Vortex reversal is a precursor of confined bacterial turbulence
Edited by Anna Balazs, University of Pittsburgh, Pittsburgh, PA; received July 28, 2024; accepted January 31, 2025
Significance
Biological and synthetic self-propelled entities, such as cultured cells, swimming bacteria, and active colloids, often exhibit complex collective motion. Controlling and rectifying such motion is crucial for developing microscopic active devices and sensors composed of swarming self-propelled particles. This study reveals how geometrical confinement transforms chaotic motion into a stabilized vortex and converts it into unsteady reversing configurations. The underlying mechanism is generic and thus applicable to various active systems. This work paves the way for design strategies for active devices grounded in robust theoretical insights.
Abstract
Active turbulence, or chaotic self-organized collective motion, is often observed in concentrated suspensions of motile bacteria and other systems of self-propelled interacting agents. To date, there is no fundamental understanding of how geometrical confinement orchestrates active turbulence and alters its physical properties. Here, by combining large-scale experiments, computer modeling, and analytical theory, we have identified a generic sequence of transitions occurring in bacterial suspensions confined in cylindrical wells of varying radii. With increasing the well’s radius, we observed that persistent vortex motion gives way to periodic vortex reversals, four-vortex pulsations, and then well-developed active turbulence. Using computational modeling and analytical theory, we have shown that vortex reversal results from the nonlinear interaction of the first three azimuthal modes that become unstable with the radius increase. The analytical results account for our key experimental findings. To further validate our approach, we reconstructed equations of motion from experimental data. Our findings shed light on the universal properties of confined bacterial active matter and can be applied to various biological and synthetic active systems.
Sign up for PNAS alerts.
Get alerts for new articles, or get an alert when an article is cited.
Interacting self-propelled particles, often termed active matter, exhibit a remarkable tendency to self-organization and the onset of collective behavior. Being intrinsically out-of-equilibrium, active matter systems exhibit a slew of collective phenomena such as the spontaneous onset of long-range order (1–5), odd viscoelasticity (6), rectifications of chaotic flows (7–10), and reduction of the effective viscosity (11, 12). One of the most visible manifestations of collective dynamics in active matter systems is the emergence of self-sustained spatiotemporal chaotic flows termed active turbulence (13–18). In stark contrast to conventional Navier–Stokes turbulence, active turbulence, occurring for essentially zero Reynolds numbers, is characterized by the well-defined characteristic length scale. In the case of bacterial turbulence, this scale corresponds to typical vortex size, which is about 40 to 50 m (14, 15). The existence of the typical vortex size allows transforming bacterial motion into stable vortex arrays under geometrical confinements (19–24) or in the presence of periodic obstacles (9, 10).
Experimental and computational studies of self-organization of bacterial and related active systems have shown that strong confinement, e.g., a cylindrical well, may suppress active turbulence and generate persistent vortex motion (19, 20, 22–24). However, a fundamental question on the nature of the transition from ordered states under strong confinement to chaotic motion in unconstrained systems remains open. Answering this question will shed light on intricate fundamental mechanisms of self-organization in a broad class of active systems under confinement.
In the context of active nematics exemplified by microtubules-motors assays, multiple experimental and numerical studies interrogated a transition from ordered quasi-stationary states to chaotic motion that occurs under the confinement in channels, rings, and wells (25–32). The primary observation is that the instability of static nematic configuration occurs via unbinding and subsequent chaotic motion of half-integer topological defects. Moreover, in the context of cytoplasmic streaming (31, 32), the onset of spontaneous circulation and consequent periodic modulations of the circulation rate were observed for the system confined in a cylindrical well. More specifically, the analysis predicted a supercritical instability of steady-state vortex with the increase in the well radius. The instability occurred via the gradual unbinding of two half-integer nematic defects with the amplitude of the oscillations vanishing at the critical radius.
However, it is unclear how these insights could be projected on polar systems, e.g. suspensions of swimming bacteria, where the polar symmetry of the systems would prohibit the above scenario. Also, in the bacterial suspensions, experimental investigations have been hindered by the difficulty in resolving the detailed dynamics very close to the transition point and the necessity of long-time measurements for evaluating the vortex stability.
Here, we examine the route to active turbulence by combining large-scale experiments, high-resolution numerical modeling, and analytical theory. We focused on a well-characterized active system: suspensions of swimming bacteria (5). We confined the suspensions into an array of isolated cylindrical wells comparable to the size of individual vortices. We systematically varied the wells’ radii to characterize the transition from stabilized vortices to bacterial turbulence. Increasing the well radius, we have detected reversals of vortex rotation as the first instability from a stable vortex. The reversals were also captured as periodic oscillations in our numerical simulations and analytical theory, unraveling a robust fundamental mechanism for the onset of polar active turbulence. In stark contrast to nematic systems (31, 32) where a local Hopf bifurcation results in small-amplitude periodic modulations of the steady-state circulation that keeps the same rotation direction (no reversals), in our case, the reversals occur via a global subcritical infinite period bifurcation with hysteresis, where the reversal period diverges at the threshold. Four-vortex pulsations follow the vortex reversal with a further increase in the radius. The observed transitions differ from the reversals caused by the viscoelasticity of the suspending fluid (33) or density gradients (34). Our analysis revealed that the reversal originate from the nonlinear interaction of the three lowest azimuthal modes near the linear instability threshold. To validate our theoretical arguments, we reconstructed equations of motion from experiential data. Our studies indicate that the vortex reversal is a generic precursor of turbulence-like behavior in bacterial and related active systems. Our findings provide insights into how geometrical confinement orchestrates spatiotemporal organization in a broad class of active systems.
