ArticlePDF Available

Quantification of Motility in Bacillus subtilis at Temperatures Up to 84°C Using a Submersible Volumetric Microscope and Automated Tracking

Frontiers
Frontiers in Microbiology
Authors:

Abstract and Figures

We describe a system for high-temperature investigations of bacterial motility using a digital holographic microscope completely submerged in heated water. Temperatures above 90°C could be achieved, with a constant 5°C offset between the sample temperature and the surrounding water bath. Using this system, we observed active motility in Bacillus subtilis up to 66°C. As temperatures rose, most cells became immobilized on the surface, but a fraction of cells remained highly motile at distances of >100 μm above the surface. Suspended non-motile cells showed Brownian motion that scaled consistently with temperature and viscosity. A novel open-source automated tracking package was used to obtain 2D tracks of motile cells and quantify motility parameters, showing that swimming speed increased with temperature until ∼40°C, then plateaued. These findings are consistent with the observed heterogeneity of B. subtilis populations, and represent the highest reported temperature for swimming in this species. This technique is a simple, low-cost method for quantifying motility at high temperatures and could be useful for investigation of many different cell types, including thermophilic archaea.
Content may be subject to copyright.
fmicb-13-836808 April 15, 2022 Time: 9:5 # 1
METHODS
published: 21 April 2022
doi: 10.3389/fmicb.2022.836808
Edited by:
Iain G. Duggin,
University of Technology Sydney,
Australia
Reviewed by:
Alex Bisson,
Brandeis University, United States
Georgia Squyres,
California Institute of Technology,
United States
*Correspondence:
Jay L. Nadeau
nadeau@pdx.edu
Specialty section:
This article was submitted to
Extreme Microbiology,
a section of the journal
Frontiers in Microbiology
Received: 16 December 2021
Accepted: 10 March 2022
Published: 21 April 2022
Citation:
Dubay MM, Johnston N,
Wronkiewicz M, Lee J,
Lindensmith CA and Nadeau JL
(2022) Quantification of Motility in
Bacillus subtilis at Temperatures Up to
84C Using a Submersible Volumetric
Microscope and Automated Tracking.
Front. Microbiol. 13:836808.
doi: 10.3389/fmicb.2022.836808
Quantification of Motility in Bacillus
subtilis at Temperatures Up to 84C
Using a Submersible Volumetric
Microscope and Automated Tracking
Megan M. Dubay1, Nikki Johnston1, Mark Wronkiewicz2, Jake Lee2,
Christian A. Lindensmith2and Jay L. Nadeau1*
1Department of Physics, Portland State University, Portland, OR, United States, 2Jet Propulsion Laboratory, California
Institute of Technology, Pasadena, CA, United States
We describe a system for high-temperature investigations of bacterial motility using a
digital holographic microscope completely submerged in heated water. Temperatures
above 90C could be achieved, with a constant 5C offset between the sample
temperature and the surrounding water bath. Using this system, we observed active
motility in Bacillus subtilis up to 66C. As temperatures rose, most cells became
immobilized on the surface, but a fraction of cells remained highly motile at distances
of >100 µm above the surface. Suspended non-motile cells showed Brownian motion
that scaled consistently with temperature and viscosity. A novel open-source automated
tracking package was used to obtain 2D tracks of motile cells and quantify motility
parameters, showing that swimming speed increased with temperature until 40C,
then plateaued. These findings are consistent with the observed heterogeneity of
B. subtilis populations, and represent the highest reported temperature for swimming
in this species. This technique is a simple, low-cost method for quantifying motility at
high temperatures and could be useful for investigation of many different cell types,
including thermophilic archaea.
Keywords: Bacillus subtilis, bacterial motility, temperature effects, heat shock, holographic microscopy,
thermophile, tracking
INTRODUCTION
The effects of elevated temperatures on bacterial motility have not been fully explored. There
are both physical and physiological effects of temperature on flagellar swimming at low Reynolds
number. Both viscosity and temperature contribute to the Stokes-Einstein equation for the diffusion
coefficient, D
D=kBT
6πηr(1)
where kBis Boltzmann’s constant, Tis the absolute temperature, ηis the dynamic viscosity of the
medium, and ris the radius of the diffusing particle. Water shows a dramatic decrease in dynamic
viscosity with temperature, described by the equation
η(T)=AeB/(TC),(2)
Frontiers in Microbiology | www.frontiersin.org 1April 2022 | Volume 13 | Article 836808
fmicb-13-836808 April 15, 2022 Time: 9:5 # 2
Dubay et al. Bacterial Motility at High Temperature
where A = 2.414 ×105Pas, B = 247.8 K, and C = 140 K. This
results in values ranging from 1.002 mPa s at 20C to 0.315 mPa
s at 90C. As a result, there is a significant change in the rate of
Brownian motion of cells at elevated temperatures: as a simple
example, 0.30 µm2/s at 33C for a 1 µm diameter cell, and 0.86
µm2/s for the same cell at 91C.
The effects of viscosity on active swimming are less clear. The
drag force is proportional to ηand thus it would be expected
that swimming speeds would increase with decreased viscosity,
but instead a surprising decrease in bacterial swimming speeds
with decreased viscosity is seen in polymer solutions. Several
papers have suggested that this is due to microstructure of the
polymer (Magariyama and Kudo, 2002;Zottl and Yeomans,
2019). In ordinary aqueous solution, a roughly linear increase
in swimming speed with temperatures up to 50C has been
reported for Escherichia coli (Maeda et al., 1976) and for multiple
other strains representing polar, bipolar, and peritrichous
flagellar arrangements (Schneider and Doetsch, 1977). One study
reported a linear increase in E. coli swimming speed with
temperature up to 40C in medium supplemented with L-serine;
in the absence of supplementation, speeds increased only up to
30C and decreased thereafter (Demir and Salman, 2012). This
linear relationship is related to increased flagellar rotation rates
at high temperatures as well as altered viscosities both inside and
outside the cell and has been modeled semi-empirically using
a large number of available motility datasets. Speed is generally
assumed to be directly proportional to flagellar rotation rate;
though this is not true in all datasets, a positive correlation is
always present (Humphries, 2013).
As a simple model, the force on a swimming cell may be
approximated as the sum of the flagellar force Fand the velocity-
dependent drag force Dv,
F=ma =FDv,(3)
where D is the drag coefficient given in Eq. (1). Although models
predict a viscosity dependence, recent studies have found that F is
independent of viscosity (Armstrong et al., 2020). This equation
of motion yields a terminal velocity of
v=F
DF
η
,(4)
where the terminal velocity should be observed during a long run
once the cell is no longer accelerating. The time-averaged velocity
over a long run should thus approximate v.
None of these models take into account the upper limits
of possible motility of different strains resulting from protein
denaturation or general organism stress. Microorganisms
show complex heat shock responses, and the expression and
maintenance of flagella can be affected by genes related to heat
stress. Motility is a complex phenotype under tight regulation
in all microorganisms that express it. Bacillus subtilis is a model
organism for which regulation of motility genes (Mukherjee and
Kearns, 2014) and heat shock responses (Schumann, 2003) have
been well studied. There is a strict dependence upon FlgN for
motility in this species (Cairns et al., 2014). Flagellar synthesis
is affected by a number of the genes involved in the heat shock
response. As temperature increases, proteins denature, and
protein degradation systems clear the damaged proteins. The
ClpCP complex in Bacillus subtilis degrades stress-damaged
proteins as well as taking part in regulatory degradation. It also
influences motility by both direct and indirect mechanisms. The
absence of Clp proteases results in defective motility, likely due to
accumulation of Spx, which suppresses flagellar gene expression.
Cells under stress do not necessarily lose motility immediately
via this mechanism, since existing flagella are not affected, rather
the production of new flagella (Moliere et al., 2016).
The heat shock responses of B. subtilis allow it to grow
to at least 53C (Warth, 1978). Much has been written about
the mechanisms of thermotolerance in this species. The sigma
factor σBprovides non-specific stress resistance, which includes
thermotolerance. Its activation induces the expression of 200
target genes (Hecker et al., 1996;Nannapaneni et al., 2012;
Young et al., 2013). Cells do not need to be exposed to heat
stress in order to gain thermotolerance via this mechanism. In
addition, there are specific heat-shock responses that turn on
under conditions of mild heat stress, leading to protection from
otherwise lethal temperatures. At least three classes of genes are
heat-shock inducible.
Because of the complexity of both the heat shock response and
regulation of motility genes, it is difficult to predict the effects
of high temperature on swimming motility. Most studies focus
on growth rather than persistence of swimming. The persistence
of the motility phenotype under conditions of heat stress has
been little explored, largely due to technical difficulties of imaging
and image analysis as temperatures rise. One study reported that
convection currents made tracking difficult above 40C (Riekeles
et al., 2021). In addition, most heated stages can only achieve
temperatures of 55–60C. We hypothesized here that some
fraction of B. subtilis cells would be capable of active motility
at temperatures above the maximum growth temperature, with
significant variations consistent with the highly heterogeneous
responses of this species to stress (Kearns and Losick, 2005;Lopez
et al., 2009;Syvertsson et al., 2021).
