschek O and Nelson D R 2010 Life at the front of an
of a genetic toggle switch in Escherichia coli Nature expanding population Evolution 64 193–206
403 339–42 [74] Ali A, Somfai E and Grosskinsky S 2012 Reproduction-time
[50] Ozbudak E M, Thattai M, Lim H N, Shraiman B I and van statistics and segregation patterns in growing populations
Oudenaarden A 2004 Multistability in the lactose utilization Phys. Rev. E 85 021923
network of Escherichia coli Nature 427 737–40 [75] Drescher K, Dunkel J, Nadell C D, van Teefelen S, Grnja I,
[51] Warren P B and ten Wolde P R 2005 Chemical models of Wingreen N S, Stone H A and Bassler B L 2016
genetic toggle switches J. Phys. Chem. B 109 6812–23 Architectural transitions in Vibrio cholerae biofilms at single-
[52] Morelli M J, Tanase-Nicola S, Allen R J and ten Wolde P R cell resolution Proc. Natl Acad. Sci. USA 113 E2066–72
2008 Reaction coordinates for the flipping of genetic [76] Lloyd D P and Allen R J 2015 Competition for space during
switches Biophys. J. 94 3413–23 bacterial colonization of a surface J. R. Soc. Interface
[53] Morelli M J, ten Wolde P R and Allen R J 2009 DNA looping 12 20150608
provides stability and robustness to the bacteriophage [77] Volfson D, Cookson S, Hasty J and Tsimring L S 2008
lambda switch Proc. Natl Acad. Sci. USA 106 8101–6 Biomechanical ordering of dense cell populations Proc.
[54] Venegas-Ortiz J and Evans M R 2011 Analytical study of an Natl Acad. Sci. USA 105 15346
exclusive genetic switch J. Phys. A: Math. Theor. 44 355001 [78] Grant M A A, Waclaw B, Allen R J and Cicuta P 2014 The
[55] Visco P, Allen R J and Evans M R 2008 Exact solution of a role of mechanical forces in the planar-to-bulk transition in
model DNA-inversion genetic switch with orientational growing Escherichia coli microcolonies J. R. Soc. Interface
control Phys. Rev. Lett. 101 118104 11 20140400
[56] Visco P, Allen R J and Evans M R 2009 Statistical physics of a [79] West S A, Griffin A S, Gardner A and Diggle S P 2006 Social
model binary genetic switch with linear feedback Phys. Rev. evolution theory for microorganisms Nat. Rev. Microbiol.
E 79 031923 4 597–607
[57] Harshey R M 2003 Bacterial motility on a surface: many ways [80] Elena S F and Lenski R E 2003 Evolution experiments
to a common goal Annu. Rev. Microbiol. 57 249–73 with microorganisms: the dynamics and genetic bases of
[58] Michell J G and Kogure K 2006 Bacterial motility: links to adaptation Nat. Rev. Genet. 3 457–69
the environment and a driving force for microbial physics [81] Allen R J and Waclaw B 2016 Antibiotic resistance: a
FEMS Microbiol. Ecol. 55 3–16 physicist’s view Phys. Biol. 13 045001
[59] Berg H C 1993 Random Walks in Biology (Princeton, NJ: [82] Volkmer B and Heinemann M 2011 Condition-dependent cell
Princeton University Press) volume and concentration of Escherichia coli to facilitate
[60] Cates M E 2012 Diffusive transport without detailed balance data conversion for systems biology modeling PLoS One 6
in motile bacteria: does microbiology need statistical e23126
physics? Rep. Prog. Phys. 75 042601 [83] Rolfe M D et al 2012 Lag phase is a distinct growth phase
[61] Lauga E and Powers T R 2009 The hydrodynamics of that prepares bacteria for exponential growth and involves
swimming microorganisms Rep. Prog. Phys. 72 096601 transient metal accumulation J. Bacteriol. 194 686–701
[62] Howse J R, Jones R A L, Ryan A J, Gough T, Vafabakhsh R [84] Kolter R, Siegele D A and Tormo A 1993 The stationary phase
and Golestanian R 2007 Self-motile colloidal particles: of the bacterial life cycle Annu. Rev. Microbiol. 47 855–74
from directed propulsion to random walk Phys. Rev. Lett. [85] Schaechter M, Williamson J P, Hood J R and Koch A L 1962
99 048102 Growth, cell and nuclear divisions in some bacteria J. Gen.