Results
Experiment.
We conducted experiments with suspensions of swimming bacteria confined in cylindrical wells, Fig. 1A and B. The height of the wells was set to 30 , which is smaller than the typical length scale of collective motion, ensuring effectively two-dimensional dynamics within each well. This constrained the range of radii for observation (SI Appendix, Supplementary Note 1G). Experiments were conducted simultaneously in an array of isolated wells of different radii (400 wells in total); see Fig. 1D and SI Appendix, Fig. S1, and Movies S1–S10. We observed stabilized vortices with steady rotational directions within the wells with small radii. For larger radii, the vortices exhibited a transition to unsteady configurations with reversing rotation directions. This observation is exemplified by the instantaneous vorticity field shown in Fig. 1C, where is the unit vector in the -direction. As one sees from Fig. 1C, the smaller wells hosted a single stabilized vortex with persistent rotation, with the velocity and vorticity profiles shown in Figs. 1E and 2A. In contrast to previous studies on bacterial suspensions confined in water-in-oil droplets (19, 20), we did not observe any counterrotating edge flows, suggesting different boundary conditions for collective motion.
Fig. 1.

Movie S1.
Movie of the whole field of view (2.1 mm × 2.1 mm), corresponding to Fig. S1. The initial 20 seconds of the data are presented at twice the real speed (2×). The spatial resolution is lowered to reduce the file size.
Movie S2.
The vorticity and velocity fields of the single stabilized vortex (R = 44.6 μm). The color bar of the vorticity field is the same as in Fig. 1. The movie is played at real-time speed. The scale bar represents 20 μm.
Movie S3.
The vorticity and velocity fields of the reversing vortices (R = 46.7 μm). The color bar of the vorticity field is the same as in Fig. 1. The movie is played at real-time speed. The scale bar represents 20 μm.
Movie S4.
The vorticity and velocity fields of the four-vortex state (R = 48.8 μm). The color bar of the vorticity field is the same as in Fig. 1. The movie is played at real-time speed. The scale bar represents 20 μm.
Movie S5.
The vorticity and velocity fields of the turbulent state (R = 49.5 μm). The color bar of the vorticity field is the same as in Fig. 1. The movie is played at real-time speed. The scale bar represents 20 μm.
Movie S6.
The velocity field of the single stabilized vortex (R = 44.6 μm). The velocity vectors are colored based on the sign of the local vorticity: yellow for positive and green for negative. The movie is played at 4× real-time speed. The scale bar represents 20 μm.
Movie S7.
The velocity field of the reversing vortices (R = 46.7 μm). The velocity vectors are colored based on the sign of the local vorticity: yellow for positive and green for negative. The movie is played at 4× real-time speed. The scale bar represents 20 μm.
Movie S8.
The velocity field of the four-vortex state (R = 48.8 μm). The velocity vectors are colored based on the sign of the local vorticity: yellow for positive and green for negative. The movie is played at 4× real-time speed. The scale bar represents 20 μm.
Movie S9.
The movie of the non-coherent random motion in the small well (R = 42.9 μm). The movie is played at real-time speed. The scale bar represents 20 μm.
Movie S10.
The vorticity and velocity fields of the reversing vortices (R = 48.0 μm). The color bar of the vorticity field is the same as in Fig. 1. The movie is played at real-time speed. The scale bar represents 20 μm.
To quantify the vortex rotation direction, we defined a spin variable for each well as,
[1]
where is the center of the -th well, and the summations run over the area of the -th well. As shown in Fig. 1 E and F, the spins for the small wells stayed almost constant and rarely flipped their signs over time, while the spins for the larger wells persistently alternated their signs, reflecting the reversals of vortices. The spin probability distribution for such a well with reversals exhibits a bimodal distribution, indicating the presence of two states with clockwise (CW, ) and counterclockwise (CCW, ) rotations (Fig. 2B); see SI Appendix, Fig. S6, Supplementary Note 1H, and Movie S10 for a well exhibiting faster reversals. Contrary to ref. (24), our setup uses two symmetric surfaces for the top and bottom to compensate for systematic bias in the rotation direction. Thus, the fraction of CW rotations as a function of the well’s radius was always 0.5, Fig. 2D. The absence of bias is crucial for characterizing the vortex reversals. The transition from a single stabilized vortex to reversing vortices was inspected through the spin correlation time, defined as the time at which the autocorrelation function of the spins decayed to ; see SI Appendix, Fig. S4. The correlation time has successfully captured the transition at the radii of approximately 46 to 48 (Fig. 2C). For large wells’ radii, four-vortex pulsating states were observed as well, Fig. 1C. The pulsation was characterized by the kinetic energy of azimuthal modes corresponding to vortices within a well (SI Appendix, Supplementary Note 4),
[2]
Fig. 2.

By this mode analysis, we probed antiphase oscillation of the modes and ; see Fig. 1G.
Further increase in the radius destabilized the four-vortex pulsations, resulting in chaotic turbulent flow (Fig. 1C and SI Appendix, Fig. S5). For radii much smaller than those for the stabilized single-vortex state, such stable vortex formation was suppressed and we observed random-like motion of individual bacteria; see SI Appendix, Fig. S4 and Movie S9.