A few studies have reported microscopic imaging systems
able to reach temperatures up to the boiling point of water,
but none fit our precise goals. A 1995 study of Thermotoga
maritima used a capillary between two Peltier elements (Gluch
et al., 1995). A more recent study reported an environmental
chamber for microbial imaging that allows for pressure and
temperature control (Nishiyama and Arai, 2017). For studying
hyperthermophilic archaea, a “Sulfoscope” with a heated cap and
stage was recently described where temperatures of 75C were
stably maintained; temperatures up to 90C were achieved but
led to significant evaporation (Pulschen et al., 2020). This design
is appropriate for fluorescence microscopy and is especially
valuable for imaging immobilized cells, a technique which was
described in detail in the study using a semi-soft Gelrite pad.
Air objectives were used. Another recent study attained higher
resolution using oil-immersion objectives heated to 65C, with
the entire chamber heated to 75C. Imaging was by differential
interference contrast (DIC) (Charles-Orszag et al., 2021).
While high resolution is sometimes desirable, our goal
was to create a simple, inexpensive system for imaging at
Frontiers in Microbiology | www.frontiersin.org 2April 2022 | Volume 13 | Article 836808
fmicb-13-836808 April 15, 2022 Time: 9:5 # 3
Dubay et al. Bacterial Motility at High Temperature
high temperatures with a particular focus on tracking of
microorganisms in a large volume of view. Thus, the goal was
to maximize the depth of field with sufficient resolution to
distinguish individual cells, but not subcellular structure. We
use a custom holographic microscope entirely submerged in a
heated water bath in order to examine motility at temperatures
between 28 and 85C, with capability of temperatures up to at
least 95C. The materials for the heated bath are inexpensive
(<$200 USD in 2022) and a complete parts list is included for
both the microscope and bath, with a total cost of <$7000 USD
for the complete system. Because the system was designed for
field use and no compound objective lenses are used, there are
minimal effects of temperature on the instrument, even up to
the boiling point.
With this system, cultures of B. subtilis were either heated
slowly over a period of 4 h or exposed rapidly by submersion into
a pre-heated bath. Individual cells were tracked, and velocities,
accelerations, turn frequency, and correlation functions were all
quantified using a custom open-source autonomous program.
The program successfully controlled for thermal currents,
classifying tracks into “motile” vs. “non-motile” based upon a
defined algorithm. All active tracks were manually confirmed
to correspond to cells. Active swimming was observed up to
66C in some fraction of the cells; increasing numbers of
immobile cells collected on the sample chamber surface as
temperatures increased. Cells up to 90C remained normal in
morphology, without signs of sporulation. Dead cells showed
little movement until temperatures >60C, at which point
convection currents became appreciable; however, these were
readily distinguished from active motility by inspection as well
as automated tracking tools. These techniques will be of interest
to anyone exploring bacterial behavior at temperatures up to the
boiling point of water.
MATERIALS AND METHODS
Microscope and Temperature Control
The microscope used in this study was a common-path
off-axis digital holographic microscope (DHM) as described
previously (Wallace et al., 2015) (U.S. Patent US20160131882A1)
(Figure 1A). Briefly, a single-mode laser (520 nm, Thorlabs,
Newton, NJ) was collimated and passed through separate
reference and sample channels held in a single plane to prevent
mis-alignment. The objective lenses were simple achromats
(numerical aperture 0.31, part number 47–689-INK, Edmund
Optics, Barrington, NJ), yielding an effective magnification of
20x and XY spatial resolution of 1.0 µm. Since focusing is
performed numerically with DHM, the sample stage was fixed in
position with the in-focus plane Z= 0 at approximately the center
of the sample chamber. The camera was a Prosilica 2460GT
(Allied Vision, purchased from Edmund Optics) monochrome
camera with a 5 MPixel, 3.45 µm pixel pitch format and 15
frames/s maximum frame rate, with acquisition windowed to
2,048 ×2,048 px. A higher frame rate (23.7 fps) can be achieved
by substituting the newer Prosilica 2560GT without further
modification of the system. Both recommended cameras use
global shutter detector readout; rolling shutter readout cameras
are not recommended unless careful consideration is taken in
avoiding distortion of fringes during readout.
To control the temperature of the sample chamber, the
entire microscope was placed into a stainless steel cooking
pot (36 quart, Bayou Classic, sold through amazon.com) and
protected by a 5-gallon, 4 mil thickness plastic bag (Ziploc
brand or autoclave bag). The laser was kept out of the bath
and coupled to the microscope through its single-mode fiber
output. The camera was kept above the water level of the bath
by the microscope tube. The pot was filled with water to a
level above the top of the sample chamber, and the temperature
of the water was gradually increased at a constant rate of
4.3C/min using a sous vide circulator (Monoprice Model#:
121594, 800 W, 4 gallon capacity) (Figures 1B,C). The water
temperature indicated on the circulator was confirmed using
a laboratory thermometer (Fisherbrand). In order to raise the
temperature above 80C, it was necessary to insulate the pot
with metallized bubble wrap (Figure 1C). A thermocouple (Gain
Express K-Type, sold through amazon.com) was taped both
above and below the chamber, in direct contact with the chamber,
in an independent set of experiments to determine how the
sample temperature corresponded to the water temperature; a
nearly constant offset of 5C was seen between the chamber
and the water temperature, and this correction was applied to
all reported chamber temperatures (Figure 1D). Supplementary
Figures 1,2show the steps involved in setup, and Supplementary
Datasheet 1 provides a parts list for the entire instrument with
purchasing links. The costliest elements are the laser and camera.
Preliminary thermal testing of elements was performed by
submerging individual parts of the setup into water in Ziploc
bags and heating the water using the sous vide controller to its
maximum temperature (just below boiling). In the case of the
objective lens holder, because of its small size and low cost, it
was placed into boiling water on the stovetop both with and
without the lenses installed. The upper temperature limit of the
microscope is set by the plastic used in the 3D-printed lens holder
and the bonding of the sample chamber; substitutions of these
components with higher temperature materials would enable
higher temperature investigations. None of the components was
tested above 100C in this study.
Samples and Chambers
Bacillus subtilis type strain (ATCC 6051) was obtained from the
American Type Culture Collection, Manassas, VA. Stocks were
maintained at 80C and periodically streaked onto lysogeny
broth (LB, Thermo Fisher Scientific, Pittsburgh, PA)-agar plates.
16–24 h before each experiment, single colonies were picked
from plates, seeded into liquid LB, and incubated at 30C
with shaking to mid-log phase (OD 0.6 as measured on a
Clariostar plate reader). After incubation, samples were diluted
1:1,000–1:100 into motility medium (10 mM phosphate buffer
pH 7.4, 10 mM NaCl, 0.1 mM EDTA) and moved to custom
sample chambers at 20 ±2C. To control for any possible
non-biological motion at high temperatures, an additional set
of experiments was performed using heat-killed B. subtilis cells
at the same concentration. The cells were exposed to boiling
Frontiers in Microbiology | www.frontiersin.org 3April 2022 | Volume 13 | Article 836808
fmicb-13-836808 April 15, 2022 Time: 9:5 # 4
Dubay et al. Bacterial Motility at High Temperature
FIGURE 1 | Temperature-controlled microscope setup. (A) Schematic of off-axis DHM used in these experiments. (B) The microscope was protected by a
heavy-duty plastic bag and submerged in heated water to control the sample temperature. (C) Photograph of the complete setup showing the insulating materials
necessary to attain the highest temperatures. The microscope is submerged into the pot with the sample chamber located several cm below the water surface. The
insulation shown is required for achieving water and sample temperatures above 70C. (D) Correspondence between the sous vide measured water temperature
(verified with a laboratory thermometer) and thermocouple measurements of the sample chamber at top and bottom, along with the average between top and
bottom. The slope of the temperature change was consistent between the bath and the chamber, with a nearly constant offset of 5C. The lines are linear fits with
slopes indicated.
water for 30 min, then pelleted and washed twice with motility
medium before imaging.
The sample chambers (product of Aline, Rancho Dominguez,
CA) are composed of two channels, one for the sample and one
as a reference channel which is filled with 0.9% saline solution
or sterile growth medium (Figure 2). These are composed of
two layers of optical quality glass with a middle acrylic layer
forming the channels. Homemade chambers using microscope
slides, coverslips, and adhesive material such as silicone or Teflon
may also be used; for detailed instructions (see Supplementary
Figures 3–5). The depth of the chamber is 1.0 mm, which allows
for motility far away from the surfaces in order to eliminate
Frontiers in Microbiology | www.frontiersin.org 4April 2022 | Volume 13 | Article 836808
fmicb-13-836808 April 15, 2022 Time: 9:5 # 5
Dubay et al. Bacterial Motility at High Temperature
FIGURE 2 | Sample chamber arrangement. (A) Schematic of the 1 mm deep chamber, showing the sample (S) and reference (R) channels that permit the off-axis
DHM geometry. The volume of view of each snapshot is 365 µm×365 µm×1 mm in x, y, and z. (B) The chambers were loaded using a 1 mL syringe and placed
onto the microscope stage at a fixed focus before immersion.
surface influences on motility (Li et al., 2008;Giacche et al., 2010;
Li et al., 2011). The sample was placed into the microscope
before immersion and left throughout the duration of the heating.
