You are on page 1of 9

SCIENCE CHINA

Physics, Mechanics & Astronomy


•Article• September 2019 Vol. 62 No. 9: 994611
https://doi.org/10.1007/s11433-019-9393-x

Tensile properties of individual multicellular Bacillus subtilis fibers


1,2 1 1 1,2*
Xuan Ye , Tao Wang , Zhuo Zhuang , and XiDe Li
1
Applied Mechanics Laboratory, Department of Engineering Mechanics, Tsinghua University, Beijing 100084, China;
2
Center for Nano and Micro Mechanics, Tsinghua University, Beijing 100084, China

Received January 30, 2019; accepted March 14, 2019; published online April 20, 2019

Microfibers formed by Bacillus subtilis (B. subtilis) have attracted interest because of their potential for use as biodegradable
fibers. In this work, an efficient method based on the micro-liquid bridge method (LBM) is proposed to investigate the
mechanical properties and the deformation evolution in individual fibers. For the first time, tensile testing of fibers of this type
containing several cells is conducted in a scanning electron microscope (SEM) chamber and the in situ deformation evolution of
the fibers and the septa is observed. Experimental results show that these fibers are almost broken at the positions of the septa at
low humidity, but also show that their fracture morphologies are different. At high humidity, local necking deformation occurs at
the septum position. To explore the deformation mechanism of an individual bacterial fiber with a diameter of several hundred
nanometers under different humidity conditions, we use the finite element method (FEM) to analyze the tensile deformation
behavior of these fibers when their septa are at various separation levels. The numerical results indicate that weak interactions
among the septa lead to the dispersion of both the fibrous tensile strength and the modulus. These results may be helpful in
understanding the deformation mechanism, thus leading to further improvements in the mechanical performance of these fibers.
Bacillus subtilis, mechanical properties, micro-liquid bridge method, finite element method

PACS number(s): 46.80.+j, 62.20.-x, 62.25.+g, 68.08.-p

Citation: X. Ye, T. Wang, Z. Zhuang, and X. D. Li, Tensile properties of individual multicellular Bacillus subtilis fibers, Sci. China-Phys. Mech. Astron. 62,
994611 (2019), https://doi.org/10.1007/s11433-019-9393-x

1 Background lengths of approximately 2 µm, these new fibers can be up to


several hundred microns in length [4]. Additionally, these
With the rapid development of micro- and nano-science and bacterial fibers can be extracted from a culture solution to
technology, it is of great research significance to study the form a bacterial thread (composed of multiple fibers ar-
mechanical properties of micro- and nano-scale materials ranged in parallel) with a length of a few centimeters and a
and structures, and to predict and analyze the mechanical diameter of a few millimeters. Performing an in-depth study
properties and reliability of microelectromechanical systems of the mechanical behavior of the individual B. subtilis fiber
(MEMS) and devices [1-3]. The Bacillus subtilis (B. subtilis) will be highly important in promoting the application of
fiber has recently appeared as a new type of biological ma- these fibers as engineering materials.
terial with nontoxic, nonpolluting and biodegradable char- However, the small-scale dimensions of the individual B.
acteristics. An individual B. subtilis fiber is a chain of subtilis fibers range from a few micrometers up to a hundred
connected cells produced by inhibiting the cell wall separa- micrometers, making measurement of their mechanical re-
tion of B. subtilis. Unlike wild-type B. subtilis cells with sponse and performance challenging. Therefore, it is of
major theoretical and practical importance to provide reliable
*Corresponding author (email: lixide@mail.tsinghua.edu.cn) measurement methods and analytical models that can char-

© Science China Press and Springer-Verlag GmbH Germany, part of Springer Nature 2019 phys.scichina.com link.springer.com
X. Ye, et al. Sci. China-Phys. Mech. Astron. September (2019) Vol. 62 No. 9 994611-2