Computational Modeling.
We performed numerical simulations using a phenomenological active fluid model, the Toner–Tu–Swift–Hohenberg equation (TTSHE) (5, 10, 15–17). The TTSHE qualitatively captures the bulk properties of polar active turbulence. It can describe the transformation of bacterial turbulence into stable vortex arrays in the presence of periodic obstacles (9, 10) and has been used to investigate the instability of the emergent order (35). In the vorticity representation, the dimensionless TTSHE is of the form (10):
[3]
where , , and are constants, is a scalar field that dampens and outside the well () without affecting the inside (), and are damping coefficients. In this dimensionless form, the vortex characteristic size is . Following ref. 10, we adopt the parameter values and impose three boundary conditions on well’s wall,
[4]
Compared with the Navier–Stokes equation, the extra boundary condition is imposed due to the higher-order differential operator () in Eq. 3. We solved Eq. 3 with the above boundary conditions in two dimensions by the pseudospectral method; see Methods.
Our simulations successfully reproduced the entire sequence of transitions observed in experiments, Fig. 3A and Movies S11–S17. We have found a single stable vortex for small radii. As the radius increases, the vortex becomes destabilized via infinite period bifurcation with hysteresis and yields a periodically reversing two-vortex state; see Fig. 3 B and C. It was demonstrated by the time series of the spin variable; see Fig. 3B. The infinite period bifurcation scenario (36) is consistent with the dependence of the correlation time in the experiment, Fig. 2C, which gradually decreases with the increase of the radius. Further increasing the radius, the reversing two-vortex state transforms into a pulsating four-vortex configuration, similarly to the experiment, Fig. 3D.
Fig. 3.

Movie S11.
Vorticity field of the single stabilized vortex (Fig. 3(a), left), numerically computed at R = 5.2.
Movie S12.
Vorticity field of the reversing two-vortex state (Fig. 3(a), the second from the left), numerically computed at R = 5.6.
Movie S13.
Vorticity field of the pulsating four-vortex state (Fig. 3(a), the third from the left), numerically computed at R = 6.4.
Movie S14.
Vorticity field of the turbulent state (Fig. 3(a), right), numerically computed at R = 7.6.
Movie S15.
Vorticity field of the reversing two-vortex state(Fig. 3(b)), numerically computed at R = 5.35.
Movie S16.
Azimuthal Fourier decomposition of the vorticity field of the reversing two-vortex state at R = 5.6 (Fig. 3(c)). The original vorticity field and the modes n = 0, 1, 2 are shown.
Movie S17.
Vorticity field of the pulsating four-vortex state (Fig. 3(d)), numerically computed at R = 6.2.
Weakly Nonlinear Analysis.
We examined the linear stability of Eq. 3 around , yielding . Its solution is of the form ,
[5]
where are the growth rates of the corresponding azimuthal modes, are constants, are the Bessel functions, . Applying the boundary conditions to and solving the characteristic equation (SI Appendix, Supplementary Note 4), one finds the growth rates vs. radius . The results are shown in Fig. 4A. For small enough , all are negative, so that no vortex is excited, similarly to our experimental observation of noncoherent, random motion of individual bacteria for small wells. For , becomes positive, corresponding to the onset of the steady-state vortical motion observed in computational modeling and experiment; see SI Appendix, Fig. S11B for quantitative agreement between the numerical and analytical solutions. Then, with the gradual increase in , higher rotational modes become unstable. We find that vortex reversal occurs at , when the first two modes () are unstable, and the mode is still stable but close to the threshold. This mode turns out to be excited due to nonlinear couplings with the modes , compare Fig. 4 A and E and see SI Appendix, Supplementary Note 4. The period of reversals diverges at the threshold radius, consistently with the infinite period bifurcation scenario; see SI Appendix, Fig. S10.
Fig. 4.

The radial vorticity and velocity profiles predicted by the linear analysis, Eq. 5, are in excellent agreement with the numerical solutions of Eq. 3 without any fittings, Fig. 4 B and C. Furthermore, fitting the theoretical expression, Eq. 5, for to experimental vorticity and velocity profiles of a stable vortex provides an excellent approximation as well; see Fig. 2A and SI Appendix, Supplementary Note 1I.
Next, we approximate the solution to Eq. 3 as a sum of the three lowest azimuthal modes with ,
[6]
Here, are the eigenfunctions obtained from linear stability analysis. For definiteness, the eigenmodes are normalized by their kinetic energy; see SI Appendix, Supplementary Note 4 and Eqs. S35 and S36. are slowly varying amplitudes that are derived from the corresponding orthogonality leading to a set of normal form Eqs. 7–9; see Methods.
Eqs. 7–9 faithfully reproduce the numerical results from Eq. 3 without further approximation; see Fig. 4. Specifically, for small radii, Eqs. 7–9 reproduce a stable vortex solution as shown in Fig. 3A. Then, with the increase in , the infinite period bifurcation to a limit cycle is faithfully captured. Furthermore, even the details of the time dependence of each azimuthal mode closely agree with those of the numerical solutions of Eq. 3; see Fig. 4D and SI Appendix, Fig. S11 C and D. In the reversing vortex state, displayed in Fig. 4E, all three amplitudes are nonzero; see Movie S18. With the further increase in , a transition from a reversing state to a pulsating four-vortex solution occurs; see Fig. 4F and Movie S19. Here, the zero mode, , is suppressed, and the first and second modes pulsate in antiphase. As shown in Fig. 1G, this antiphase relation was indeed observed experimentally, further demonstrating the quantitative agreements among the experimental, numerical, and analytical results. The normal form analysis indicates that the transition to vortex reversals and other time-dependent states is a result of resonant nonlinear interaction among the three lowest azimuthal modes. This behavior only exists for sufficiently large values of the nonlinear advection term in Eq. 3, which controls the resonant three-mode interaction. No limit cycles were found for .