Replicate experiments (2–4) were performed on different days
with independent B. subtilis cultures in order to confirm the
reproducibility of the results. Cell density in the chambers varied
from 105to 106cells/mL.
Acquisition and Reconstruction of
Holograms
Data were acquired using a custom, open-source software
package, DHMx.1Recordings were between 30 and 60 s long,
with a maximum frame rate of 15 frames per second (fps).
Recordings were obtained every 5C until reaching a maximum
sample temperature of 84C. Additional tests were performed
which subjected the bacteria to rapid increases in temperature.
Samples were loaded into chambers at 20C and submerged
into pre-heated water at varying temperatures. For these “heat
shock” experiments, recordings were obtained immediately after
submersion as well as 2 and 10 min afterward. The 10 min
datasets were chosen for analysis.
The holograms for each recording were either median
subtracted (as we described previously Bedrossian et al., 2020)
or frame-to-frame subtracted then reconstructed in amplitude
using Fiji (ImageJ) (RRID:SCR_002285) (Schindelin et al., 2012).
Reconstructions were performed using the angular spectrum
method (Mann et al., 2005) implemented in a custom plug-in
described in detail elsewhere (Cohoe et al., 2019) and available
from our update site.2The choice of frame-to-frame subtraction
was necessary at temperatures >50C in order to eliminate
noise due to non-stationary cells on the chamber surface. The
z thickness chosen for reconstruction ranged from 400 to 800
µm and varied somewhat among datasets depending upon the
location of active cells. The resulting stacks were maximum
projected in Z using Fiji to create a 2D time series of all of the
cells of interest.
1https://github.com/dhm-org/dhm_suite
2https://github.com/sudgy/
High Resolution Light and Fluorescence
Microscopy
In order to carefully evaluate cell morphology and possible
presence of spores, cells were examined on an Olympus IX-
71 inverted microscope with a 100x, NA = 1.4 oil immersion
objective using either brightfield illumination or fluorescence
illumination with a Hg lamp and a 450/50 nm excitation filter
and 520 nm longpass (Chroma Technology, Bellows Falls, VT).
Fraction Motile
Non-motile cells were difficult to count from images, especially
at higher temperatures as cells began to cluster, so the fraction
of motile cells was estimated as the number of motile cells per
volume of view averaged over the length of the recording, divided
by the total cell concentration (as measured by original OD
divided by dilution factor; the relationship between OD and
cell concentration was established using a hemocytometer). The
instantaneous volume of view is 0.365 ×0.365 ×1 mm3, or
0.13 µL, corresponding to 100 cells/frame at 106cells/mL. The
number of motile cells per frame was calculated using frame-
to-frame subtracted projections so that non-motile cells did not
interfere with the analysis. Fiji Analyze Particles3was used to
count the number of cells per frame.
Tracking and Statistics
Bacteria were tracked using a custom software package,
Holographic Examination for Life-like Motility (HELM),4which
was developed to autonomously detect, track, and characterize
motile cells. HELM identifies pixel changes in sequential DHM
images, tracks clusters of change as particle movement, and
classifies particles as motile or non-motile based on their
movement patterns. The pixel changes are computed by simple
background subtraction of the median video image. Clusters
of pixel changes are identified using DBSCAN. Tracks are
then generated using the Linear Assignment Problem (LAP)
tracker method (Jaqaman et al., 2008). With the spatiotemporal
3https://imagej.net/imaging/particle-analysis
4https://github.com/JPLMLIA/OWLS-Autonomy
Frontiers in Microbiology | www.frontiersin.org 5April 2022 | Volume 13 | Article 836808
fmicb-13-836808 April 15, 2022 Time: 9:5 # 6
Dubay et al. Bacterial Motility at High Temperature
points, HELM computes approximately two dozen metrics that
form a feature vector for each track. These include features
like mean speed, mean turn angle, total track displacement,
etc. A random forest classifier (Breiman, 2001) is then used
to classify each track as motile or non-motile. The classifier
was trained using manually labeled tracks from both prepared
laboratory and field-acquired ocean water samples, both with
and without fluid flow in the sample chamber. HELM can be
used on raw, unreconstructed holograms or on 2D projections of
reconstructed holograms. Analysis presented here was performed
on projections of reconstructed holograms as described in section
“Motility Analysis.”
HELM also generates multiple contextual products to support
assessment of a DHM recording. One of these, the Motion
History Image (MHI), summarizes a full video in one image by
color mapping each pixel to the time index of largest intensity
change. The MHI image allows a rapid understanding of how
many particles were present in the recording as well as the
presence and characteristics of potential motile organisms.
Diffusion coefficients were measured using NanoTrackJ
(Wagner et al., 2014). A video of at least 30 s containing at least 10
trackable particles was analyzed using the Maxima and Gaussian
Fit center estimator and the covariance diffusion coefficient
estimator. Parameters used were: minimum estimated particle
size, 20 pixels; minimum number of steps per track, 5; pixel size,
178 nm; and frame rate 15 frames/s.
Statistical analysis of HELM outputs and graphing were
performed using Prism 9 (GraphPad Software, San Diego,
CA). Statistical relevance was estimated using the ANOVA
package in Prism after ensuring that distributions were Gaussian.
The correlation matrix was computed using the Multivariable
Analysis package in Prism.
RESULTS
Microscope Function and Maintenance
at High Temperatures
We tested all elements of the system for their ability to withstand
temperatures up to 100C. The DHM optics are robust, as
they contain no compound objectives, adhesives, or other heat-
sensitive elements. The 3D printed lens holder and objectives
withstood boiling up to 30 min. The most sensitive element
of the setup was the fiber optic cable. Under preliminary tests,
the cable housing degraded, allowing light to escape, when the
cable was heated to 90C. This could be prevented by shielding
the cable from the heat using an aluminum mini-box as shown
in Supplementary Figure 1. On occasion, condensation would
form on the sample chamber, objective lens, or collimating lens.
This could be removed by wiping with lens paper. After multiple
rounds of experiments or if image quality appeared poor, all of
the lenses were removed and cleaned with 100% isopropanol. It
was also important to keep the camera outside of the hot area to
prevent condensation on the window.
Morphology and Brownian Motion
There was no significant difference in size or overall morphology
of motile cells as the temperature increased (Figures 3,4A).
Non-motile cells showed increased rates of Brownian motion
consistent with scaling of diffusion rates with temperature (kBT)
FIGURE 3 | General appearance and size of B. subtilis at normal and elevated temperatures. Shown are phase contrast (top row) and autofluorescence (bottom
row); scale bar = 5 µm. (A) 30C. (B) 60C. (C) 90C.
Frontiers in Microbiology | www.frontiersin.org 6April 2022 | Volume 13 | Article 836808
fmicb-13-836808 April 15, 2022 Time: 9:5 # 7
Dubay et al. Bacterial Motility at High Temperature
FIGURE 4 | Cell sizes and diffusion coefficients with temperature. (A) Cell area as estimated by microscopy, shown as mean ±standard deviation. The differences in
the means from one temperature to the next was not significant. (B) Measured diffusion coefficients at 3 temperatures (black dots), with measured standard
deviations, compared with predicted values according to Eq. (1) for particle radii of 2.0, 2.5, and 3.0 µm.
FIGURE 5 | Data reconstruction and processing for tracking. (A) A portion of the field of view of a median-subtracted hologram from the 34C dataset. The inset
shows part of the image magnified 4x to show the fringes. (B) Amplitude reconstruction of the same field of view in (A) at +50 µm. (C) Maximum Z projection of the
same field of view, representing reconstructions from 200 to +200 µm in steps of 10 µm.
and dynamic viscosity of water (η). The Stokes-Einstein relation
(Eq. 1) gave an excellent fit to the measured data for particle
radius r= 2.1 µm (Figure 4B). This was consistent with the sizes
measured by direct imaging, with a large amount of variability
seen in both measured size and diffusion coefficient due to
the presence of elongated and clustered cells. For temperatures
higher than 44C, Brownian motion was impossible to evaluate
because of either thermal currents or large numbers of cells
immobilized on the glass surface.
Motility Analysis
Motility analysis was performed on reconstructions and
projections of holograms. An example hologram is shown in
Figure 5A, and a single-plane amplitude reconstruction in
Figure 5B. Single plane reconstructions provided qualitative
insight into cell morphologies, speeds, and fraction of motile
cells; however, their signal to noise ratio was insufficient for
automated tracking. Maximum projections through 40–80 Z
planes, representing 400–800 µm sample depth (Figure 5C),
permitted automated tracking to create 2D trajectories. Motion
history images (MHIs, as described in “Materials and Methods”)
were used to identify tracks of motile organisms from these
projections (see Supplementary Video 1). Figure 6 shows
time-coded MHIs for temperatures from 28 to 84C for a
selected set of experiments. The spacing between points on each
track gives a rough estimate of cell speed. Where thermal drift
was significant, motile cells could be identified by their paths
of travel against the drift current. Only cells moving counter
to the thermal drift were selected for tracking, as they clearly
represented active motility.