acterize the different factors influencing their deformations, subtilis fibers. We observed the deformation evolution pro-
including the fiber microstructures, their elastic and viscoe- cesses of the fibers in real time and found various fracture
lastic parameters, and environmental parameters. At the cell morphologies at the septum positions under low humidity
scale, Touhami et al. [5], Hoh and Schoenenberger [6], and conditions and necking evolution at the septum positions for
Zhao et al. [7] obtained the average spring constant of the the fibers under high humidity conditions. In particular, by
cell wall of different cells, while Vadillo-Rodriguez et al. [8- considering the complex associational behaviors between the
10] measured the viscoelastic parameters using atomic force biostructure and the fiber’s mechanical properties, along
microscopy (AFM). Tuson et al. [11] proposed a method with the failure processes under low and high humidity, we
called “cell length analysis of mechanical properties” applied the finite element method (FEM) to analyze the
(CLAMP) to measure the mechanical properties of a bac- tensile deformation behavior of these fibers when their septa
terium in vivo. Amir et al. [12] applied a controllable hy- were at various separation levels.
drodynamic force to bend the B. subtilis cells and thus
obtained the stresses that regulate the cell wall growth me-
chanism. At the bacterial thread scale, Thwaites et al. [13-15] 2 Experiments and results
stretched bacterial threads that contained 20000 parallel
bacterial fibers under various relative humidities using an Test system. Uniaxial tensile tests of the fibers were per-
instrument similar to standard machines (e.g., the Instron formed on our in-house-built multiscale material testing
material testing system) and obtained the elastic modulus of system (MMTS) in an SEM (Quanta 450 FEG, FEI, Hills-
the threads. Recently, Zhang et al. [16] measured the dy- boro, OR, USA); the system was composed of a microscale
namic mechanical properties of B. subtilis threads using a material testing module (m-MTM) and a nanoscale material
cantilever-probe system. testing module (n-MTM) [17]. The n-MTM contained a pair
While the mechanical behavior of the cells or the bacterial of symmetrical wedge-grip units and a specimen connection
threads has been widely measured experimentally, as de- and support unit (SCSU), which was manufactured on a si-
scribed above, previous studies involving AFM, CLAMP, licon wafer using well-known silicon deep etching technol-
and hydrodynamic force have mainly focused on studies of ogy. In the SCSU, the n-MTM was available in two sizes,
the mechanical properties of the cell walls or the membranes with specimen spacing of 10 μm and 100 μm. In the uniaxial
of a single cell, indicating that they are unsuitable for me- tensile tests, we first fixed the symmetrical wedge-grip unit
chanical testing of individual multicellular fibers containing to the support stages of the m-MTM and then installed the
several to hundreds of mutant cells. In addition, the test re- specimen on the clamping ends of the SCSU. Depending on
sults for bacterial threads are not suitable for prediction of the measurement method and the measurement environment
the mechanical properties of the cell walls or the individual used, the specimen clamping method could include gluing,
cellular fibers. It thus remains challenging to manipulate and van der Waals interactions, electron beam-induced deposi-
measure the properties of the individual fibers quantitatively tion (EBID), and capillary adhesion. Next, the double sup-
because of their nanoscale diameters and high slenderness port bars of the SCSU were cut off using a small rotating
ratios. diamond saw, and the m-MTM provided an external load and
Recently, we performed experiments that involved a force sensing function to achieve the required uniaxial
stretching individual fibers with an average diameter of stretching or compression of the specimen. The complete
0.7 µm and a length range of 25.7-254.3 µm (where the fi- experimental system is shown in Figure 1.
bers contained tens of bacterial cells up to several hundreds) Liquid bridge method. This material is moisture-sensi-
using an in-house-built multiscale material testing system tive and thus we naturally think of clamping using capillary
under an optical microscope (OM) [17, 18]. We found that force [19, 20]. However, because the previously developed
the mechanical properties of the fibers were strongly de- liquid drop method (LDM) [18] under an OM is relatively
pendent on the relative humidity. The tensile strengths of the complex and is not suitable for clamping of the test specimen
fibers were highly dispersed, which we speculated to be in an SEM environment, the LBM has been proposed here to
caused by the septa. Regrettably, we could not clearly ob- provide the clamping connection for the multicellular fibers.
serve the fibers’ deformation evolution due to the spatial The LBM implementation process is described as follows
resolution limitation of the OM. (see Figure 2). First, a small number of freeze-dried multi-
In order to finely measure the deformation and reveal the cellular fibers is added to deionized water to obtain a sus-
evolution process, in particular the local deformation me- pension with an appropriate concentration. Then, a tiny drop
chanism of the septa, the B. subtilis fibers containing several of the fiber suspension is transferred using a dropper pipette
cells were uniaxially stretched in the SEM. A novel method into the gap between the clamping ends of the n-MTM
called the liquid bridge method (LBM) was proposed in this (Figure 2(b) and (f)). At this stage, a liquid bridge is formed
paper for the clamping and alignment of multicellular B. in the gap by the adhesion between the liquid and the solid
X. Ye, et al. Sci. China-Phys. Mech. Astron. September (2019) Vol. 62 No. 9 994611-3