Movie S18.
Vorticity field and its azimuthal Fourier decomposition of the analytically calculated reversing two-vortex state at R = 5.9 (Fig. 4(e)). The definitions of the modes are the same as in Supplementary Movie 12 and Fig. 3(c).
Movie S19.
Vorticity field and its azimuthal Fourier decomposition of the analytically calculated pulsating four-vortex state at R = 7.0 (Fig. 4(f)). The definitions of the modes are the same as in Supplementary Movie 12 and Fig. 3(c).
Validating Equations of Motion.
The use of the TTSHE was validated through regression analysis of our experimental data; see SI Appendix, Supplementary Note 2 for details. Similar approaches were used in refs. 37 and 38. In addition to the TTSHE, we tested another model for bacterial turbulence, the Nikolaevskiy equation, which includes term but no cubic nonlinearity nor linear term (39–41). The TTSHE outperformed the Nikolaevskiy equation in terms of the residuals, justifying our numerical and theoretical approaches; see SI Appendix, Figs. S7–S9 and Tables S1–S3. The regression for the two-vortex reversing state shown in Fig. 1F yields for the dimensional TTSHE, proving the presence of the advection term with , larger than for the Naiver-Stokes equation. Transforming the TTSHE into the form of Eq. 3 with characteristic values in the unconstrained bacterial turbulence (velocity 50 , length scale 40 , and time scale 0.5 s) yields in the dimensionless TTSHE. It is consistent with our theoretical prediction of for the onset of oscillations.
Concluding Remarks
We observed a generic route to active turbulence in confined suspensions of swimming bacteria: a single steady vortex gives way to a reversing vortex pair, four pulsating vortices, and then to a well-developed spatiotemporal chaos. The fact that the entire bifurcation sequence is reproduced by a generic phenomenological model for active turbulence reveals the universal fundamental mechanism governing the transition: resonant interaction of the three lowest azimuthal modes associated with cylindrical confinement. Furthermore, the onset of the periodic reversal relies on the finite value of the Navier–Stokes-like advection term in the phenomenological model of active turbulence (10, 15, 17, 42). The regression of experimental data also reliably corroborates the presence of the advection term with its coefficient in the effective equation. These findings suggest that the observed transitions should also occur in a broad class of active self-propelled systems under confinement. This robust mechanism is presumably responsible for the onset of reversing edge currents numerically observed in ref. 43 and is not sensitive to the details of boundary conditions or geometry (42). Furthermore, the observed transitions occur in a Newtonian fluid environment with homogeneous activity and density. Viscoelasticity or anisotropy may only affect the details of the transitions (33, 44, 45). The oxygen supply through the top and bottom polydimethylsiloxane (PDMS) membranes realized the homogeneity of the system, excluding the previously proposed reversal mechanism driven by density gradients arising from nonuniform oxygen supply limited to the circumference (34). This generic mechanism is based on the three-mode resonant interaction and should be relevant for the variety of biological and synthetic active systems, e.g., Janus colloids (3, 4, 46).
Another intriguing aspect is the effect of chirality. Since bacteria are chiral objects due to counterrotation of the body and helical flagella (47–50), there could be an asymmetry between CW/CCW rotating vortices (24). In this work, a sustained effort was undertaken to make the upper/bottom surfaces of the wells as identical as possible to suppress the asymmetry. While a minor chiral shift does not affect the transition sequence, it could introduce slightly different thresholds for the onsets of vortex oscillations of opposite chirality.
The current experiment is unavoidably susceptible to strong fluctuations discarded in the theoretical description. For example, the number of bacteria within a single microscopic well is about 10 bacteria/well. The dynamics of such a small bacterial population is intrinsically stochastic. Therefore, understanding how noise affects the nature of transitions and exploring ways to tame and control the fluctuating active dynamics would be of interest to future studies.
Finally, the controls and rectification of vortices in confined active matter open up possibilities for engineering out-of-equilibrium systems. For instance, weak coupling between neighboring wells may realize a “bacterial lattice clock,” in which reversing vortex pairs synchronize and exhibit higher regularity and persistence. The reversing or pulsating vortices may be useful for mixing at low Reynolds numbers. Taming the fluctuations in active systems based on the fundamental instability uncovered in this work provides design principles for functioning active devices, such as biosensors or microrobotic swarms for targeted drug delivery, precision surgery, or detoxification (51, 52).
Methods
Materials and Methods
Experimental Details.