Frontiers in Microbiology | www.frontiersin.org 7April 2022 | Volume 13 | Article 836808
fmicb-13-836808 April 15, 2022 Time: 9:5 # 8
Dubay et al. Bacterial Motility at High Temperature
FIGURE 6 | MHI analysis shows changes in motility patterns with increasing temperature and assists in identifying motile tracks. The tracks are time-coded, with the
time index indicating frame number (15 frames/s). The scale bar applies to all panels. (A–F) Full field of view of all identified tracks in selected datasets. (A) 28C,
showing a distribution of motile and non-motile cells and distinct swimming patterns. (B) 34C, showing increased motility and speed. (C) 44C, showing nearly all
cells motile at high speed. (D) 61C, showing a reduction in the number of motile cells, but some swimming at high speed. (E) 66C, showing a large amount of
thermal drift with a few motile cells. (F) 84C, with all motion due to thermal drift. (G–J) Selected examples of swimming types identified in the tracks. Magenta
indicates tracks identified as motile by the software; tracks in cyan were identified as non-motile. (G) Helical swimming. (H) Long straight runs. (I) Circling or
spinning. (J) Run and tumble.
Frontiers in Microbiology | www.frontiersin.org 8April 2022 | Volume 13 | Article 836808
fmicb-13-836808 April 15, 2022 Time: 9:5 # 9
Dubay et al. Bacterial Motility at High Temperature
TABLE 1 | Motility parameters.
Temperature (C) Fraction of
cells
motile ±SD
Total # tracks
analyzed
(# replicates)
# fast runs/
tumbles/helices/
circles
Mean ±SD speed
(µm/s) (range)
Mean ±SD
acceleration
(µm/s2)
Mean ±SD
displacement
(µm/s) (range)
28 20 ±1 36 (2) 10/8/10/8 16 ±7 (3–32) 150 ±70 5 ±4 (0.2–16)
34 24 ±2 66 (2) 24/26/8/8 23 ±6 (12–42)* 200 ±60* 9 ±5 (0.7–21)
44 61 ±2* 57 (2) 18/30/5/4 24 ±7 (8–45)* 210 ±90* 10 ±6 (0.1–25)*
61 6 ±2* 16 (2) 10/5/1/0 24 ±7µm/s (14–38)* 230 ±90* 10 ±7 (1.3–27)
66 <1* 35 (2) 18/13/2/0 23 ±6 (12–41)* 220 ±70* 8 ±5 (1.3–26)
51 heat shock Not done 21 (1) 2/2/10/7 28 ±6 (15–47)* 250 ±110* 10 ±7 (2–24)
60 heat shock Not done 24 (1) 2/12/8/2 22 ±6 (13–37)* 160 ±70 8 ±5 (1.1–18)
“Replicates” represent independent experiments done on different days with fresh cultures of B. subtilis. Tracks were characterized according to Figure 6 by visual
inspection and correlation with parameters as in Figure 7. SD, standard deviation. (*), significantly different from value at 28C(p<0.01).
The fraction of motile cells (see Table 1 for statistics) increased
from 20% at 28C (Figure 6A and Supplementary Video 2) to
25% at 34C (Figure 6B and Supplementary Video 3) to >60%
44C (Figure 6C and Supplementary Video 4). However, at 61
and 66C, the fraction of motile cells was greatly reduced, to <6%
and <1%, respectively. Motile tracks could still be identified
in single-plane reconstructions (Supplementary Video 5) and
in MHI analysis projections through Z (Figures 6D,E and
Supplementary Video 6). At 84C, only thermal currents and
drift were observed, with no active counter-current motility;
the fraction of motile cells was deemed 0% (Figure 6F and
Supplementary Video 7). Qualitative analysis of these tracks on
a cell-by-cell basis revealed that there were 4 basic track types:
(1) helical tracks (Figure 6G); (2) long straight runs (Figure 6H);
(3) circular tracks (spinning) (Figure 6I); and (4) runs with
tumbles and directional changes, which could feature either
straight or helical swimming or some combination of both; any
track with distinct directional changes was classified as run and
tumble (Figure 6J).
Heat-killed cells showed no active motility, and the
MHI images correspondingly showed few features at the
lower temperatures (Supplementary Figure 6). Amplitude
reconstructions of the holograms revealed cells undergoing
slight drift and Brownian motion (Supplementary Video 8
shows 28C). At temperatures of 60C and higher, substantial
thermal drift and convection became apparent. At the highest
temperatures, multiple drift planes could be observed, but
in all cases motion was clearly due to bulk flow without any
evidence of run and tumble events (Supplementary Figure 6
and Supplementary Video 9 shows 84C). HELM’s classification
algorithm was 100% successful at classifying tracks as non-
motile at temperatures <50C, and >95% successful at higher
temperatures (Supplementary Video 10 shows classification of
tracks at 84C). This classification is not based upon any one
parameter, and correlating its results with physical parameters
will be the subject of a future study. Visual inspection is always
required to correlate tracks with cells and to perceive patterns
suggesting active motility.
Quantitative analysis of tracks was performed by using the
automated tracker combined with the MHI to identify tracks
that represented valid cell trajectories. Tracks that were not
following organisms or which were fewer than 15 frames long
were excluded from analysis. Table 1 gives classification of the
analyzed motile tracks for selected datasets. Parameters extracted
from these tracks are shown in Table 1 and Figure 7. Data
from independent experiments were consistent, so full tracking
was performed using selected tracks from pools of replicates
(see Supplementary Figure 7 for comparisons of replicate
experiments). Plotted in Figure 7 are selected parameters
where significant differences were seen among the different
temperatures or where parameters assisted in classifying tracks.
The full datasets are available in Supplementary Datasheets 2–7.
The mean speed was defined for each trajectory as
v=1
N
N
X
i=i
||
xi
xi1||
1t,(5)
where Nis all of the points in the identified trajectory and 1t
is the (constant) frame rate. There was a significant increase
in mean speed between 28C and all of the other elevated
temperatures. There was a statistically significant difference
(p<0.001) for comparison between 28C and the other
temperatures, and not significant between any other pairs. The
fastest speeds were seen in the 51C heat shock sample (p<0.001
for comparison with 28, 35, and 60C heat shock). The values for
the 60C heat shock were comparable to those at all other elevated
temperatures (significantly different from 28 to 51C heat shock,
all others non-significant). Values are given in Table 1 and means
with standard errors of the mean are plotted in Figure 7A for
both live and killed cells; distributions were Gaussian at each
temperature. The speeds of the killed cells were significantly less
than those of the live cells, even when drift was significant. There
was no significant difference in maximum trajectory speed seen
between any pairs of data sets of live cells (not shown; means
80–100 µm/s for maximum instantaneous speeds).
The mean speed times viscosity, which should give an
approximate measure of flagellar force as given in Eq. (4),
decreased essentially linearly at higher temperatures. This is
consistent with denaturation of the proteins as the temperatures
rise, representing a typical optimum performance curve with the
optimal temperature near 310 K (37C) (Figure 7B). Because
heat denaturation is a time-dependent process involving multiple
parameters (Peterson et al., 2007), additional time points at each
Frontiers in Microbiology | www.frontiersin.org 9April 2022 | Volume 13 | Article 836808
fmicb-13-836808 April 15, 2022 Time: 9:5 # 10
Dubay et al. Bacterial Motility at High Temperature
FIGURE 7 | Selected motility parameters. Definitions of the parameters, values and statistical significance are given in the text. Error bars shown are
mean ±standard error of the mean unless noted. Numbers of analyzed tracks and their classifications are given in Table 1. When error bars do not appear, they are
smaller than the symbols. (A) Mean speed. (B) Mean speed times viscosity of water at that temperature. The lines are linear fits to temperatures above or below 310
K. (C) Total displacement normalized to track length and average speed <v>.(D) Histogram of selected values of D/ <v>T, showing classification into long runs vs.
run-and-tumble traces. (E) Sinuosity (no error bars given as the distributions were not Gaussian and the mean value had little significance). These values are plotted
on a Log_2 scale for ease of visualization of the ranges involved. Circular tracks were identified as those with sinuosity >20. (F) Correlation matrix of the measured
parameters.
temperature would be needed to extract meaningful physical
parameters from this temperature dependence.
As distinguished from mean speed, displacement (µm/s)
looks at the whole trajectory rather than each frame, and is
the norm of the total XY path, D = q(4x)2+(4y)2. For
a straight track, displacement will equal mean speed <v>
multiplied by total time T of the track lifetime; for a circular track,
displacement will be close to 0 regardless of mean speed. Dividing
the displacement by <v>Tgives a dimensionless number that
can be used to classify tracks. Figure 7C shows the average
D/<v>Tvs. temperature, and (Figure 7D) shows a histogram
of displacements for selected measured tracks at multiple
temperatures; values of D/<v>T0.55 indicated long runs.