adhere to the clamping ends through the interface interaction


between them. In the experiments, we did not observe the
slippage or peeling off between the specimens and the
clamping ends. During this process, the fibers may be in a
pre-stretched state. In order to eliminate the initial tensile
stress generated by the LBM, we first relaxed the fiber to
release the initial tensile stress generated during the drying
process using the fine-tuning function of the m-MTM, then
stretched the fiber at a loading speed of 1.0 μm/s. The entire
stretching process was recorded by video recording. The
elongation of the fiber was obtained by analyzing the SEM
image sequences, and the load was calculated from the de-
flection of the tip of the force-sensing probe based on the
beam theory [18]; considering that the end of the probe was
fixed on one clamping end of the MTS, the deflection of the
tip was entirely caused by the tension of the fiber.
From these experiments, we also found that the flat
clamping ends of the n-MTM were suitable for clamping
Figure 1 (Color online) Experimental system. (a) Entire system under fibers with lengths of 30-500 μm (Figure 3(a)), while for
OM. (b) Nanoscale material testing module observed in an SEM [17]. shorter fibers (e.g., with a length of 10 μm), sharp clamping
ends should be used (Figure 3(b)). This is because a rela-
(the clamping ends of the n-MTM), along with the cohesion tively short and thick liquid bridge is formed between the
of the liquid itself. The flow characteristics in the liquid two flat clamping ends when the gap between the clamping
bridge between two rods pulled apart have been investigated ends is small (Figure 3(c)), and the fibers in the liquid bridge
[21]. In our experiment, as the liquid evaporates, it should will be broken or pulled off at the clamping ends because of
create a similar flow effect which then accounts for the the large capillary force that is provided in such a situation.
straightening of the fibers along the axial direction of the With regard to the sharp clamping ends, when the gap is
liquid bridge. At the beginning, we can obtain a fiber with small, a slender liquid bridge can be formed to realize
ends that adhered to the clamping ends of the n-MTM by clamping of shorter fibers (Figure 3(d)). It should be noted
surface tension (Figure 2(c) and (g)). To ensure that only a that this does not mean that the sharp clamping ends are only
single fiber was clamped, we need to control the con- suitable for short fibers; these ends are suitable for clamping
centration of the suspension and verified it by the observa- of fibers with lengths in the range from 10 to 500 μm.
tion in the SEM. It was worth mentioning that the entire The LBM is suitable for use with moisture-sensitive low-
process took only 15 s. Subsequently, the entire experimental dimensional materials (e.g., one-dimensional materials with
system remained stationary until the liquid at the clamping lengths ranging from a few microns to hundreds of microns
ends completely evaporated and the fiber was dried (Figure 2 and diameters on the micron and submicron scales). When
(d) and (h)). At this time, the fibers adhered to the clamping compared with the LDM, the LBM is easier and more effi-
ends by the combination of the adhesion which was similar cient, can be performed without the aid of a manipulator, and
to the colloid from the culture medium and van der Waals avoids complex procedures such as picking up, transferring
force. As the water evaporated, the fiber would strongly and clamping samples.

Figure 2 (Color online) LBM procedure. (a) Ultrasonically cleaned clamping ends of the n-MTM viewed under the OM. (b) The moment at which a tiny
drop of the fiber suspension was transferred to the clamping ends of the n-MTM. (c) Fiber caught by the clamping ends of the n-MTM under capillary force.
(d) A dried individual fiber. (e)-(h) Schematic diagrams corresponding to Figure 2(a)-(d), respectively.
X. Ye, et al. Sci. China-Phys. Mech. Astron. September (2019) Vol. 62 No. 9 994611-4