[63] Schwarz-Linek J, Arlt J, Jepson A, Dawson A, Vissers T, Microbiol. 29 421
Miroli D, Pilizota T, Martinez V A and Poon W C K 2016 [86] Murray J D 2002 Mathematical Biology (Berlin: Springer)
Escherichia coli as a model active colloid: a practical [87] Egli T 1991 On multiple-nutrient-limited growth of
introduction Colloid Surf. B 137 2–16 microorganisms, with special reference to dual limitation by
[64] Valeriani C, Allen R J and Marenduzzo D 2010 Non- carbon and nitrogen substrates Antonie van Leeuwenhoek
equilibrium dynamics of an active colloidal ‘chucker’ 60 225–34
J. Chem. Phys. 132 204904 [88] Baranyi J, McClure P J, Sutherland J P and Roberts T A 1993
[65] Kiviet D, Nghe P, Walker N, Boulineau S, Sunderlikova V and Modeling bacterial growth responses J. Ind. Microbiol.
Tans S J 2014 Stochasticity of metabolism and growth at 12 190–4
the single-cell level Nature 514 376 [89] Buchanan R L, Whiting R C and Damert W C 1997 When
[66] Kennard A S, Osella M, Javer A, Grilli J, Nghe P, Tans S J, is simple good enough: a comparison of the Gompertz,
Cicuta P and Cosentino Lagomarsino M 2016 Individuality Baranyi, and three-phase linear models for fitting bacterial
and universality in the growth-division laws of single growth curves Food Microbiol. 14 313–26
E. coli cells Phys. Rev. E 93 012408 [90] Pruitt K M and Kamau D N 1993 Mathematical models of
[67] Hoffman H and Frank M E 1965 Synchrony of division in bacterial growth, inhibition and death under combined
clonal microcolonies of Escherichia coli J. Bacteriol. stress conditions J. Ind. Microbiol. 12 221–31
89 513–7 [91] Zwietering M H, Jongenburger I, Rombouts F M and Van’t
[68] Cutler R G and Evans J E 1966 Synchronization of bacteria by Riet K 1990 Modeling of the bacterial growth curve Appl.
a stationary-phase method J. Bacteriol. 91 469–76 Environ. Microbiol. 56 1875–81
[69] Balaban N Q, Merrin J, Chait R, Kowalik L and Leibler S [92] Monod J 1949 The growth of bacterial cultures Annu. Rev.
2004 Bacterial persistence as a phenotypic switch Science Microbiol. 3 371–94
305 1622–5 [93] Contois D E 1959 Kinetics of bacterial growth: relationship
[70] Balaban N Q, Gerdes K, Lewis K and McKinney J D 2013 A between population density and specific growth rate of
problem of persistence: still more questions than answers? continuous cultures J. Gen. Microbiol. 21 40–50
Nat. Rev. Microbiol. 11 587–91 [94] Ratkowsky D A, Lowry R K, McMeekin T A, Stokes A N and
[71] Ghosh P, Mondal J, Ben-Jacob E and Levine H 2015 Chandler R E 1983 Model for bacterial culture growth rate
Mechanically-driven phase separation in a growing throughout the entire biokinetic temperature range
bacterial colony Proc. Natl Acad. Sci. USA 112 E2166—73 J. Bacteriol. 154 1222–6

24
Rep. Prog. Phys. 82 (2019) 016601 Review

[95] Kompala D S, Ramkrishna D, Jansen N B and Tsao G T 1986 [118] McKane A J and Newman T J 2005 Predator-prey cycles
Investigation of bacterial growth on mixed substrates: from resonant amplification of demographic stochasticity
experimental evaluation of cybernetic models Biotechnol. Phys. Rev. Lett. 94 218102
Bioeng. 28 1044–55 [119] Lugo C A and McKane A J 2008 Quasicycles in a spatial
[96] Presser K A, Ratkowsky D A and Ross T 1997 Modelling predator-prey model Phys. Rev. E 78 051911
the growth rate of Escherichia coli as a function of pH [120] Boland R P, Galla T and McKane A J 2008 How limit cycles