Sinuosity is a measure of movement inefficiency defined as
end-to-end displacement over total path length in µm. For a
straight path, sinuosity will equal 1. For a serpentine or circular
movement pattern, sinuosity will be 1. Values of >20 were
only seen at lower temperatures (Figure 7E). The cell-by-cell
classifications in Table 1 agreed with this analysis, as circular
tracks were not found in the elevated temperatures.
Acceleration (µm/s2) is measured as inter-frame differences in
velocity:
a=1
N
N
X
i=i
||
vi
vi1||
4t,(6)
and could be used along with mean speed to classify tracks. Long,
smooth runs showed low values of acceleration, while tracks with
multiple tumble events showed high values. Similarly, the step
angle measures how much a particle turns at each time point.
It gives the angle that a particle deviated from a straight path
per inter-frame interval (in radians). Passively drifting particles
should not very much in their direction from frame to frame,
while highly motile particles that swerve and turn regularly will
show large changes in step angle. Figure 7F shows a correlation
Frontiers in Microbiology | www.frontiersin.org 10 April 2022 | Volume 13 | Article 836808
fmicb-13-836808 April 15, 2022 Time: 9:5 # 11
Dubay et al. Bacterial Motility at High Temperature
FIGURE 8 | Collections of non-motile cells on the chamber surface at 66C. (A) Still image; the circle indicates a cluster of cells. (B) Schematic of surface clustered
cells with relative positions of selected motile tracks. (C) Corresponding image of the slide surface with the killed cell dataset at 66C.
matrix of the measured parameters, showing that D/<v>T
showed a negative correlation with nearly all measured values.
Autocorrelations of speed, velocity, and turn angle at 1 and 2
s did not vary significantly among datasets (not shown; available
in Supplementary Datasheets 2–7).
Surface Clustering
At higher temperatures (60C and above), cells were seen to be
swimming above large patches of surface-adherent bacteria. The
Z projections could not determine how far the highly motile cells
were from the patches, or whether they were interacting. Thus, it
was necessary to use full Z plane reconstructions to identify the
relative position of motile cells vs. collections of non-motile cells
at the chamber bottom. It was seen that the highly motile cells
were moving at focal planes tens to hundreds of microns away
from the chamber bottom (Figures 8A,B and Supplementary
Video 11). While some active motility was seen near the clustered
cells at the surface, it was significantly slower as expected near a
surface (Li et al., 2008, 2011). Similar groups of cells were not seen
with the killed sample (Figure 8C).
DISCUSSION
The novel immersion system reported in this paper allows
for recordings of bacterial motility up to nearly the boiling
point of water. The custom software package that we report
for the first time here aids in tracking of motile organisms as
well as identification of non-motile tracks resulting from drift.
Distinguishing drift from motility was straightforward by visual
inspection, as active motility could occur perpendicular to the
thermal drift. However, the automated algorithm did generate
a large number of false positive tracks, so manual validation
was essential to accurately identify motile organisms. The MHI
images were used to help identify tracks that corresponded to
organisms. Although human intervention is needed, this method
is significantly easier than manual tracking. As HELM was
designed for use on spacecraft, HELM can process a sample in
several minutes on an ordinary laptop computer, and manual
confirmation of motile tracks may be performed in less than an
hour. It is important to note that the current implementation of
the software assumes (and checks for) a constant frame rate, so
it is necessary to ensure that the acquisition camera and software
do not drop frames. Future versions of the software will enable
input of a timestamps file to accommodate varying frame rates.
An approach to extracting tracks directly from the MHI traces is
also in development.
For this work, tracking was only performed on 2D projections
of the 3D tracks. The errors introduced in velocities and turn
angles by this approximation have been reported (Taute et al.,
2015). Given that the axial resolution of the instrument is >2µm,
we decided that in this analysis, the additional computational and
time cost of 3D tracking does not contribute significantly to the
analysis of velocities, as we have shown in a previous analysis
(Acres and Nadeau, 2021). In 2D, helices appear as spirals, with all
of the parameters of the helix readily extractable from the 3D data
(Gurarie et al., 2011). For organisms or instruments that show
higher contrast, HELM could also be used on individual Z planes.
However, with our current implementation, the signal to noise of
B. subtilis is too low for tracking on a single plane, so projections
over multiple planes were used. Since most literature data is
based upon 2D tracks, this also allows for easier comparison
of our measured speeds with those found in other studies. The
speeds we observed, 15–25 µm/s, are consistent with the recent
reports (Turner et al., 2016). However, large variations from
study to study have made general conclusions about parameters
as simple as mean velocity for a given strain largely impossible.
A new database, named BOSO-Micro (Rodrigues et al., 2021), is
aiming to increase standardization of motility experiments and
parameters. Such databases will be important for summarizing
the large amounts of data being generated in the rapidly emerging
field of bacterial tracking.
Using this system and approach, we found that B. subtilis
was capable of active motility up to 66C under conditions of
constant heating over a time course of 4 h. Cell size and shape
did not significantly change with temperature. The changes in
speed with temperature were consistent with previous studies up
to 50C (Schneider and Doetsch, 1977), with data unavailable
past that point. We observed active motility to 66C, but with
a minority of cells being motile at temperatures above 45C.
The majority of cells were found clustered on the lower surface
Frontiers in Microbiology | www.frontiersin.org 11 April 2022 | Volume 13 | Article 836808
fmicb-13-836808 April 15, 2022 Time: 9:5 # 12
Dubay et al. Bacterial Motility at High Temperature
of the chamber at high temperatures. This is consistent with
cell population heterogeneity in this species (Kearns and Losick,
2005;Lopez et al., 2009;Syvertsson et al., 2021). In addition, some
of the cells in this recording showed a change in the observed
motility type, most commonly a change from run-and-tumble
to long runs. These parameters were indicated by turn angles,
total displacement, and sinuosity in our analysis. Tracks with
high sinuosity, indicating circling, were seen only at the lower
temperatures. Long runs could be readily identified from total
displacement divided by average velocity and time length of
the track. The observed effect may result from the highly non-
linear viscosity of water at high temperatures, but additional
experiments are necessary to attempt to deconvolve viscosity
and temperature.
After 66C, movement in the field was due to thermal drift
alone. The thermocouples used to confirm the temperature of
the chamber showed a slight difference between the top and
bottom glass of the chamber, which could have led to increased
thermal currents, most notably at temperatures past the limit
for autonomous motion. This drift was clearly distinguishable
from active motility on the MHI traces, but attention to correct
classification of tracks as motile or non-motile was essential
during analysis. The algorithm performed well with the killed
samples, identifying nearly all of the tracks as non-motile despite
drift, although it did yield a small number of false positives at
the highest temperature (84C). The random forest classifier used
to classify tracks as motile or non-motile relies upon numerous
parameters, no single one of which is sufficient to classify a
track as motile or non-motile. Further analysis will correlate
physical parameters with motility; this may require analysis
beyond the use of means and standard deviations across an
entire track, but close attention to instantaneous parameters.
This analysis will also aid in classifying motile tracks with
regard to run length and tumble frequency. Custom training of
the classification software may be necessary for datasets with
substantially different flow profiles.
The MHI traces are a unique feature of HELM, which was
designed with spaceflight missions in mind, where the detection
of possible signs of life with the least possible processing power is
required. The MHI allows visualization of tracks in low signal-
to-noise recordings where objects cannot easily be identified
by tracking algorithms. These traces may be used as a guide
for selection of complete motile tracks identified by HELM.
However, compared to other software packages such TrackMate,
HELM’s particle identification and tracking is less interactive.
When the algorithms do not perform well, the MHI may be
used as a guide to finding tracks manually or using other
particle detection algorithms in other packages. When HELM’s
detection and tracking works well, tracks are exported in.json
files which may be imported into other packages for stitching or
further analysis.
Reconstruction at individual Z planes at higher temperatures
showed rapidly swimming cells tens to hundreds of µm above
the surface, with large numbers of cells on the surface, some
of which exhibited active motility. The presence of active
motility in mesophilic organisms at temperatures beyond those
at which they can grow is somewhat of a surprising result.
Direct visualization of the motion of individual cells in this
work makes this result unambiguous. The use of killed control
cells eliminated any possibility that the motion was due to
complex thermal currents, and the disappearance of any signs
of active motility above 66C further indicates that the “live”
cells became inactive at this point. When reduced viscosity at
these temperatures was considered, there was a temperature-
dependent reduction in flagellar force. This is likely due to protein
denaturation, and longer incubation times at these elevated
temperatures may eventually lead to complete loss of motility.
More detailed experiments with controlled incubation times at
specific temperatures may provide useful models of thermal
stability of motility-related proteins (Daniel and Danson, 2010).
This paper represents the first step toward evaluating the upper
limits of temperature on mesophile motility.