Figure 3 (Color online) (a) Long fiber with length of 203.6 μm clamped
at the flat clamping ends. (b) Short fiber with length of 28.4 μm clamped at
the sharp clamping ends. (c) Short and thick liquid bridge formed between
the flat clamping ends. (d) Slender liquid bridge formed between the sharp
clamping ends. Figure 4 (Color online) Stress-strain curves of B. subtilis fibers under
low and high humidity conditions. The mutation in the tensile curve derives
from the change in humidity (there are only two humidity states in this
Experimental results. Previous studies have shown that study).
the B. subtilis fibers are sensitive to humidity, but we cannot
establish a tunable humidity environment within the chamber
of a conventional SEM. Therefore, we regulated each B.
subtilis fiber’s dryness to simulate the specimen under low or
high humidity conditions. The method used to obtain a
specimen with low humidity is to naturally dry the fiber for
20 min after evaporation of the liquid bridge. In contrast, to
obtain the mechanical properties of fibers under high hu-
midity conditions, we performed the tensile tests in the SEM
vacuum chamber immediately after evaporation of the liquid
bridge. At this time, the fiber cells still contain water and it Figure 5 Tensile testing of B. subtilis fibers when using (a) flat clamping
will not be lost immediately during the short-time experi- ends and (b) sharp clamping ends in an SEM chamber.
ments. If the specimen is kept in high vacuum for a long
time, water will gradually lose. Figure 4 shows the typical showed significant elongation both in parts of the cell wall
tensile testing results obtained when using flat and sharp and at the septum positions, and ultimately the fibers necked
clamping ends in the SEM (see Figure 5). The tensile results at the septum positions (Figure 6(d)-(e)).
show that the multicellular fibers have profoundly different
mechanical properties that depend on the different humid-
ities of the fibers (there are only two humidity states in this 3 Finite element simulations of stretched B.
study because of the difficulty of controlling the humidity in subtilis fiber
the SEM) (see Figure 4). For the fibers with low humidity,
the experimental stress-strain curves were fitted using a The fibrous mechanical parameters described above, such as
linear spring model and had an average elastic modulus of the elastic modulus and the viscosity coefficient, are given
(8.7±1.2) GPa and fracture stress of (86.0±22.3) MPa. At based on theoretical models by combining the experimental
high humidity, these fibers were viscoelastic materials with a test results, which assume that the B. subtilis fibers are one-
certain degree of nonlinearity during the initial loading stage. dimensional continuous hollow cylinders. In fact, these fi-
We obtained an average elastic modulus E of (62.7±11.7) MPa bers exhibit a continuous cylindrical wall with invagination
and an average viscosity η of (83.4±32.7) MPa·s based on at the septum positions at varying separation levels. These
fitting of the experimental stress-strain curves using the theoretical models cannot adequately characterize the way in
Kelvin-Voigt model [18]. For the fibers with low humidity, which the septa affect the deformation and fracture of these
fracture only occured at the septum position, although the fibers. In particular, it is not possible to provide adequate
fracture morphologies of the cross-sections of multicellular information on the inelastic deformation evolution and the
fibers vary and include vertical fracture (Figure 6(a)), micro- fracture mechanism of the B. subtilis fibers using the ex-
curved fracture (Figure 6(b)) and large-curved fracture perimentally measured stress-strain curves alone. Here, we
(Figure 6(c)). For fibers with high humidity, viscoelasticity use the FEM to simulate the uniaxial stretching process of
dominated the fiber deformation process. These fibers the B. subtilis fibers.
X. Ye, et al. Sci. China-Phys. Mech. Astron. September (2019) Vol. 62 No. 9 994611-5

Figure 6 (Color online) Typical failures of B. subtilis fibers. (a)-(c) Scanning electron micrographs of fracture cross-sections of multicellular fibers with
low humidity, demonstrating that the B. subtilis fibers fracture at the septum positions with various fracture morphologies. (d)-(e) Scanning electron
micrographs of fibers with high humidity that have necked at the septum positions.