and lactic acid concentration Appl. Environ. Microbiol. and quasi-cycles are related in systems with intrinsic noise
63 2355–60 J. Stat. Mech. P09001
[97] Jin Q and Bethke C M 2007 The thermodynamics and [121] Ghose S and Adhikari R 2010 Endogenous quasicycles
kinetics of microbial metabolism Am. J. Sci. 307 643–77 and stochastic coherence in a closed epidemic model
[98] Smith H L and Waltman P E 1995 The Theory of the Phys. Rev. E 82 021913
Chemostat: Dynamics of Microbial Competition [122] Galla T 2009 Intrinsic noise in game dynamical learning
(Cambridge: Cambridge University Press) Phys. Rev. Lett. 103 198702
[99] Ferenci T 2008 Bacterial physiology, regulation and [123] Bladon A J, Galla T and McKane A J 2010 Evolutionary
mutational adaptation in a chemostat environment dynamics, intrinsic noise and cycles of cooperation
Adv. Microbial Physiol. 53 169 Phys. Rev. E 81 066122
[100] Butler G J and Waltman P 1981 Bifurcation from a limit [124] Alonso D, McKane A J and Pascual M 2007 Stochastic
cycle in a 2 predator one prey ecosystem modeled on a amplification in epidemics J. R. Soc. Interface 4 575–82
chemostat J. Math. Biol. 12 295–310 [125] McKane A J, Nagy J D, Newman T J and Stefanini M O
[101] Bohannan B J M and Lenski R E 1997 Effect of resource 2007 Amplified biochemical oscillations in cellular
enrichment on a chemostat community of bacteria and systems J. Stat. Phys. 128 165–91
bacteriophage Ecology 73 2303–15 [126] Dauxois T, Di Patti F, Fanelli D and McKane A J 2009
[102] Khatri B, Free A and Allen R J 2012 Oscillating microbial Enhanced stochastic oscillations in autocatalytic systems
dynamics driven by small populations, limited nutrient Phys. Rev. E 79 036112
supply and high death rates J. Theor. Biol. 314 120–9 [127] Jakiela S, Kaminski T S, Cybulski O, Weibel D B and
[103] Xia H, Wolkowicz G S K and Wang L 2005 Transient Garstecki P 2013 Bacterial growth and adaptation
oscillations induced by delayed growth response in the in microdroplet chemostats Angew. Chem. Int. Ed.
chemostat J. Math. Biol. 50 489–530 52 8908–11
[104] Wang P, Robert L, Pelletier J, Dang W L, Taddie F, Wright A [128] Evans C G T, Herbert D and Tempest D W 1970 The
and Jun S 2010 Robust growth of Escherichia coli Curr. continuous culture of microorganisms. 2. Construction of
Biol. 20 1099–103 a chemostat Methods in microbiology vol 2, ed J R Norris
[105] Keymer J E, Galajda P, Muldoon C, Park S and Austin R E and D W Ribbons (New York: Academic) pp 277–327
2006 Bacterial metapopulations in nanofabricated [129] van der Woude M W and Bäumler A 2004 Phase and antigenic
landscapes Proc. Natl Acad. Sci. USA 103 17290–5 variation in bacteria Clin. Microbiol. Rev. 17 581–611
[106] Balagadde F K, You L C, Hansen C L, Arnold F H and [130] Henderson I R, Owen P and Nataro J P 1999 Molecular
Quake S R 2005 Long-term monitoring of bacteria switches—the on and off of bacterial phase variation
undergoing programmed population control in a Mol. Microbiol. 33 919–32
microchemostat Science 309 137–40 [131] Ackermann M 2015 A functional perspective on phenotypic
[107] Long Z, Nugent E, Javer A, Cicuta P, Sclavi B, heterogeneity in microorganisms Nat. Rev. Microbiol.
Lagomarsino M C and Dorfman D K 2013 Microfluidic 13 497–508
chemostat for measuring single cell dynamics in bacteria [132] Hallet B 2001 Playing Dr Jekyll and Mr Hyde: combined
Lab Chip 13 947–54 mechanisms of phase variation in bacteria Curr. Opin.