The authors hope that this simple setup will encourage
others to reproduce these experiments and examine other strains
of bacteria and archaea. In contrast to bacteria, especially
test strains such as E. coli and B. subtilis, hyperthemophilic
archaea have not been frequently imaged. The setup we
report here should facilitate studies of thermophilic organisms,
including those such as Pyrococcus furiosus which require
temperatures near the boiling point of water for optimal motility
(Herzog and Wirth, 2012).
The detailed parts list, combined with the open-source
software, should be sufficient to enable duplication of the system
by anyone who wishes to perform these experiments. The only
custom parts are a 3D printed stage and objective lens holder
(plastic) and a custom machined stage (aluminum) to allow
for easy changing of sample chambers without requiring stage
realignment. CAD drawings of these can be provided on request,
and users are encouraged to tailor designs to their own specific
applications. It is important to ensure that these elements are
made from materials that can withstand high temperatures;
any materials chosen should ideally be tested beforehand by
submerging them into hot water at the desired temperatures
before use on the microscope.
DATA AVAILABILITY STATEMENT
The raw holograms for the datasets used here are deposited
in a public depository at Data Dryad, accession https://doi.
org/10.5061/dryad.ns1rn8pv6. Other data are available from the
authors upon request. The software packages used are all open
source and are available at the following sites: DHMx: https:
//github.com/dhm-org/dhm_suite; HELM: https://github.com/
JPLMLIA/OWLS-Autonomy; and Reconstruction Fiji plug-ins:
https://github.com/sudgy/.
AUTHOR CONTRIBUTIONS
MD and NJ: construction and calibration of thermal control
apparatus, growth and maintenance of bacteria, data collection,
data analysis, data archiving, and writing. MW and JL:
development of HELM tracking software, troubleshooting and
debugging of software, and addition of new features upon
request. CL: original concept design and acquisition of funding.
Frontiers in Microbiology | www.frontiersin.org 12 April 2022 | Volume 13 | Article 836808
fmicb-13-836808 April 15, 2022 Time: 9:5 # 13
Dubay et al. Bacterial Motility at High Temperature
JN: acquisition of funding, supervision and planning of
experiments, data analysis, and writing. All authors: editing and
approval of final draft.
FUNDING
The authors acknowledge the support of the National
Science Foundation (Grant No. 1828793). Portions of this
work were supported under a contract from, or performed
at, the Jet Propulsion Laboratory, California Institute of
Technology, under a contract with the National Aeronautics and
Space Administration.
SUPPLEMENTARY MATERIAL
The Supplementary Material for this article can be found
online at: https://www.frontiersin.org/articles/10.3389/fmicb.
2022.836808/full#supplementary-material
REFERENCES
Acres, J., and Nadeau, J. (2021). 2D vs 3D tracking in bacterial motility analysis.
AIMS Biophys. 8, 385–399.
Armstrong, D. J., Nieminen, T. A., Favre-Bulle, I., Stilgoe, A. B., Lenton, I. C. D.,
Schembri, M. A., et al. (2020). Optical force measurements illuminate dynamics
of Escherichia coli in viscous media. Front. Phys. 8:575732. doi: 10.3389/fphy.
2020.575732
Bedrossian, M., Wallace, J. K., Serabyn, E., Lindensmith, C., and Nadeau, J. (2020).
Enhancing final image contrast in off-axis digital holography using residual
fringes. Opt. Express 28, 16764–16771. doi: 10.1364/OE.394231
Breiman, L. (2001). Random forests. Mach. Learn. 45, 5–32. doi: 10.1023/A:
1010933404324
Cairns, L. S., Marlow, V. L., Kiley, T. B., Birchall, C., Ostrowski, A., Aldridge, P. D.,
et al. (2014). FlgN is required for flagellum-based motility by Bacillus subtilis.
J. Bacteriol. 196, 2216–2226. doi: 10.1128/jb.01599-14
Charles-Orszag, A., Lord, S. J., and Mullins, R. D. (2021). High-temperature
live-cell imaging of cytokinesis, cell motility, and cell-cell interactions in the
thermoacidophilic crenarchaeon Sulfolobus acidocaldarius.Front. Microbiol.
12:707124. doi: 10.3389/fmicb.2021.707124
Cohoe, D., Hanczarek, I., Wallace, J. K., and Nadeau, J. (2019). Multiwavelength
imaging and unwrapping of protozoa in amplitude and phase using custom Fiji
plug-ins. Front. Phys. 7:94.
Daniel, R. M., and Danson, M. J. (2010). A new understanding of how temperature
affects the catalytic activity of enzymes. Trends Biochem. Sci. 35, 584–591.
Demir, M., and Salman, H. (2012). Bacterial thermotaxis by speed modulation.
Biophys. J. 103, 1683–1690. doi: 10.1016/j.bpj.2012.09.005
Giacche, D., Ishikawa, T., and Yamaguchi, T. (2010). Hydrodynamic entrapment
of bacteria swimming near a solid surface. Phys. Rev. E Stat. Nonlin. Soft Matter
Phys. 82:056309. doi: 10.1103/PhysRevE.82.056309
Gluch, M. F., Typke, D., and Baumeister, W. (1995). Motility and thermotactic
responses of Thermotoga maritima.J. Bacteriol. 177, 5473–5479. doi: 10.1128/
jb.177.19.5473-5479.1995
Gurarie, E., Grunbaum, D., and Nishizaki, M. T. (2011). Estimating 3D movements
from 2D observations using a continuous model of helical swimming. Bull.
Math. Biol. 73, 1358–1377. doi: 10.1007/s11538-010-9575- 7
Hecker, M., Schumann, W., and Volker, U. (1996). Heat-shock and general stress
response in Bacillus subtilis.Mol. Microbiol. 19, 417–428. doi: 10.1046/j.1365-
2958.1996.396932.x
Herzog, B., and Wirth, R. (2012). Swimming behavior of selected species of archaea.
Appl. Environ. Microbiol. 78, 1670–1674. doi: 10.1128/aem.06723-11
Humphries, S. (2013). A physical explanation of the temperature dependence of
physiological processes mediated by cilia and flagella. Proc. Natl. Acad. Sci.
U. S. A. 110, 14693–14698. doi: 10.1073/pnas.1300891110
Jaqaman, K., Loerke, D., Mettlen, M., Kuwata, H., Grinstein, S., Schmid, S. L., et al.
(2008). Robust single-particle tracking in live-cell time-lapse sequences. Nat.
Methods 5, 695–702. doi: 10.1038/nmeth.1237
Kearns, D. B., and Losick, R. (2005). Cell population heterogeneity during growth
of Bacillus subtilis.Genes Dev. 19, 3083–3094. doi: 10.1101/gad.1373905
Li, G., Bensson, J., Nisimova, L., Munger, D., Mahautmr, P., Tang, J. X., et al. (2011).
Accumulation of swimming bacteria near a solid surface. Phys. Rev. E Stat.
Nonlin. Soft Matter Phys. 84:041932. doi: 10.1103/PhysRevE.84.041932
Li, G., Tam, L. K., and Tang, J. X. (2008). Amplified effect of Brownian motion
in bacterial near-surface swimming. Proc. Natl. Acad. Sci. U. S. A. 105, 18355–
18359. doi: 10.1073/pnas.0807305105
Lopez, D., Vlamakis, H., and Kolter, R. (2009). Generation of multiple cell types
in Bacillus subtilis.Fems Microbiol. Rev. 33, 152–163. doi: 10.1111/j.1574-6976.
2008.00148.x
Maeda, K., Imae, Y., Shioi, J. I., and Oosawa, F. (1976). Effect of temperature
on motility and chemotaxis of Escherichia-coli.J. Bacteriol. 127, 1039–1046.
doi: 10.1128/jb.127.3.1039-1046.1976
Magariyama, Y., and Kudo, S. (2002). A mathematical explanation of an increase in
bacterial swimming speed with viscosity in linear-polymer solutions. Biophys. J.
83, 733–739. doi: 10.1016/s0006-3495(02)75204- 1
Mann, C., Yu, L., Lo, C. M., and Kim, M. (2005). High-resolution quantitative
phase-contrast microscopy by digital holography. Opt. Express 13, 8693–8698.
doi: 10.1364/opex.13.008693
Moliere, N., Hossmann, J., Schafer, H., and Turgay, K. (2016). Roleo f Hsp100/Clp
protease complexes in controlling the regulation of motility in Bacillus subtilis.
Front. Microbiol. 7:315. doi: 10.3389/fmicb.2016.00315
Mukherjee, S., and Kearns, D. B. (2014). “The structure and regulation of flagella
in Bacillus subtilis,” in Annual Review of Genetics, Vol. 48, ed. B. L. Bassler
(Bloomington: Indiana University), 319–340.
Nannapaneni, P., Hertwig, F., Depke, M., Hecker, M., Mader, U., Volker, U.,
et al. (2012). Defining the structure of the general stress regulon of Bacillus
subtilis using targeted microarray analysis and random forest classification.
