Collective buoyancy-driven dynamics in swarming enzymatic nanomotors

Nature

Collective buoyancy-driven dynamics in swarming enzymatic nanomotors"


Play all audios:

Loading...

ABSTRACT Enzymatic nanomotors harvest kinetic energy through the catalysis of chemical fuels. When a drop containing nanomotors is placed in a fuel-rich environment, they assemble into


ordered groups and exhibit intriguing collective behaviour akin to the bioconvection of aerobic microorganismal suspensions. This collective behaviour presents numerous advantages compared


to individual nanomotors, including expanded coverage and prolonged propulsion duration. However, the physical mechanisms underlying the collective motion have yet to be fully elucidated.


Our study investigates the formation of enzymatic swarms using experimental analysis and computational modelling. We show that the directional movement of enzymatic nanomotor swarms is due


to their solutal buoyancy. We investigate various factors that impact the movement of nanomotor swarms, such as particle concentration, fuel concentration, fuel viscosity, and vertical


confinement. We examine the effects of these factors on swarm self-organization to gain a deeper understanding. In addition, the urease catalysis reaction produces ammonia and carbon


dioxide, accelerating the directional movement of active swarms in urea compared with passive ones in the same conditions. The numerical analysis agrees with the experimental findings. Our


findings are crucial for the potential biomedical applications of enzymatic nanomotor swarms, ranging from enhanced diffusion in bio-fluids and targeted delivery to cancer therapy. SIMILAR


CONTENT BEING VIEWED BY OTHERS LOCOMOTION AND DISAGGREGATION CONTROL OF PARAMAGNETIC NANOCLUSTERS USING WIRELESS ELECTROMAGNETIC FIELDS FOR ENHANCED TARGETED DRUG DELIVERY Article Open


access 23 July 2021 CHEMICALLY-POWERED SWIMMING AND DIFFUSION IN THE MICROSCOPIC WORLD Article 27 May 2021 ION-EXCHANGE ENABLED SYNTHETIC SWARM Article 11 January 2021 INTRODUCTION


Collective behaviour is widespread in nature. While individual units of a group obey simple rules, they present complex and intriguing collective behaviour when assembling into highly


ordered structures1. Living organisms use distributed or swarm intelligence to accomplish sophisticated tasks to survive. Examples range from collective cell migration2, honeybees adapting


to repeated shaking to maintain mechanical stability of the swarm3, to emperor penguins packing in a huddle in a highly coordinated manner to survive cold winter4. Multiple synthetic


swarming systems have been developed with inspiration from nature including: (1) applying one or multiple external forces, such as magnetic fields5,6,7,8, light9,10, ultrasound11,12,


electric fields13,14, (2) utilizing chemicals as signals15,16,17, (3) combining biological microswimmers, such as sperm cells and algae, into artificial moieties as a hybrid


integration18,19,20, (4) exploiting DNA base-pair interactions21,22. These well-designed swarms show many advantages compared to single-unit functionalities, like enhanced coverage and fluid


mixing, intelligent multitasking, collective chemotaxis and perception, and environmental adaptation. Micro/nanomotors (MNMs) are synthetic active devices achieving self-propulsion through


converting various types of energy into mechanical motion23,24. Earlier works on enzyme-powered MNMs have demonstrated the motion of single particles25,26,27 and small clusters28, as well as


proof-of-concept studies in drug delivery29,30,31,32,33 and sensing34,35. Nonetheless, recent reports have shifted focus to the collective motion of these particles. Recently, Hortelao et


al.36 reported the emergent swarming behaviour of enzymatic nanomotors. The urease-powered nanomotors show collective migration in urea, demonstrating the ability to swim across complex


paths compared to the inactive nanomotors. Furthermore, the active collective dynamics, combined with advanced imaging technologies, position them as promising tools in the field of


biomedicine. For example, swarms of radio-labelled nanobots have shown an eightfold increase in tumour penetration and approximately a 90% reduction in tumour size during radionuclide


therapy37. Swarms of catalase-powered nanobots overcome and disrupt mucus layer, resulting in a 60-fold increase in mucus barrier penetration, through in vitro and ex vivo validation38.