[108] Athreya K B and Ney P E 2004 Branching Processes (North Microbiol. 4 570–81
Chelmsford, MA: Courier) [133] Visco P, Allen R J, Majumdar S N and Evans M R 2010
[109] Nowak M A 2006 Evolutionary Dynamics (Cambridge, MA: Switching and growth for microbial populations in
Harvard University Press) catastrophic responsive environments Biophys. J.
[110] Durrett R 2015 Branching process models of cancer 98 1099–108
Branching Process Models of Cancer (Mathematical [134] Tadrowski A C, Evans M R and Waclaw B 2018 Phenotypic
Biosciences Institute Lecture Series) (Cham: Springer) switching can speed up biological evolution of microbes
pp 1–63 Sci. Rep. 8 8941
[111] Nicholson M D and Antal T 2016 Universal asymptotic clone [135] Diard M, Garcia V, Maier L, Remus-Emsermann M N P,
size distribution for general population growth Bull. Math. Regoes R, Ackermann M and Hardt W-D 2013
Biol. 78 2243–76 Stabilization of cooperative virulence by the expression of
[112] Strogatz S H 1994 Nonlinear Dynamics and Chaos (Boulder, an avirulent phenotype Nature 494 353–6
CO: Westview) [136] Arnoldini M, Avalos Vizcarra I, Peña-Miller R, Stocker N,
[113] Cohen S S and Barner H D 1954 Studies on unbalanced Diard M, Vogel V, Beardmore R E, Hardt W-D and
growth in Escherichia coli Proc. Natl Acad. Sci. USA Ackermann M 2014 Bistable expression of virulence
40 885–93 genes in salmonella leads to the formation of an antibiotic-
[114] Gillespie D T 1976 A general method for numerically tolerant subpopulation PLoS Biol. 12 e1001928
simulating the stochastic time evolution of coupled [137] Seger J and Brockman H J 1987 What is bet-hedging?
chemical reactions J. Comput. Phys. 22 403–34 Oxford Surveys in Evolutionary Biology (Oxford: Oxford
[115] Gillespie D T 1977 Exact stochastic simulation of coupled University Press)
chemical reactions J. Phys. Chem. 81 2340–61 [138] Lachman M and Jablonka E 1996 The inheritance of
[116] van Kampen N G 1981 Stochastic Processes in Physics and phenotypes: an adaptation to fluctuating environment
Chemistry (Amsterdam: North-Holland) J. Theor. Biol. 181 1
[117] Lu T, Volfson D, Tsimring L and Hasty J 2004 Cellular [139] Ishii K, Matsuda H, Iwasa Y and Sasaki A 1989
growth and division in the gillespie algorithm Syst. Biol. Evolutionarily stable mutation rate in a periodically
1 121–8 changing environment Genetics 121 163

25
Rep. Prog. Phys. 82 (2019) 016601 Review

[140] Thattai M and van Oudenaarden A 2004 Stochastic gene [165] Fasham M J R, Ducklow H W and McKelvie S M 1990 A
expression in fluctuating environments Genetics 167 523 nitrogen-based model of plankton dynamics in the ocean
[141] Gander M J, Mazza C and Rummler H 2007 Stochastic gene mixed layer J. Mar. Res. 48 591–639
expression in switching environments J. Math. Biol. 55 259 [166] Rang H P, Dale M M, Ritter J M and Moore P K 2003
[142] Ribeiro A S 2008 Dynamics and evolution of stochastic Pharmacology 5th edn (London: Churchill Livingstone)
bistable gene networks with sensing in fluctuating [167] Hallatschek O and Korolev K S 2009 Fisher waves in the
environments Phys. Rev. E 78 061902 strong noise limit Phys. Rev. Lett. 103 108103