Microbiology 158, 696–707. doi: 10.1099/mic.0.055434-0
Nishiyama, M., and Arai, Y. (2017). Tracking the movement of a single prokaryotic
cell in extreme environmental conditions. Methods Mol. Biol. 1593, 175–184.
doi: 10.1007/978-1- 4939-6927-2_13
Peterson, M. E., Daniel, R. M., Danson, M. J., and Eisenthal, R. (2007).
The dependence of enzyme activity on temperature: determination and
validation of parameters. Biochem. J. 402, 331–337. doi: 10.1042/BJ2006
1143
Pulschen, A. A., Mutavchiev, D. R., Culley, S., Sebastian, K. N., Roubinet, J.,
Roubinet, M., et al. (2020). Live imaging of a hyperthermophilic archaeon
reveals distinct roles for two ESCRT-III homologs in ensuring a robust
and symmetric division. Curr. Biol. 30:e2854. doi: 10.1016/j.cub.2020.0
5.021
Riekeles, M., Schirmack, J., and Schulze-Makuch, D. (2021). Machine learning
algorithms applied to identify microbial species by their motility. Life 11:44.
doi: 10.3390/life11010044
Rodrigues, M. F. V., Lisicki, M., and Lauga, E. (2021). The bank of swimming
organisms at the micron scale (BOSO-Micro). PLoS One 16:e0252291. doi:
10.1371/journal.pone.0252291
Schindelin, J., Arganda-Carreras, I., Frise, E., Kaynig, V., Longair, M., Pietzsch, T.,
et al. (2012). Fiji: an open-source platform for biological-image analysis. Nat.
Methods 9, 676–682. doi: 10.1038/nmeth.2019
Schneider, W. R., and Doetsch, R. N. (1977). Temperature effects on bacterial
movement. Appl. Environ. Microbiol. 34, 695–700. doi: 10.1128/aem.34.6.695-
700.1977
Schumann, W. (2003). The Bacillus subtilis heat shock stimulon. Cell Stress
Chaperones 8, 207–217.
Syvertsson, S., Wang, B., Staal, J., Gao, Y., Kort, R., and Hamoen, L. W.
(2021). Different resource allocation in a Bacillus subtilis population
displaying bimodal motility. J. Bacteriol. 203:e0003721. doi: 10.1128/JB.000
37-21
Taute, K. M., Gude, S., Tans, S. J., and Shimizu, T. S. (2015). High-throughput 3D
tracking of bacteria on a standard phase contrast microscope. Nat. Commun.
6:8776. doi: 10.1038/ncomms9776
Frontiers in Microbiology | www.frontiersin.org 13 April 2022 | Volume 13 | Article 836808
fmicb-13-836808 April 15, 2022 Time: 9:5 # 14
Dubay et al. Bacterial Motility at High Temperature
Turner, L., Ping, L., Neubauer, M., and Berg, H. C. (2016). Visualizing flagella while
tracking bacteria. Biophys. J. 111, 630–639. doi: 10.1016/j.bpj.2016.05.053
Wagner, T., Lipinski, H. G., and Wiemann, M. (2014). Dark field nanoparticle
tracking analysis for size characterization of plasmonic and non-plasmonic
particles. J. Nanopart. Res. 16:2419. doi: 10.1007/s11051-014-2419-x
Wallace, J. K., Rider, S., Serabyn, E., Kuhn, J., Liewer, K., Deming, J., et al. (2015).
Robust, compact implementation of an off-axis digital holographic microscope.
Opt. Express 23, 17367–17378. doi: 10.1364/oe.23.017367
Warth, A. D. (1978). Relationship between the heat resistance of spores and the
optimum and maximum growth temperatures of Bacillus species. J. Bacteriol.
134, 699–705. doi: 10.1128/jb.134.3.699-705.1978
Young, J. W., Locke, J. C. W., and Elowitz, M. B. (2013). Rate of environmental
change determines stress response specificity. Proc. Natl. Acad. Sci. U. S. A. 110,
4140–4145. doi: 10.1073/pnas.1213060110
Zottl, A., and Yeomans, J. M. (2019). Enhanced bacterial swimming speeds
in macromolecular polymer solutions. Nat. Phys. 15, 554–558. doi: 10.1038/
s41567-019- 0454-3
Conflict of Interest: The authors declare that the research was conducted in the
absence of any commercial or financial relationships that could be construed as a
potential conflict of interest.
Publisher’s Note: All claims expressed in this article are solely those of the authors
and do not necessarily represent those of their affiliated organizations, or those of
the publisher, the editors and the reviewers. Any product that may be evaluated in
this article, or claim that may be made by its manufacturer, is not guaranteed or
endorsed by the publisher.
Copyright © 2022 Dubay, Johnston, Wronkiewicz, Lee, Lindensmith and Nadeau.
This is an open-access article distributed under the terms of the Creative Commons
Attribution License (CC BY). The use, distribution or reproduction in other forums
is permitted, provided the original author(s) and the copyright owner(s) are credited
and that the original publication in this journal is cited, in accordance with accepted
academic practice. No use, distribution or reproduction is permitted which does not
comply with these terms.
Frontiers in Microbiology | www.frontiersin.org 14 April 2022 | Volume 13 | Article 836808
... We recently presented tracking results for DHM recordings of cultures of Bacillus subtilis swimming at different temperatures, achieved using the LAP tracker (Dubay et al., 2022b). To obtain the tracks, the holograms were reconstructed into Z stacks, then a maximum projection was taken through the Z stack to yield a higher signal-to-noise X,Y time series. ...
Article
Full-text available
Quantitative tracking of rapidly moving micron-scale objects remains an elusive challenge in microscopy due to low signal-to-noise. This paper describes a novel method for tracking micron-sized motile organisms in off-axis Digital Holographic Microscope (DHM) raw holograms and/or reconstructions. We begin by processing the microscopic images with the previously reported Holographic Examination for Life-like Motility (HELM) software, which provides a variety of tracking outputs including motion history images (MHIs). MHIs are stills of videos where the frame-to-frame changes are indicated with color time-coding. This exposes tracks of objects that are difficult to identify in individual frames at a low signal-to-noise ratio. The visible tracks in the MHIs are superior to tracks identified by all tested automated tracking algorithms that start from object identification at the frame level, particularly in low signal-to-noise ratio data, but do not provide quantitative track data. In contrast to other tracking methods, like Kalman filter, where the recording is analyzed frame by frame, MHIs show the whole time span of particle movement at once and eliminate the need to identify objects in individual frames. This feature also enables post-tracking identification of low-SNR objects. We use these tracks, rather than object identification in individual frames, as a basis for quantitative tracking of Bacillus subtilis by first generating MHIs from X, Y, and t stacks (raw holograms or a projection over reconstructed planes), then using a region-tracking algorithm to identify and separate swimming pathways. Subsequently, we identify each object's Z plane of best focus at the corresponding X, Y, and t points, yielding ap full description of the swimming pathways in three spatial dimensions plus time. This approach offers an alternative to object-based tracking for processing large, low signal-to-noise datasets containing highly motile organisms.
Article
Motivation Many organisms’ survival and behavior hinge on their responses to environmental signals. While research on bacteria-directed therapeutic agents has increased, systematic exploration of real-time modulation of bacterial motility remains limited. Current studies often focus on permanent motility changes through genetic alterations, restricting the ability to modulate bacterial motility dynamically on a large scale. To address this gap, we propose a novel real-time control framework for systematically modulating bacterial motility dynamics. Results We introduce MotGen, a deep learning approach leveraging Generative Adversarial Networks (GANs) to analyze swimming performance statistics of motile bacteria based on live cell imaging data. By tracking objects and optimizing cell trajectory mapping under environmentally altered conditions, we trained MotGen on a comprehensive statistical dataset derived from real image data. Our experimental results demonstrate MotGen’s ability to capture motility dynamics from real bacterial populations with low mean absolute error in both simulated and real datasets. MotGen allows us to approach optimal swimming conditions for desired motility statistics in real-time. MotGen’s potential extends to practical biomedical applications, including immune response prediction, by providing imputation of bacterial motility patterns based on external environmental conditions. Our short-term, in-situ interventions for controlling motility behavior offer a promising foundation for the development of bacteria-based biomedical applications. Availability MotGen is presented as a combination of Matlab image analysis code and a machine learning workflow in Python. Codes are available at https://github.com/bgmseo/MotGen, for cell tracking and implementation of trained models to generate bacterial motility statistics. Supplementary information Supplementary data are available at Bioinformatics online.
Article
Full-text available
Digital holographic microscopy provides the ability to observe throughout a large volume without refocusing. This capability enables simultaneous observations of large numbers of microorganisms swimming in an essentially unconstrained fashion. However, computational tools for tracking large 4D datasets remain lacking. In this paper, we examine the errors introduced by tracking bacterial motion as 2D projections vs. 3D volumes under different circumstances: bacteria free in liquid media and bacteria near a glass surface. We find that while XYZ speeds are generally equal to or larger than XY speeds, they are still within empirical uncertainties. Additionally, when studying dynamic surface behavior, the Z coordinate cannot be neglected.