Hyaluronidase and urease nanomotor swarms work synergistically for enhanced diffusion in viscous media, such as synovial fluid, paving the way for treating joint injuries39. Similarly,


collagenase-powered MNMs40,41 and urease-powered iron oxide nanomotor swarms42 were exploited to disrupt collagen fibres, serving as a model of the extracellular medium. This disruption


facilitates cell spheroids penetration and enhances the delivery efficiency of a second swarm of nanomotors by 10-fold. Although enzyme-powered MNMs have primarily demonstrated their


potential in biomedical applications, the mechanisms underlying the emergent collective behaviour remain to be clarified. Inspired by nature, the intriguing collective phenomenon bears a


resemblance to bioconvection. Bioconvection is a self-organized and self-sustained vortex motion that arises naturally in suspensions of microorganisms43. It visually resembles the


Rayleigh–Bénard convection in fluid heated from below44. The bioconvection emerges due to the unstable density gradients resulting from the accumulation of buoyant microorganisms45. Each


microorganism plays a pivotal role in driving accumulation and fluid flow. Certain gravitactic algae or aerotactic bacteria exhibit upward swimming. In the presence of an upper surface, they


form a thin boundary layer of microorganism-rich heavier fluid, which becomes unstable, leading to the formation of falling plumes46. In synthetic MNMs, buoyancy-driven convection has been


employed for directional motility and cargo delivery. One approach utilizes incident light to generate convective flow through the photothermal effect. This convective flow can drive TiO2


micromotors to aggregate and form clusters47,48, or enable magnetic colloidal collectives to drift using fluidic currents49. Another approach involves enzymes fixed on a surface, which


catalyze fuels, inducing density variations between the reactants and products of chemical reactions. For instance, urease-attached macroscale sheets exhibit clockwise or counterclockwise


rotation via a solutal buoyancy mechanism arising from enzyme pumps50. Additionally, urease attached to a lipid membrane induces fluid flow and the directional transport of tracer


particles51. These studies underscore the significant implications of convective flow on the collective movement of micro/nanoparticles. Here, the buoyancy-driven convective dynamics of


collective enzymatic nanomotors arises from the density difference between the product-rich particulate and the denser environment. We attach enzymes onto the surface of nanoparticles, which


move dynamically with the fluid flow. We describe the emergent collective behaviour of enzymatic nanomotor in three dimensions (3D) to unveil the underlying mechanisms. We model a swarm as


light particulates immersed in a denser fluid environment. The particulate swarm moves upward, creating a convective flow in a closed fuel-filled space. Control factors for directional and


collective mobility of enzymatic nanomotors, such as particle and fuel concentration and media viscosity, are investigated. In 3D space, convective flows develop complex patterns such as


vortices. To further validate our findings, we vertically confine the system to two dimensions, where convective flow is constrained to a plane, inhibiting the possible flow configurations.


The convective dynamics resemble bioconvection and provide insights into the mechanisms of collective behaviour observed in enzymatic nanomotors. RESULTS AND DISCUSSION In our study, we view


the collective behaviour of enzymatic nanomotors from the side or from the top. The enzymatic nanomotors are based on mesoporous silica nanoparticles with urease attached (UrNMs). Detailed


characterization can be found in the Supplementary Materials (Fig. S1) and the Methods section. The molecularly unbalanced distribution of enzymes generates net motion for single


nanomotors52,53 in urea, Fig. S2. From the side, upon introducing a drop of particulate in a fuel-filled chamber, the drop shows upward motion against gravity, generating a convective flow


within the closed space, Fig. 1a. During an ascending stage, there are two counter-rotating vortices within a droplet. A characteristic hydrodynamic flow pattern is displayed in Fig. S3. As


the particulate reaches the upper boundary, it spreads to balance the mean upward force, forming a layer of unstable particle-rich fluid. The layer then sinks in the form of falling plumes.


Trajectory tracking of UrNMs in a 22 × 8 × 1.6 mm (length × height × width) chamber shows the spreading, sinking, and swirling stages of the convective dynamics, Fig. 1b–g. These


trajectories on the left and right are not perfectly symmetric due to experimental limitations. The upward movement of a nanomotor swarm is due to buoyancy arising from the density


difference between the reaction product-rich particulate and the media with fuel. We state that individual nanomotors perform urease catalysis reaction and generate ammonia and carbon


dioxide, making the particulate lighter. Since the temperature change during chemical reactions is not obvious (Fig. S4), we rule out heat effect on the upward movement. In addition, the


analysis of enzymatic activity and the upward velocity of UrNMs in urea at physiological temperature (37 °C) shows no significant difference compared to room temperature (25 °C), Fig. S5a–c.