[143] Wolf D M, Vazirani V V and Arkin A P 2005 Diversity in [168] Conclon J G and Doering C R 2005 On travelling waves for
times of adversity: probabilistic strategies in microbial the stochastic Fisher–Kolmogorov–Petrovsky–Piscunov
survival games J. Theor. Biol. 234 227 equation J. Stat. Phys. 120 421–77
[144] Kussell E and Leibler S 2005 Phenotypic diversity, [169] Hallatschek O 2011 The noisy edge of traveling waves
population growth, and information in fluctuating Proc. Natl Acad. Sci. USA 108 1783–7
environments Science 309 2075 [170] Venegas-Ortiz J, Allen R J and Evans M R 2014 Speed of
[145] Kussell E, Kishony R, Balaban N Q and Leibler S 2005 invasion of an expanding population by a horizontally
Bacterial persistence: a model of survival in changing transmitted trait Genetics 196 497–507
environments Genetics 169 1807 [171] Hermsen R, Deris J B and Hwa T 2012 On the rapidity
[146] Acar M, Mettetal J T and van Oudenaarden A 2008 of antibiotic resistance evolution facilitated by a
Stochastic switching as a survival strategy in fluctuating concentration gradient Proc. Natl Acad. Sci. USA
environments Nat. Genet. 40 471 109 10775–80
[147] Stewart P S 2003 Diffusion in biofilms J. Bacteriol. [172] Greulich P, Waclaw B and Allen R J 2012 Mutational
185 1485–91 pathway determines whether drug gradients accelerate
[148] Baquero F and Negri M C 1997 Challenges: selective evolution of drug-resistant cells Phys. Rev. Lett.
compartments for resistant microorganisms in antibiotic 109 088101
gradients Bioessays 19 731–6 [173] Kragh K N et al 2016 Role of multicellular aggregates in
[149] Costerton J W, Lewandowski Z, Caldwell D E, Korber D R biofilm formation mBio 7 e00237
and Lappin-Scott H M 1995 Microbial biofilms Annu. Rev. [174] Klapper I and Dockery J 2010 Mathematical description of
Microbiol. 49 711–45 microbial biofilms SIAM Rev. 52 221–65
[150] Watnick P and Kolter R 2000 Biofilm, city of microbes [175] Klapper I and Dockery J 2002 Finger formation in biofilm
J. Bacteriol. 182 2675–9 layers SIAM J. Appl. Math. 62 853–69
[151] Parsek M R and Singh P K 2003 Bacterial biofilms: an [176] Tolker-Nielsen T and Sternberg C 2014 Methods for studying
emerging link to disease pathogenesis Annu. Rev. biofilm formation: flow cells and confocal laser scanning
Microbiol. 57 677–701 microscopy Methods Mol. Biol. 1149 615–29
[152] Bixler G D and Bhushan B 2012 Biofouling: lessons from [177] Klausen M, Aaes-Jorgensen A, Molin S and Tolker-
nature Phil. Trans. R. Soc. A 1967 2381–417 Nielsen T 2003 Involvement of bacterial migration in
[153] Ben-Jacob E, Cohen I and Levine H 2000 Cooperative self- the development of complex multicellular structures
organization of microorganisms Adv. Phys. 49 395–554 in Pseudomonas aeruginosa biofilms Mol. Microbiol.
[154] Ben-Jacob E, Cohen I and Gutnick D L 1998 Cooperative 50 61–8
organization of bacterial colonies: from genotype to [178] Giverso C, Verani M and Ciarletta P 2016 Emerging
morphotype Annu. Rev. Microbiol. 52 779–806 morphologies in round bacterial colonies: comparing
[155] Dockery J and Kapper I 2001 Finger formation in biofilm volumetric versus chemotactic expansion Biomech. Model.