Bacteria Bacillus subtilis (strain: 1085) were grown in Terrific Broth (T9179, Sigma-Aldrich) growth medium until optical density (OD) achieved 1. After concentrating the suspension 180-fold, it was sandwiched between two thin PDMS membranes to facilitate sufficient oxygen supply for sustaining high bacterial motility. The bottom PDMS membrane was patterned with 30--deep multiple microscopic wells with the radius ranging from 44 to 51 with 0.5 increments (Fig. 1A, B, and D and SI Appendix, Figs. S1 and S2). To overcome systematic errors arising from different preparations of bacterial cultures and slight density variations caused during the confinement process, we simultaneously observed 400 wells (19 radii, 20 wells for each radius) in a single field of view by using an inverted microscope equipped with a large-sensor sCMOS camera (Kinetix, Teledyne Photometrics, 3,200 3,200 pixels) and a 10 objective lens, realizing the 2.1 mm 2.1 mm field of view (SI Appendix, Fig. S1). It allowed resolving the bacterial dynamics very close to the transition point. We captured the movies at 50 fps for 150 s and analyzed the bacterial velocity fields using the particle image velocimetry (PIV). The duration was limited to 150 s to ensure statistical stationarity, which is eventually spoiled by the gradual decrease in bacterial activity. See SI Appendix, Supplementary Note 1A–D for the detailed protocols.
Computational Details.
Eq. 3 was solved by the pseudospectral method in a two-dimensional periodic domain, discretized as the 8,192 8,192 square lattice. Spatial derivatives were handled by the fast Fourier transform; see SI Appendix, Supplementary Note 3. Time update was performed in the Fourier space, with the time step . To accelerate simulations, we performed the whole computation on GPUs (NVIDIA RTX A6000 or A100).
The damping wall implemented in Eq. 3 with the kernel permits some leakage outside of the well radius . We calibrated to account for the leakage and defined the effective radius , where the velocity and vorticity vanish. is calculated as the root of (the zeroth azimuthal mode); see SI Appendix, Fig. S11. We obtained . The zeroth mode amplitude in (Eq. 6) provides a convenient measure of the spin ((Eq. 1)) up to a certain prefactor. For the details of the numerical mode decomposition and related quantities, see SI Appendix, Supplementary Note 4.
Normal form Equations.
Substituting Eq. 6 into Eq. 3, and implementing the orthogonality conditions, we obtain the set of equations for amplitudes ,
[7]
[8]
[9]
where are the linear growth rates; other coefficients are integrals over the nonlinearities, SI Appendix, Supplementary Note 4. All coefficients are calculated using a Mathematica script provided as a SI Appendix.
Data, Materials, and Software Availability
All the relevant experimental and numerical data, the MATLAB codes for analyzing the experimental data, Python scripts for the numerical simulations of the TTSHE, and the Mathematica code for analytical theory are deposited on Zenodo (53). Although only the initial parts of the experimental movies were deposited due to the size limit of the repository, the complete image sequences can be obtained by contacting the authors at [email protected]. The Mathematica code for analytical theory is also included in SI Appendix.
Acknowledgments
This work is supported in part by Japan Society for the Promotion of Science (JSPS) KAKENHI Grant Numbers JP19H05800, JP20K14426, JP23K25838, JP24K00593, JP24KJ0890, Japan Science and Technology Agency (JST) PRESTO Grant Number JPMJPR21O8, JST FOREST Grant Number JPMJFR2364, Japan, and the JSPS Core-to-Core Program “Advanced core-to-core network for the physics of self-organizing active matter (JPJSCCA20230002).” The research of I.S.A. was supported by the NSF award PHY-2140010. I.S.A. also acknowledges the visiting professorship for the Global Science Graduate Course–B, at The University of Tokyo, where the main part of this work was done. S.S. acknowledges support from The Forefront Physics and Mathematics Program to Drive Transformation, World-leading Innovative Graduate Study Program, the University of Tokyo. The computation in this paper was done partly by the nekoya/ai cluster at the Institute for Physics of Intelligence, The University of Tokyo, and the facilities of the Supercomputer Center, the Institute for Solid State Physics, The University of Tokyo.
Author contributions
D.N. conceived the project; D.N. initiated the project; D.N. performed the experiments; D.N. analyzed the experimental data; S.S. performed the numerical simulations; S.S. analyzed the numerical data; I.S.A. developed the analytical theory; D.N., S.S., K.A.T., and I.S.A. discussed the results; D.N., S.S., K.A.T., and I.S.A. interpreted the results; and D.N., S.S., K.A.T., and I.S.A. wrote the paper.
Competing interests
The authors declare no competing interest.
Supporting Information
Appendix 01 (PDF)
- Download
- 6.48 MB
Movie S1.
Movie of the whole field of view (2.1 mm × 2.1 mm), corresponding to Fig. S1. The initial 20 seconds of the data are presented at twice the real speed (2×). The spatial resolution is lowered to reduce the file size.
- Download
- 52.25 MB
Movie S2.
The vorticity and velocity fields of the single stabilized vortex (R = 44.6 μm). The color bar of the vorticity field is the same as in Fig. 1. The movie is played at real-time speed. The scale bar represents 20 μm.
- Download
- 45.43 MB
Movie S3.
The vorticity and velocity fields of the reversing vortices (R = 46.7 μm). The color bar of the vorticity field is the same as in Fig. 1. The movie is played at real-time speed. The scale bar represents 20 μm.
- Download
- 46.35 MB
Movie S4.
The vorticity and velocity fields of the four-vortex state (R = 48.8 μm). The color bar of the vorticity field is the same as in Fig. 1. The movie is played at real-time speed. The scale bar represents 20 μm.
- Download
- 44.71 MB
Movie S5.
The vorticity and velocity fields of the turbulent state (R = 49.5 μm). The color bar of the vorticity field is the same as in Fig. 1. The movie is played at real-time speed. The scale bar represents 20 μm.
- Download
- 47.29 MB
Movie S6.