We performed computational modelling based on two-fluid hydrodynamics and compared the computational results to the experiments; a good qualitative agreement is obtained, see Fig. 1h, i. To


verify the universality of our mechanisms, we synthesized catalase-powered nanomotors (CatNMs) and observed the convective dynamics of these enzymatic nanomotors in hydrogen peroxide (H2O2)


(Fig. S6 and video S1). Notably, the instant chemical reaction of CatNMs in H2O2 results in a burst of oxygen bubbles, which drives the CatNMs to move upward against gravity within one


second. We expect that this buoyancy-driven mechanism will also apply to other asymmetric motors, such as Janus motors. Previous studies in our group have verified the difference between the


Janus structure and the patchy-like structure. The motion of a single out-of-equilibrium particle arises from the asymmetric distribution of ions, which generates an ionic gradient54. A


theoretical study has shown that Janus particles exhibit higher velocities compared to patchy-like motors55. Additionally, it has been reported that micron-sized hollow urease motors present


a 3D motion at a single particle level56, thus it is expected that large populations of these particles will also show collective motion. However, if the fabrication of Janus structures


involves heavier materials like platinum or gold, sedimentation may neglect buoyancy57,58,59, impeding the upward collective movement. That suggests that the buoyancy-driven mechanism could


be universal for various types of motors across different length scales, provided that the gravitational effects would not suppress the buoyancy-driven motion. CONTROLLING COLLECTIVE


BEHAVIOUR OF URNMS Collective behaviour can be viewed from the side. We studied the influence of three main control factors, UrNM concentration, urea concentration, and viscosity mediated by


hyaluronic acid (HA) concentration, on the collective behaviour. As illustrated in Fig. 2a, there are three stages of the collective behaviour of enzymatic nanomotors, i.e., ascending (1, 2


and 3), spreading (2 and 3), and sinking (3). When a swarm of nanomotors seeded in the bottom of the chamber that is filled with fuel, they show directional mobility against gravity. Under


different conditions, they display various forms of collective behaviour and velocity. Figure 2b shows the _z_-component of the particulate centre of mass as a function of time within 32 s.


The velocity difference can be deducted from Fig. S7 showing that for higher UrNM concentrations, particulates reach a lower _z_ position in 5 s. As the UrNM concentration increases to 20 


mg/mL, the majority of the nanomotor swarms cannot get to the upper boundary due to gravity (Fig. 2b, c and video S2). The velocity field was analyzed by front-tracking the particulate based


on custom Python code. As expected, compared with passive nanoparticles (MSNPs), active nanomotors show enhanced upward speeds, Fig. S8 and video S3. We assume that enzymatic catalysis of


urea produces microbubbles60 and the product, ammonia, makes this particulate less dense. Although the product quantity may be larger with higher UrNM concentration, the density of


particulate increases as well when we increase the concentration of nanoparticles. We state that there should be a competition between the two opposite conditions, after which the effect of


increased particulate density takes the lead, and the upward particulate velocity decreases with the increased UrNM concentration. Buoyancy, the main driving force, is strongly influenced by


fuel concentration. Figure 2d, e show that the upward speeds increase with the fuel concentration. One can clearly observe the upward motion of particulates at concentrations of 150 mM urea


and above. However, in the presence of 100 mM urea concentration, particulate almost stays at the seeding point, and there is no difference between the upward motion of active swarms and


passive particulates. We argue that this is because in low urea concentration, density difference resulting in a buoyancy force is not sufficient to lift the particulate. To distinguish the


behaviours between active nanomotors and passive nanoparticles, we introduce phenol red, a pH indicator, into the urea solution. The chemical reactions that occur during the collective


movement of active nanomotors lead to a pH change in the surrounding solution, resulting in a colour shift from light yellow to pink, video S4 and Fig. S9. This colour change indicates the