layers SIAM J. Appl. Math. 62 853–69 mechanobiol. 15 643–61
[156] Nadell C D, Bucci V, Drescher K, Levin S A, Bassler B L [179] Wang X, Stone H A and Golestanian R 2017 Shape of the
and Xavier J B 2013 Cutting through the complexity of growing front of biofilms New J. Phys. 19 125007
cell collectives Proc. R. Soc. B 280 20122770 [180] Doostmohammadi A, Thampi S P and Yeomans J M 2016
[157] Melaugh G, Hutchison J, Kragh K N, Irie Y, Roberts A, Defect-mediated morphologies in growing cell colonies
Bjarnsholt T, Diggle S P, Gordon V D and Allen R J 2016 Phys. Rev. Lett. 117 168101
Shaping the growth behaviour of biofilms initiated from [181] You Z, Pearce D J G, Sengupta A and Giomi L 2018
bacterial aggregates PLoS One 11 e0149683 Geometry and mechanics of micro-domains in growing
[158] Bonachela J A, Nadell C D, Xavier J B and Levin S A 2011 bacterial colonies Phys. Rev. X. 8 031065
Universality in bacterial colonies J. Stat. Phys. 144 303–15 [182] Plischke M and Rácz Z 1985 Dynamic scaling and the
[159] Krug J 1997 Origins of scale invariance in growth processes surface structure of Eden clusters Phys. Rev. A
Adv. Phys. 46 139–282 32 3825–8
[160] Maunuksela J, Myllys M, Kahkonen O-P, Timonen J, [183] Kertesz J and Wolf D E 1988 Noise reduction in Eden
Provatas N, Alava M J and Ala-Nissila T 1997 Kinetic models: II. Surface structure and intrinsic width J. Phys.
roughening in slow combustion of paper Phys. Rev. Lett. A: Math. Gen. 21 747
79 1515–8 [184] Alves S G and Ferreira S C 2012 Eden clusters in three
[161] Takeuchi K A and Sano M 2010 Universal fluctuations of dimensions and the Kardar–Parisi–Zhang universality
growing interfaces: evidence in turbulent liquid crystals class J. Stat. Mech. P10011
Phys. Rev. Lett. 104 230601 [185] Corwin I 2012 The Kardar–Parisi–Zhang equation and
[162] Takeuchi K A, Sano M, Sasamoto T and Spohn H 2011 universality class Random Matrices 1 1130001
Growing interfaces uncover universal fluctuations behind [186] Vicsek T, Cserz M and Horváth V K 1990 Self-affine growth
scale invariance Sci. Rep. 1 34 of bacterial colonies Physica A 167 315–21
[163] Takeuchi K A and Sano M 2012 Evidence for geometry- [187] Wakita J, Itoh H, Matsuyama T and Matsushita M 1997 Self-
dependent universal fluctuations of the Kardar–Parisi– affinity for the growing interface of bacterial colonies
Zhang interfaces in liquid-crystal turbulence J. Stat. Phys. J. Phys. Soc. Japan 66 67–72
147 853–90 [188] Gralka M, Stiewe F, Farrell F, Moebius W, Waclaw B
[164] Farrell F D C, Hallatschek O, Marenduzzo D and Waclaw B and Hallatschek O 2016 Allele surfing promotes
2013 Mechanically driven growth of quasi-two- microbial adaptation from standing variation Ecol. Lett.
dimensional microbial colonies Phys. Rev. Lett. 111 19 889–98

26
Rep. Prog. Phys. 82 (2019) 016601 Review

[189] Ohgiwara M, Matsushita M and Matsuyama T 1992 [201] Korolev K S, Xaview J D, Nelson D R and Foster K R
Morphological changes in growth phenomena of bacterial 2011 A quantitative test of population genetics using
colony patterns J. Phys. Soc. Japan 61 816–22 spatiogenetic patterns in bacterial colonies Am. Nat.
[190] Fugikawa H and Matsushita M 1989 Fractal growth of Bacillus 178 538–52
subtilis on agar plates J. Phys. Soc. Japan 58 3875–8 [202] Korolev K S, Müller M J I, Karahan N, Murray A W,
[191] Ruzicka M C, Fridrich M and Burkhard M 1995 A bacterial Hallatschek O and Nelson D R 2012 Selective
colony is not self-similar Physica A 216 382–5 sweeps in growing microbial colonies Phys. Biol.