3.1 Cell division of B. subtilis and formation of a sep-


tum

Typical transmission electron microscopy (TEM) images


that illustrate the cell division process of B. subtilis are
shown in Figure 7, while a corresponding schematic diagram
is shown in Figure 8. Before cell division, the rod-shaped B.
subtilis is elongated to approximately twice its cell length by
inserting a new peptidoglycan (indicated by the red lines in
Figure 8) into the lateral cell wall. The septum (indicated by
the blue lines in Figure 8) then begins to grow inwards at the
middle of the double-length cell cylinder until the mother Figure 7 Typical TEM images of the B. subtilis cell division process. (a)
cell synthesizes a complete septum that contains two flat New peptidoglycan is synthesized, and the cylindrical part of the cell is
elongated. (b) A portion of a septum is synthesized in the middle of the
parallel plates with a thin interface layer between them mother cell. (c) A complete septum containing two flat parallel plates with
(Figure 7(a)-(c)). Finally, the mother cell splits into two a thin interface layer between them is synthesized. (d) The mother cell
daughter cells under the action of hydrolase [22] (Figure begins to divide under the action of hydrolase. (e) Two daughter cells with
partially connected ends. (f) The daughter cells when almost entirely se-
7(d)-(f)). In this study, an individual multicellular fiber is a parated [18].
chain of cells that is connected through these septa by in-
hibiting cell separation [23, 24]. In addition, the septa of the several septa at different separation levels (from left to right:
same multicellular fiber can have different separation levels. initial stage, late stage, late stage, and initial stage), as illu-
strated in Figure 9(e).
3.2 Finite element models of fibers at varying separa-
tion levels 3.3 Material properties
Previous studies have shown that cells exhibit different The fracture behavior of the fibers with low humidity was
mechanical behaviors at various stages in their physiological studied using the elastic-fracture constitutive and extended
processes [25]. To study the mechanical properties and the finite element method (XFEM) [26, 27]. The average elastic
deformation evolution of these multicellular fibers with septa modulus for the cell wall from the experiments is 8.7 GPa.
at different separation levels, we established four three-di- The intermediate interface layer is set to be softer than the
mensional finite element models, as shown in Figure 9(a)- cell wall (by 0.2 times) [28]. Because the fiber structure near
(d), which corresponded to the initial (Figure 7(c)), early the septum is similar to a circumferentially crack-notched
(Figure 7(d)), middle (Figure 7(e)), and late (Figure 7(f))
round bar [29], we use the stress intensity factor K I to de-
stages of the cell separation process, respectively. The geo-
scribe the fracture behavior of an individual fiber, which can
metric sizes of these fibers were obtained from the TEM
be calculated as follows:
images shown in Figure 7(c)-(f). Considering that an actual
fiber was connected via the septa at various separation levels, D F
K I = 1.72 1.27 , (1)
d d 3/2
we also established a typical fibrous model that contained
X. Ye, et al. Sci. China-Phys. Mech. Astron. September (2019) Vol. 62 No. 9 994611-6

Figure 10 (Color online) Calculated results for multicellular fibers with


septa at the (a) initial, (b) early, (c) middle, (d) late, and (e) multiple
different separation stages of the cell separation process under low hu-
midity conditions. The fibers all break at the septum positions, which is
consistent with the experimental results shown in Figure 6(a)-(c).

uniaxial tension and the calculated fracture morphologies of


Figure 8 (Color online) Schematic of B. subtilis cell division process the fibers are consistent with the experimental results, which
corresponding to the images in Figure 7. indicate that the septa are weak connections. The simulated
load-displacement curves and the stress-strain curves of the
fibers at the different separation stages can also be obtained
(Figure 11). We find that the fibrous stress-strain curves and
the corresponding mechanical properties (i.e., the elastic
modulus and the fracture strength) have high dispersibility
because of the local structural changes (septa) that occur in
the fibers.
The simulation results for the five fibrous models with
high humidity are shown in Figure 12, illustrating that these
fibers are elongated both in parts of the cell wall and at the
Figure 9 (Color online) Fibrous models for septa at the (a) initial, (b) septum positions, while the necking instability also occurs at
early, (c) middle, and (d) late separation stages in the cell separation pro-
cess. (e) Fibrous model containing several septa at different separation the septum positions. For the fibers in the later separation
stages (from left to right: initial stage, late stage, late stage, and initial stages (Figure 12(c) and (d)), the necking at the septum
stage). positions is more obvious. For the fiber that contains several
septa at different separation levels (initial-late-late-initial),
where D, d , and F denote the diameters of the cylinder and the calculated tensile result (Figure 12(e)) is close to the
the septum and the fracture force, respectively. The calcu-
experimental results (Figure 6(d) and (e)), which shows that
lated average fracture toughness is K IC = 33.6 kN/m3/2 and an uneven distribution of the septa accelerates necking fail-
the corresponding energy release rate is G IC = 0.2 N/m. Un- ure of the multicellular fiber.
der high humidity conditions, the viscoelastic constitutive The simulated stress-strain curves of multicellular fibers
relationship (the Kelvin-Voigt model) was used to study the under high humidity condition with multiple different se-
deformation behavior of the fibers. The elastic modulus and paration stages can also be obtained (Figure 13). The stress-
the viscosity coefficient of the cell wall are 62.7 MPa and strain curves of these fibers still show wide dispersion be-
83.4 MPa·s respectively. The finite element calculations are cause of the different separation levels of the septa within
then performed using uniaxial displacement loading (the these fibers.
details of the finite element model is shown in Appendix).