The velocity field of the single stabilized vortex (R = 44.6 μm). The velocity vectors are colored based on the sign of the local vorticity: yellow for positive and green for negative. The movie is played at 4× real-time speed. The scale bar represents 20 μm.
- Download
- 41.90 MB
Movie S7.
The velocity field of the reversing vortices (R = 46.7 μm). The velocity vectors are colored based on the sign of the local vorticity: yellow for positive and green for negative. The movie is played at 4× real-time speed. The scale bar represents 20 μm.
- Download
- 42.69 MB
Movie S8.
The velocity field of the four-vortex state (R = 48.8 μm). The velocity vectors are colored based on the sign of the local vorticity: yellow for positive and green for negative. The movie is played at 4× real-time speed. The scale bar represents 20 μm.
- Download
- 41.78 MB
Movie S9.
The movie of the non-coherent random motion in the small well (R = 42.9 μm). The movie is played at real-time speed. The scale bar represents 20 μm.
- Download
- 24.09 MB
Movie S10.
The vorticity and velocity fields of the reversing vortices (R = 48.0 μm). The color bar of the vorticity field is the same as in Fig. 1. The movie is played at real-time speed. The scale bar represents 20 μm.
- Download
- 39.68 MB
Movie S11.
Vorticity field of the single stabilized vortex (Fig. 3(a), left), numerically computed at R = 5.2.
- Download
- 216.79 KB
Movie S12.
Vorticity field of the reversing two-vortex state (Fig. 3(a), the second from the left), numerically computed at R = 5.6.
- Download
- 1.91 MB
Movie S13.
Vorticity field of the pulsating four-vortex state (Fig. 3(a), the third from the left), numerically computed at R = 6.4.
- Download
- 1.77 MB
Movie S14.
Vorticity field of the turbulent state (Fig. 3(a), right), numerically computed at R = 7.6.
- Download
- 3.12 MB
Movie S15.
Vorticity field of the reversing two-vortex state(Fig. 3(b)), numerically computed at R = 5.35.
- Download
- 1.61 MB
Movie S16.
Azimuthal Fourier decomposition of the vorticity field of the reversing two-vortex state at R = 5.6 (Fig. 3(c)). The original vorticity field and the modes n = 0, 1, 2 are shown.
- Download
- 7.42 MB
Movie S17.
Vorticity field of the pulsating four-vortex state (Fig. 3(d)), numerically computed at R = 6.2.
- Download
- 1.35 MB
Movie S18.
Vorticity field and its azimuthal Fourier decomposition of the analytically calculated reversing two-vortex state at R = 5.9 (Fig. 4(e)). The definitions of the modes are the same as in Supplementary Movie 12 and Fig. 3(c).
- Download
- 15.07 MB
Movie S19.
Vorticity field and its azimuthal Fourier decomposition of the analytically calculated pulsating four-vortex state at R = 7.0 (Fig. 4(f)). The definitions of the modes are the same as in Supplementary Movie 12 and Fig. 3(c).
- Download
- 2.28 MB
References
1
D. Nishiguchi, K. H. Nagai, H. Chaté, M. Sano, Long-range nematic order and anomalous fluctuations in suspensions of swimming filamentous bacteria. Phys. Rev. E 95, 020601(R) (2017).
2
H. Chaté, Dry aligning dilute active matter. Annu. Rev. Condens. Matter Phys. 11, 189–212 (2020).
3
J. Iwasawa, D. Nishiguchi, M. Sano, Algebraic correlations and anomalous fluctuations in ordered flocks of Janus particles fueled by an AC electric field. Phys. Rev. Res. 3, 043104 (2021).
4
D. Nishiguchi, Deciphering long-range order in active matter: Insights from swimming bacteria in quasi-2D and electrokinetic janus particles. J. Phys. Soc. Jpn. 92, 121007 (2023).
5
I. S. Aranson, Bacterial active matter. Rep. Prog. Phys. 85, 076601 (2022).
6
M. Fruchart, C. Scheibner, V. Vitelli, Odd viscosity and odd elasticity. Annu. Rev. Condens. Matter Phys. 14, 471–510 (2023).
7
A. Sokolov, M. M. Apodaca, B. A. Grzybowski, I. S. Aranson, Swimming bacteria power microscopic gears. Proc. Natl. Acad. Sci. U.S.A. 107, 969–974 (2010).
8
C. O. Reichhardt, C. Reichhardt, Ratchet effects in active matter systems. Annu. Rev. Condens. Matter Phys. 8, 51–75 (2017).
9
D. Nishiguchi, I. S. Aranson, A. Snezhko, A. Sokolov, Engineering bacterial vortex lattice via direct laser lithography. Nat. Commun. 9, 4486 (2018).
10
H. Reinken et al., Organizing bacterial vortex lattices by periodic obstacle arrays. Commun. Phys. 3, 76 (2020).
11
A. Sokolov, I. S. Aranson, Reduction of viscosity in suspension of swimming bacteria. Phys. Rev. Lett. 103, 148101 (2009).
12
V. A. Martinez et al., A combined rheometry and imaging study of viscosity reduction in bacterial suspensions. Proc. Natl. Acad. Sci. U.S.A. 117, 2326–2331 (2020).
13
A. Sokolov, I. S. Aranson, J. O. Kessler, R. E. Goldstein, Concentration dependence of the collective dynamics of swimming bacteria. Phys. Rev. Lett. 98, 158102 (2007).