[192] Farrell F D, Gralka M, Hallatschek O and Waclaw B 2017 9 026008
Mechanical interactions in bacterial colonies and the [203] Gardiner C 2009 Stochastic Methods: a Handbook for
surfing probability of beneficial mutations J. R. Soc. the Natural and Social Sciences (Springer Series in
Interface 14 20170073 Synergetics) 4th edn (Berlin: Springer)
[193] Rudge T J, Steiner P J, Phillips A and Haseloff J 2012 [204] Redner S 2001 A Guide to First-passage Processes
Computational modeling of synthetic microbial biofilms (Cambridge: Cambridge University Press)
ACS Synth. Biol. 1 345–52 [205] Pastuszak J and Waclaw B 2017 A physics-explicit model
[194] Storck T, Picioreanu C, Virdis B and Batstone D J 2014 of plasmid conjugation in an expanding bacterial colony
Variable cell morphology approach for individual-based (arXiv:1712.01956)
modeling of microbial communities Biophys. J. 106 [206] Freese P D, Korolev K S, Jimenez J I and Chen I A 2014
2037–48 Genetic drift suppresses bacterial conjugation in spatially
[195] Head D A 2013 Linear surface roughness growth and flow structured populations Biophys. J. 106 944–54
smoothing in a three-dimensional biofilm model Phys. [207] Moebius W, Murray A W and Nelson D R 2015 How
Rev. E 88 032702 obstacles perturb population fronts and alter their genetic
[196] Kearns D B 2010 A field guide to bacterial swarming motility structure PLoS Comput. Biol. 11 e1004615
Nat. Rev. Microbiol. 8 634–44 [208] Najafi A and Golestanian R 2004 Simple swimmer at low
[197] Liotta L A and Kohn E C 2001 The microenvironment of the Reynolds number: three linked spheres Phys. Rev. E
tumour-host interface Nature 411 375–9 69 062901
[198] Wolpert L, Beddington R, Brockes J, Jessel T, Lawrence P [209] Golestanian R, Liverpool T B and Adjari A 2007
and Meyerowitz E (ed) 1998 Principles of Development Designing phoretic micro- and nano-swimmers New J.
(Oxford: Oxford University Press) Phys. 9 126
[199] Su P-T, Liao C, Roan J-R, Wang S-H, Chiou A and Syu W-J [210] Pooley C M, Alexander G P and Yeomans J M 2007
2012 Bacterial colony from two-dimensional division to Hydrodynamic interaction between two swimmers at low
three-dimensional development PLoS One 7 e48098 Reynolds number Phys. Rev. Lett. 99 228103
[200] Beroz F, Yan J, Meir Y, Sabass B, Stone H A, Bassler B L [211] Watson C, Hush P, Williams J, Dawson A, Ojkic N,
and Wingreen N S 2018 Verticalization of bacterial Titmuss S and Waclaw B 2018 Europhys. Lett.
biofilms Nat. Phys. 14 954–60 123 68001

Rosalind J. Allen is a professor of biological physics at Edinburgh University, where she has worked since 2006. She obtained
her PhD in theoretical chemistry from Cambridge University in 2003 under the supervision of Professor Jean-Pierre Hansen, and
was a postdoc in the Biochemical Networks group of Professor Pieter Rein ten Wolde at AMOLF in Amsterdam from 2003-6.
She uses a combination of lab experiments, computer simulations and theory to understand how microbes grow, interact with
their environment and with each other, respond to antibiotics and evolve antibiotic resistance. URL: https://www2.ph.ed.
ac.uk/~rallen2/.

Bartłomiej Waclaw is a reader in the School of Physics & Astronomy, Edinburgh University. He received a PhD in theoretical
physics in 2007 from Jagellonian University in Krakow. During and shortly after his PhD he worked on statistical physics
problems: random matrix theory, complex networks, random walks, and non-equilibrium statistical mechanics models. Since his
arrival in Edinburgh in 2009 he has been working on applications of statistical and soft matter physics to biological evolution:
non-equilibrium phase transitions in models of growing bacterial populations, the structure of fitness landscapes, genetic
heterogeneity in cancer, the evolution of antibiotic resistance in bacteria, and the growth of bacterial colonies. BW’s recent
research interests include the interplay between molecular mechanisms by which antibiotics kill bacteria and antibiotic resistance
spreading in bacterial populations, experimental and theoretical models of bacterial infections, and computer models of cancer
progression and treatment. In his research he uses a combination of experiments, computer modelling, and analytical calculations. URL: https://
bartekwaclaw.wordpress.com/.

27

You might also like