3.4 Numerical results


4 Discussion

For the fibers with low humidity, the experimental results In this study, a liquid bridge was formed between the two
show that the multicellular fibers all fracture at the septa, clamping ends of the n-MTM (Figure 14) by the adhesion of
while the fracture morphologies at the cross-sections of the the liquids to the side surfaces of the clamping ends. The side
multicellular fibers differ, as shown in Figure 6(a)-(c). We surface of the liquid presented a concave meniscus surface
speculate that it is related to the various separation levels of because of the adhesion of the solids (i.e., the clamping ends)
the septa. Here, the numerical results (Figure 10) again in- at both ends and the surface tension of the liquid. This
dicate that the fibers fracture at the septum positions under concave meniscus surface caused a pressure difference be-
X. Ye, et al. Sci. China-Phys. Mech. Astron. September (2019) Vol. 62 No. 9 994611-7

Figure 14 (Color online) Schematic of a liquid bridge between two


clamping ends.

composed of two parts, where one part represents the action


of the negative pressure in the liquid along the direction of
the liquid bridge axis in the wetted region of the solid, while
Figure 11 (Color online) Numerical results for multicellular fibers at the the other part represents the action of the surface tension on
initial, early, middle, late, and multiple different separation stages under the three-phase line along the liquid bridge axis, the capillary
low humidity conditions. The wide dispersion of the stress-strain curves is
caused by the different separation levels of the septa within the fibers. force between the two plates separated by a liquid bridge is
then given by
f = A p + lsin
= 2 Lt cos / D 0 + 2 (L + t )sin , (3)
in which A is the area over which the upper and lower par-
allel plate surfaces are wetted by the liquid, l is the wetted
area perimeter, L is the wetted length, and t is the wetted
thickness.
Figure 12 (Color online) Numerical results for multicellular fibers at (a) For the liquid bridge that is formed between the two silicon
initial, (b) early, (c) middle, (d) late, and (e) multiple different separation
stages under high humidity conditions. plates (i.e., the clamping ends), the water surface energy γ is
0.072 N/m, and θ is 10.0° (with an oxide thickness of ap-
proximately 3 nm) [32]. We assume that the wetted length L
is equal to the length of the clamping ends of the SCSU, and
that the wetted thickness t is equal to the thickness of the
clamping ends. Because the SCSU is manufactured on a si-
licon wafer using well-known silicon deep etching technol-
ogy, t is known to be 25 μm. For the flat clamping ends, L is
80 μm, and thus the value of the capillary force is 5.5 μN
when D0 is equal to 100 μm, and the value of capillary force
is 31.0 μN when D0 is equal to 10 μm. For the sharp
clamping ends, L is 4 μm, and the capillary force is thus
0.9 μN when D0 is equal to 100 μm, and the value is 2.1 μN
when D0 is equal to 10 μm. We thus obtain capillary force
ranges of 5.5-31.0 μN for the flat clamping ends and
0.9-2.1 μN for sharp clamping ends, where the latter range is
much smaller.
Figure 13 (Color online) Calculated stress-strain curves for fibers with
multiple different separation levels under high humidity conditions. From eq. (3), if D0 remains constant and L decreases, the
capillary force between the two silicon plates separated by
the liquid bridge is then also reduced. This explains why the
tween the interior and the exterior of the liquid. According to flat clamping ends cannot clamp the short specimen well
the Laplace-Young equation [30, 31], the hydrostatic pres- when compared with the sharp champing ends. Because of
sure difference is given by the excessive capillary force between the flat clamping ends,
p = 2 cos / D 0, (2) the fibers in the liquid bridge are broken (previous experi-
where γ is the liquid surface energy, θ is the contact angle, mental results indicate that the maximum breaking load of an
and D0 is the distance between the clamping ends. individual fiber is 12 μN). If L remains constant and D0 de-
Considering that the capillary force between two plates is creases, the capillary force is then increased, which explains
X. Ye, et al. Sci. China-Phys. Mech. Astron. September (2019) Vol. 62 No. 9 994611-8