14
A. Sokolov, I. S. Aranson, Physical properties of collective motion in suspensions of bacteria. Phys. Rev. Lett. 109, 248109 (2012).
15
H. H. Wensink et al., Meso-scale turbulence in living fluids. Proc. Natl. Acad. Sci. U.S.A. 109, 14308–14313 (2012).
16
J. Dunkel et al., Fluid dynamics of bacterial turbulence. Phys. Rev. Lett. 110, 228102 (2013).
17
J. Dunkel, S. Heidenreich, M. Bär, R. E. Goldstein, Minimal continuum theories of structure formation in dense active fluids. New J. Phys. 15, 045016 (2013).
18
K. Qi, E. Westphal, G. Gompper, R. G. Winkler, Emergence of active turbulence in microswimmer suspensions due to active hydrodynamic stress and volume exclusion. Commun. Phys. 5, 49 (2022).
19
H. Wioland, F. G. Woodhouse, J. Dunkel, J. O. Kessler, R. E. Goldstein, Confinement stabilizes a bacterial suspension into a spiral vortex. Phys. Rev. Lett. 110, 268102 (2013).
20
E. Lushi, H. Wioland, R. E. Goldstein, Fluid flows created by swimming bacteria drive self-organization in confined suspensions. Proc. Natl. Acad. Sci. U.S.A. 111, 9733–9738 (2014).
21
H. Wioland, E. Lushi, R. E. Goldstein, Directed collective motion of bacteria under channel confinement. New J. Phys. 18, 075002 (2016).
22
H. Wioland, F. G. Woodhouse, J. Dunkel, R. E. Goldstein, Ferromagnetic and antiferromagnetic order in bacterial vortex lattices. Nat. Phys. 12, 341–345 (2016).
23
K. Beppu et al., Geometry-driven collective ordering of bacterial vortices. Soft Matter 13, 5038–5043 (2017).
24
K. Beppu et al., Edge current and pairing order transition in chiral bacterial vortices. Proc. Natl. Acad. Sci. U.S.A. 118, e2107461118 (2021).
25
K. T. Wu et al., Transition from turbulent to coherent flows in confined three-dimensional active fluids. Science 355, 1284 (2017).
26
A. Opathalage et al., Self-organized dynamics and the transition to turbulence of confined active nematics. Proc. Natl. Acad. Sci. U.S.A. 116, 4788–4797 (2019).
27
J. Hardoüin et al., Reconfigurable flows and defect landscape of confined active nematics. Commun. Phys. 2, 121 (2019).
28
J. Hardoüin et al., Active boundary layers in confined active nematics. Nat. Commun. 13, 6675 (2022).
29
C. D. Schimming, C. Reichhardt, C. Reichhardt, Friction-mediated phase transition in confined active nematics. Phys. Rev. E 108, L012602 (2023).
30
M. Ravnik, J. M. Yeomans, Confined active nematic flow in cylindrical capillaries. Phys. Rev. Lett. 110, 026001 (2013).
31
F. G. Woodhouse, R. E. Goldstein, Spontaneous circulation of confined active suspensions. Phys. Rev. Lett. 109, 168105 (2012).
32
F. G. Woodhouse, R. E. Goldstein, Cytoplasmic streaming in plant cells emerges naturally by microfilament self-organization. Proc. Natl. Acad. Sci. U.S.A. 110, 14132–14137 (2013).
33
S. Liu, S. Shankar, M. C. Marchetti, Y. Wu, Viscoelastic control of spatiotemporal order in bacterial active matter. Nature 590, 80–84 (2021).
34
A. E. Hamby, D. K. Vig, S. Gaines, C. W. Wolgemuth, Swimming bacteria power microspin cycles. Sci. Adv. 4, eaau0125 (2018).
35
H. Reinken, S. Heidenreich, M. Bär, S. H. L. Klapp, Pattern selection and the route to turbulence in incompressible polar active fluids. New J. Phys. 26, 063026 (2024).
36
S. H. Strogatz, Ed., “Nonlinear dynamics and chaos: With applications to physics, biology, chemistry, and engineering (studies in nonlinearity)” in Nonlinear Dynamics and Chaos: With Applications to Physics, Biology, Chemistry, and Engineering (Studies in Nonlinearity) (CRC Press, 2001), pp. 262–264.
37
S. L. Brunton, J. L. Proctor, J. N. Kutz, Discovering governing equations from data by sparse identification of nonlinear dynamical systems. Proc. Natl. Acad. Sci. U.S.A. 113, 3932–3937 (2016).
38
R. Supekar et al., Learning hydrodynamic equations for active matter from particle simulations and experiments. Proc. Natl. Acad. Sci. U.S.A. 120, e2206994120 (2023).
39
I. A. Beresnev, V. N. Nikolaevskiy, A model for nonlinear seismic waves in a medium with instability. Phys. D Nonlinear Phenom. 66, 1–6 (1993).
40
M. I. Tribelsky, K. Tsuboi, New scenario for transition to turbulence? Phys. Rev. Lett. 76, 1631 (1996).
41
J. Słomka, J. Dunkel, Spontaneous mirror-symmetry breaking induces inverse energy cascade in 3D active fluids. Proc. Natl. Acad. Sci. U.S.A. 114, 2119–2124 (2017).