Article
Full-text available
Significant technical challenges have limited the study of extremophile cell biology. Here we describe a system for imaging samples at 75°C using high numerical aperture, oil-immersion lenses. With this system we observed and quantified the dynamics of cell division in the model thermoacidophilic crenarchaeon Sulfolobus acidocaldarius with unprecedented resolution. In addition, we observed previously undescribed dynamic cell shape changes, cell motility, and cell-cell interactions, shedding significant new light on the high-temperature lifestyle of this organism.
Article
Full-text available
Unicellular microscopic organisms living in aqueous environments outnumber all other creatures on Earth. A large proportion of them are able to self-propel in fluids with a vast diversity of swimming gaits and motility patterns. In this paper we present a biophysical survey of the available experimental data produced to date on the characteristics of motile behaviour in unicellular microswimmers. We assemble from the available literature empirical data on the motility of four broad categories of organisms: bacteria (and archaea), flagellated eukaryotes, spermatozoa and ciliates. Whenever possible, we gather the following biological, morphological, kinematic and dynamical parameters: species, geometry and size of the organisms, swimming speeds, actuation frequencies, actuation amplitudes, number of flagella and properties of the surrounding fluid. We then organise the data using the established fluid mechanics principles for propulsion at low Reynolds number. Specifically, we use theoretical biophysical models for the locomotion of cells within the same taxonomic groups of organisms as a means of rationalising the raw material we have assembled, while demonstrating the variability for organisms of different species within the same group. The material gathered in our work is an attempt to summarise the available experimental data in the field, providing a convenient and practical reference point for future studies.
Article
Full-text available
To cope with sudden changes in their environment, bacteria can use a bet-hedging strategy by dividing the population into cells with different properties. This so-called bimodal or bistable cellular differentiation is generally controlled by positive feedback regulation of transcriptional activators. Due to the continuous increase in cell volume, it is difficult for these activators to reach an activation threshold concentration when cells are growing exponentially. This is one reason why bimodal differentiation is primarily observed from the onset of the stationary phase when exponential growth ceases. An exception is the bimodal induction of motility in Bacillus subtilis , which occurs early during exponential growth. Several mechanisms have been put forward to explain this, including double negative-feedback regulation and the stability of the mRNA molecules involved. In this study, we used fluorescence-assisted cell sorting to compare the transcriptome of motile and non-motile cells and noted that expression of ribosomal genes is lower in motile cells. This was confirmed using an unstable GFP reporter fused to the strong ribosomal rpsD promoter. We propose that the reduction in ribosomal gene expression in motile cells is the result of a diversion of cellular resources to the synthesis of the chemotaxis and motility systems. In agreement, single-cell microscopic analysis showed that motile cells are slightly shorter than non-motile cells, an indication of slower growth. We speculate that this growth rate reduction can contribute to the bimodal induction of motility during exponential growth. IMPORTANCE To cope with sudden environmental changes, bacteria can use a bet-hedging strategy and generate different types of cells within a population, so called bimodal differentiation. For example, a Bacillus subtilis culture can contain both motile and non-motile cells. In this study we compared the gene expression between motile and non-motile cells. It appeared that motile cells express less ribosomes. To confirm this, we constructed a ribosomal promoter fusion that enabled us to measure expression of this promoter in individual cells. This reporter fusion confirmed our initial finding. The re-allocation of cellular resources from ribosome synthesis towards synthesis of the motility apparatus results in a reduction in growth. Interestingly, this growth reduction has been shown to stimulate bimodal differentiation.
Article
Full-text available
(1) Background: Future missions to potentially habitable places in the Solar System require bio-chemistry-independent methods for detecting potential alien life forms. The technology was not advanced enough for onboard machine analysis of microscopic observations to be performed in past missions, but recent increases in computational power make the use of automated in-situ analyses feasible. (2) Methods: Here, we present a semi-automated experimental setup, capable of distinguishing the movement of abiotic particles due to Brownian motion from the motility be-havior of the bacteria Pseudoalteromonas haloplanktis, Planococcus halocryophilus, Bacillus subtilis, and Escherichia coli. Supervised machine learning algorithms were also used to specifically identify these species based on their characteristic motility behavior. (3) Results: While we were able to distinguish microbial motility from the abiotic movements due to Brownian motion with an ac-curacy exceeding 99%, the accuracy of the automated identification rates for the selected species does not exceed 82%. (4) Conclusions: Motility is an excellent biosignature, which can be used as a tool for upcoming life-detection missions. This study serves as the basis for the further devel-opment of a microscopic life recognition system for upcoming missions to Mars or the ocean worlds of the outer Solar System.
Article
Full-text available
Escherichia coli and many other bacteria swim through media with the use of flagella, which are deformable helical propellers. When the viscosity of media is increased, a peculiar phenomenon can be observed in which the organism's motility appears to improve. This improvement in the cell's swimming speed has previously been explained by modified versions of resistive force theory (RFT) which accounts for the interaction between flagella and molecules associated with the viscosity increase. Using optical tweezers, we measure the swimming force of individual E. coli in solutions of varying viscosity. By using probe-free force measurements, we are able to quantitatively validate and compare RFT and proposed modifications to the theory. We find that the force produced by the flagellum remains relatively constant even when the viscosity of the medium increases by approximately two orders of magnitude, contrary to predictions of RFT and variants. We conclude that the observed swimming forces can be explained by allowing the flagella geometry to deform as the viscosity of the surrounding medium is increased.
Article
Full-text available
Live-cell imaging has revolutionized our understanding of dynamic cellular processes in bacteria and eukaryotes. Although similar techniques have been applied to the study of halophilic archaea [1, 2, 3, 4, 5], our ability to explore the cell biology of thermophilic archaea has been limited by the technical challenges of imaging at high temperatures. Sulfolobus are the most intensively studied members of TACK archaea and have well-established molecular genetics [6, 7, 8, 9]. Additionally, studies using Sulfolobus were among the first to reveal striking similarities between the cell biology of eukaryotes and archaea [10, 11, 12, 13, 14, 15]. However, to date, it has not been possible to image Sulfolobus cells as they grow and divide. Here, we report the construction of the Sulfoscope, a heated chamber on an inverted fluorescent microscope that enables live-cell imaging of thermophiles. By using thermostable fluorescent probes together with this system, we were able to image Sulfolobus acidocaldarius cells live to reveal tight coupling between changes in DNA condensation, segregation, and cell division. Furthermore, by imaging deletion mutants, we observed functional differences between the two ESCRT-III proteins implicated in cytokinesis, CdvB1 and CdvB2. The deletion of cdvB1 compromised cell division, causing occasional division failures, whereas the ΔcdvB2 exhibited a profound loss of division symmetry, generating daughter cells that vary widely in size and eventually generating ghost cells. These data indicate that DNA separation and cytokinesis are coordinated in Sulfolobus, as is the case in eukaryotes, and that two contractile ESCRT-III polymers perform distinct roles to ensure that Sulfolobus cells undergo a robust and symmetrical division.
Article
Full-text available
We show that background fringe-pattern subtraction is a useful technique for removing static noise from off-axis holographic reconstructions and can enhance image contrast in volumetric reconstructions by an order of magnitude in the case for instruments with relatively stable fringes. We demonstrate the fundamental principle of this technique and introduce some practical considerations that must be made when implementing this scheme, such as quantifying fringe stability. This work also shows an experimental verification of the background fringe subtraction scheme using various biological samples.
Article
Full-text available
The locomotion of swimming bacteria in simple Newtonian fluids can successfully be described within the framework of low Reynolds number hydrodynamics. The presence of polymers in biofluids generally increases the viscosity, which is expected to lead to slower swimming for a constant bacterial motor torque. Surprisingly, however, several experiments have shown that bacterial speeds increase in polymeric fluids, and there is no clear understanding why. Therefore we perform extensive coarse-grained simulations of a bacterium swimming in explicitly modeled solutions of macromolecular polymers of different lengths and densities. We observe an increase of up to 60% in swimming speed with polymer density and demonstrate that this is due to a depletion of polymers in the vicinity of the bacterium leading to an effective slip. However this in itself cannot predict the large increase in swimming velocity: coupling to the chirality of the bacterial flagellum is also necessary.
Chapter
Many bacterial species move toward favorable habitats. The flagellum is one of the most important machines required for the motility in solution and is conserved across a wide range of bacteria. The motility machinery is thought to function efficiently with a similar mechanism in a variety of environmental conditions, as many cells with similar machineries have been isolated from harsh environments. To understand the common mechanism and its diversity, microscopic examination of bacterial movements is a crucial step. Here, we describe a method to characterize the swimming motility of cells in extreme environmental conditions. This microscopy system enables acquisition of high-resolution images under high-pressure conditions. The temperature and oxygen concentration can also be manipulated. In addition, we also describe a method to track the movement of swimming cells using an ImageJ plugin. This enables characterization of the swimming motility of the selected cells.