why the specimen cannot be clamped when the two flat the B. subtilis fibers and providing the TEM images of B. subtilis cell di-
vision.
clamping ends are relatively closely spaced.
The mechanical properties of the multicellular B. subtilis
1 Z. P. Xu, and Q. S. Zheng, Sci. China-Phys. Mech. Astron. 61, 074601
fibers may be affected by different factors, including in-
(2018).
dividual organisms, changes in the loading rates, the biolo- 2 W. Yang, H. T. Wang, T. F. Li, and S. X. Qu, Sci. China-Phys. Mech.
gical pretreatment methods used, and in particular, the Astron. 62, 14601 (2019).
moisture contents and saturability of the cells in the fibers 3 W. Qiu, C. L. Cheng, R. R. Liang, C. W. Zhao, Z. K. Lei, Y. C. Zhao,
L. L. Ma, J. Xu, H. J. Fang, and Y. L. Kang, Acta Mech. Sin. 32, 805
and the retained septa with various separation levels, which (2016).
changes the composition and structure of cells (or the ske- 4 L. Zhao, J. Ye, J. Fu, and G. Q. Chen, Bioresour. Technol. 248, 238
leton of cells). The experimental results indicated that the (2018).
humidity seriously affected the mechanical properties of the 5 A. Touhami, B. Nysten, and Y. F. Dufrêne, Langmuir 19, 4539 (2003).
6 J. H. Hoh, and C. A. Schoenenberger, J. Cell Sci. 107, 1105 (1994).
fibers because of the presence of multiple hydrogen bonds 7 L. Zhao, D. Schaefer, and M. R. Marten, Appl. Environ. MicroBiol.
between the peptides of the peptidoglycan contained in the 71, 955 (2005).
fibers, which made the polymer networks rigid under dry 8 V. Vadillo-Rodriguez, T. J. Beveridge, and J. R. Dutcher, J. Bacteriol.
conditions. As the degree of hydration increased, water then 190, 4225 (2008).
9 V. Vadillo-Rodriguez, S. R. Schooling, and J. R. Dutcher, J. Bacteriol.
competed for the hydrogen bond sites, which made the net- 191, 5518 (2009).
works more flexible and resulted in reductions in both the 10 V. Vadillo-Rodríguez, and J. R. Dutcher, Soft Matter 7, 4101 (2011).
elastic modulus and the strength of the fibers. Both the ex- 11 H. H. Tuson, G. K. Auer, L. D. Renner, M. Hasebe, C. Tropini, M.
Salick, W. C. Crone, A. Gopinathan, K. C. Huang, and D. B. Weibel,
perimental and computational results reveal that the septa
Mol. MicroBiol. 84, 874 (2012).
among the pairs of microbial cells are the weak links in the 12 A. Amir, F. Babaeipour, D. B. McIntosh, D. R. Nelson, and S. Jun,
multicellular fibers where failures are most likely to occur, Proc. Natl. Acad. Sci. USA 111, 5778 (2014), arXiv: 1305.5843.
and the failure strength is dependent on the strength of the 13 J. J. Thwaites, and N. H. Mendelson, Int. J. Biol. Macromol. 11, 201
(1989).
weakest septum. On one hand, we hope to perform in-situ
14 J. J. Thwaites, U. C. Surana, and A. M. Jones, J. Bacteriol. 173, 204
stretching of single-cell and dual-cell in SEM to observe the (1991).
deformation evolution process of cell cylinder and the septa 15 J. J. Thwaites, and N. H. Mendelson, Proc. Natl. Acad. Sci. USA 82,
in real time and reveal their failure mechanism. On the other 2163 (1985).
16 X. Zhang, X. Ye, and X. Li, Meas. Sci. Technol. 27, 085006 (2016).
hand, for this type of fiber with diameter at such a small 17 X. Ye, Z. Cui, H. Fang, and X. Li, Sensors 17, 1800 (2017).
length scale, the size effect and surface effect also have a 18 X. Ye, L. Zhao, J. Liang, X. Li, and G. Q. Chen, Sci. Rep. 7, 46052
certain influence on the mechanical properties [33, 34]. (2017).
19 A. Lutfurakhmanov, G. K. Loken, D. L. Schulz, and I. S. Akhatov,
Appl. Phys. Lett. 97, 124107 (2010).
20 P. Lambert, F. Seigneur, S. Koelemeijer, and J. Jacot, J. Micromech.
5 Conclusions Microeng. 16, 1267 (2006).
21 W. Schwalb, T. W. Ng, J. K. K. Lye, O. W. Liew, and B. H. P. Cheong,