42
S. Shiratani, K. A. Takeuchi, D. Nishiguchi, Route to turbulence via oscillatory states in polar active fluid under confinement. arXiv [Preprint] (2023). http://arxiv.org/abs/2304.03306 (Accessed 28 July 2024).
43
H. Matsukiyo, J.-I. Fukuda, Oscillating edge current in polar active fluid. Phys. Rev. E 109, 054604 (2024).
44
W. Liao, I. S. Aranson, Viscoelasticity enhances collective motion of bacteria. PNAS Nexus 2, pgad291 (2023).
45
S. Zhou, A. Sokolov, O. D. Lavrentovich, I. S. Aranson, Living liquid crystals. Proc. Natl. Acad. Sci. U.S.A. 111, 1265 (2014).
46
D. Nishiguchi, M. Sano, Mesoscopic turbulence and local order in Janus particles self-propelling under an ac electric field. Phys. Rev. E 92, 052309 (2015).
47
W. R. DiLuzio et al., Escherichia coli swim on the right-hand side. Nature 435, 1271–1274 (2005).
48
Marcos, H. C. Fu, T. R. Powers, R. Stocker, Bacterial rheotaxis. Proc. Natl. Acad. Sci. U.S.A. 109, 4780–4785 (2012).
49
G. Jing, A. Zöttl, É. Clément, A. Lindner, Chirality-induced bacterial rheotaxis in bulk shear flows. Sci. Adv. 6, eabb2012 (2020).
50
D. Grober et al., Unconventional colloidal aggregation in chiral bacterial baths. Nat. Phys. 19, 1680–1688 (2023).
51
J. Li, B. Esteban-Fernández de Ávila, W. Gao, L. Zhang, J. Wang, Micro/nanorobots for biomedicine: Delivery, surgery, sensing, and detoxification. Sci. Robot. 2, eaam6431 (2017).
52
J. Li, M. Pumera, 3d printing of functional microrobots. Chem. Soc. Rev. 50, 2794–2838 (2021).
53
D. Nishiguchi, S. Shiratani, I. S. Aranson, K. A. Takeuchi, Codes and data used in “vortex reversal is a precursor of confined bacterial turbulence”. Zenodo. https://doi.org/10.5281/zenodo.14553935. Accessed 25 December 2024.
Information & Authors
Information
Published in
Classifications
Copyright
Copyright © 2025 the Author(s). Published by PNAS. This open access article is distributed under Creative Commons Attribution License 4.0 (CC BY).
Data, Materials, and Software Availability
All the relevant experimental and numerical data, the MATLAB codes for analyzing the experimental data, Python scripts for the numerical simulations of the TTSHE, and the Mathematica code for analytical theory are deposited on Zenodo (53). Although only the initial parts of the experimental movies were deposited due to the size limit of the repository, the complete image sequences can be obtained by contacting the authors at [email protected]. The Mathematica code for analytical theory is also included in SI Appendix.
Submission history
Received: July 28, 2024
Accepted: January 31, 2025
Published online: March 14, 2025
Published in issue: March 18, 2025
Change history
April 10, 2025: The numbering of Supporting Movies S1–S19 have been updated and Movie S2 has been removed to coincide with a formal Correction.
Keywords
Acknowledgments
This work is supported in part by Japan Society for the Promotion of Science (JSPS) KAKENHI Grant Numbers JP19H05800, JP20K14426, JP23K25838, JP24K00593, JP24KJ0890, Japan Science and Technology Agency (JST) PRESTO Grant Number JPMJPR21O8, JST FOREST Grant Number JPMJFR2364, Japan, and the JSPS Core-to-Core Program “Advanced core-to-core network for the physics of self-organizing active matter (JPJSCCA20230002).” The research of I.S.A. was supported by the NSF award PHY-2140010. I.S.A. also acknowledges the visiting professorship for the Global Science Graduate Course–B, at The University of Tokyo, where the main part of this work was done. S.S. acknowledges support from The Forefront Physics and Mathematics Program to Drive Transformation, World-leading Innovative Graduate Study Program, the University of Tokyo. The computation in this paper was done partly by the nekoya/ai cluster at the Institute for Physics of Intelligence, The University of Tokyo, and the facilities of the Supercomputer Center, the Institute for Solid State Physics, The University of Tokyo.
Author contributions
D.N. conceived the project; D.N. initiated the project; D.N. performed the experiments; D.N. analyzed the experimental data; S.S. performed the numerical simulations; S.S. analyzed the numerical data; I.S.A. developed the analytical theory; D.N., S.S., K.A.T., and I.S.A. discussed the results; D.N., S.S., K.A.T., and I.S.A. interpreted the results; and D.N., S.S., K.A.T., and I.S.A. wrote the paper.
Competing interests
The authors declare no competing interest.
Notes
This article is a PNAS Direct Submission.
Authors
Metrics & Citations
Metrics
Altmetrics
Citations
Cite this article
Vortex reversal is a precursor of confined bacterial turbulence, Proc. Natl. Acad. Sci. U.S.A.
122 (11) e2414446122,
https://doi.org/10.1073/pnas.2414446122
(2025).
Copied!
Copying failed.
Export the article citation data by selecting a format from the list below and clicking Export.
Cited by
Loading...
View Options
View options
PDF format
Download this article as a PDF file
DOWNLOAD PDFLogin options
Check if you have access through your login credentials or your institution to get full access on this article.
Personal login Institutional LoginRecommend to a librarian
Recommend PNAS to a LibrarianPurchase options
Purchase this article to access the full text.