We performed direct stretching of individual B. subtilis fi- J. Colloid Interface Sci. 365, 314 (2012).
22 R. Ohnishi, S. Ishikawa, and J. Sekiguchi, J. Bacteriol. 181, 3178
bers in situ in an SEM chamber using the new developed (1999).
liquid bridge method. By taking the effects of the bios- 23 D. W. Adams, and J. Errington, Nat. Rev. Microbiol. 7, 642 (2009).
tructure into account, a finite element analysis revealed the 24 R. D. Turner, W. Vollmer, and S. J. Foster, Mol. MicroBiol. 91, 862
(2014).
reason for the wide dispersion in the fiber tensile behavior
25 M. P. Stewart, Y. Toyoda, A. A. Hyman, and D. J. Müller, Nat. Protoc.
under low and high humidity conditions. The numerical re- 7, 143 (2012).
sults showed that the different degrees of separation of the 26 N. Moës, J. Dolbow, and T. Belytschko, Int. J. Numer. Meth. Eng. 46,
septa contribute to changes in both the elastic modulus and 131 (2015).
27 J. H. Song, P. M. A. Areias, and T. Belytschko, Int. J. Numer. Meth.
the strength in these fibers, while the weakest septum
Eng. 67, 868 (2010).
dominated the mechanical properties of the individual fibers. 28 X. Zhou, D. K. Halladin, E. R. Rojas, E. F. Koslover, T. K. Lee, K. C.
We expect that our findings, at the level of the individual Huang, and J. A. Theriot, Science 348, 574 (2015).
bacterial fiber, could provide a mechanism to improve fi- 29 W. F. Brown, and J. E. Srawley, Plane Strain Crack Toughness Testing
of High-strength Metallic Materials (ASTM STP410, New York,
brous mechanical performance in engineering applications. 1966).
To be used as mass-produced materials, the weak septa in 30 T. Young, Philos. Trans. R. Soc. Lond. 95, 65 (1805).
these fibers must be reinforced locally to enhance the load- 31 G. Whyman, E. Bormashenko, and T. Stein, Chem. Phys. Lett. 450,
bearing capacity. 355 (2008).
32 R. Williams, and A. M. Goodman, Appl. Phys. Lett. 25, 531 (1974).
33 Y. P. Zhao, Nano and Mesoscopic Mechanics (Science Press, Beijing,
This work was supported by the National Natural Science Foundation of 2014).
China (Grant Nos. 11872035, 11472151, 11632010, and 11227202). The 34 C. Xu, T. Xue, W. Qiu, and Y. Kang, ACS Appl. Mater. Interfaces 8,
authors thank Prof. Guo-Qiang Chen and Dr. Liang Zhao for synthesizing 27099 (2016).
X. Ye, et al. Sci. China-Phys. Mech. Astron. September (2019) Vol. 62 No. 9 994611-9

Appendix The details of finite element model used: the Youngʼs modulus of the cell wall is 8.7 GPa,
Poissonʼs ratio is 0.45, and fracture energy is G IC = 0.2 N/m.
A1 Elastic model at low humidity The Youngʼs modulus of the interface layer is 0.2 times that
As mentioned above, B. subtilis fibers show obvious brittle of the cell wall. We adopt XFEM technology to simulate the
fracture characteristics at low humidity. In order to better crack propagation in the commercial finite element code
understand their mechanical behavior and whole deforma- Abaqus.
tion and fracture process, we established finite element
models of different cell separation stages and their combi- A2 Viscoelastic model at high humidity
nations. The profile of a typical finite element model is
shown in Figure a1. The following material parameters are B. subtilis fibers exhibit viscoelastic behavior at high hu-
midity. They are stretched to produce large deformations
accompanied by necking. We established a finite element
model with the same geometric parameters and boundary
conditions as those in the previous section. The viscoelastic
constitutive model (Kelvin-Voigt model) was used to capture
the large deformation and necking phenomena. The corre-
Figure a1 (Color online) The finite element model of a B. subtilis fiber sponding Young’s modulus is 62.7 MPa and viscosity para-
under a tension load. meter η is 83.4 MPa·s.

You might also like