location of the active UrNMs. For comparison, passive nanoparticles were tested in the same condition (10 mg/mL, 200 mM urea). They expand along the bottom plane at time 0, and the pH of


their surrounding solution remains unchanged. We added hyaluronic acid into the fuel to change the media viscosity observing that the upward speeds of particulate decreases with the increase


in concentration of hyaluronic acid, Fig. 2f, g. As it was shown above, active swarms show enhanced speed compared to passive particulates in viscous media. When the concentration of


hyaluronic acid increases to 3 mg/mL, both active swarms and passive particulates remain at the seeding point because higher viscosities inhibit fluid convection. We conducted particulate


velocity analysis at elevated heights in the middle of the chamber. In Fig. 2h, active particles move slightly faster in the middle of their paths and decrease their speeds when approaching


the upper boundary in different groups, while passive particles keep decreasing their speeds (Fig. S10). For instance, a particulate of 5 mg/mL UrNMs moves upward at 1.74 ± 0.09 mm/s at 4 mm


height, 1.93 ± 0.14 mm/s at 5 mm height, and 1.70 ± 0.08 mm/s at 7 mm height, while a particulate of the same concentration of passive nanoparticles moves at 0.96 ± 0.03 mm/s at 4 mm


height, 0.68 ± 0.02 mm/s at 5 mm height, and 0.47 ± 0.02 mm/s at 7 mm height. The acceleration process of active particulate could be due to the density changes caused by chemical reaction


products. Additionally, during the spreading stage, active nanomotors form a thin boundary layer of particle-rich fluid that continues to spread until it meets the side boundary. In


contrast, passive particles form a less stable boundary layer, leading to the formation of larger falling plumes earlier, causing them to sink before reaching the side boundary (see videos 


S2 and S3). We assume this occurs because the products of chemical reactions make the active particles less dense, and the faster upward movement of active nanomotors creates a more dynamic


environment with increased fluid flow, making it less likely to form large falling plumes. PRODUCTS OF URNMS CATALYSIS REACTION ACCELERATE THE DIRECTIONAL MOVEMENT Urease catalyzes the


decomposition of urea into ammonia (NH3) and carbon dioxide (CO2). On the one hand, NH3 is highly soluble in water due to the formation of hydrogen bonds with water molecules. This


interaction results in a smaller density of the solution61,62. On the other hand, the released NH3 dissolves in water, resulting in an alkaline solution (Fig. 3a) and promoting CO2 to


dissolve. Under proper fuel concentration, the formation of NH3 and CO2 microbubbles can be observed60. However, in acidic buffers, such as acetate buffer (pH = 4.6, Fig. 3b), CO2 may exist


because the abundant hydrogen ions inhibit the dissolution of CO2 and the ionization of carbonic acid. The main reaction rate constants for CO2 and NH3 in phosphate buffer saline (PBS)


buffer and acetate buffer are presented in table S1. In PBS buffer, the rate constant for CO2 dissolution63 in basic solutions (k1 = 1.21 × 104 M−1s−1) is much higher than the reverse rate


constant (k−1 = 4.0 × 10−4 s−1). In acetate buffer, ammonia dissolves in acidic solutions. The rate constant for CO2 dissolution in water (k4 = 0.037 s−1) is much smaller than the reverse


constant of HCO3- combining with H+ (k−4 = 1.24 × 105 M−1 s−1). We conducted experiments to verify the existence of NH3 and CO2. In Fig. 3c, cover papers were pre-dipped in phenol red


solutions. Upon adding urease or UrNMs into urea solution, NH3 is produced and volatilizes until it dissolves in the cover paper that contains phenol red, the colour change of which from


light yellow to pink indicates the presence of NH3. The production of CO2 can be observed in acetate buffer, which maintains an acidic environment during the urease catalysis reaction, Fig. 


3b. CO2 bubbles produced by UrNMs reacting with urea dissolving in acetate buffer can be observed on the wall of a cuvette (video S5). We filled cuvettes with 300 mM urea solutions that were


dissolved either in PBS buffer or in acetate buffer. Then UrNMs or urease solutions were added to the cuvettes, respectively. Video S5 shows clearly the convective flow from the turbidity


while for smaller urease molecules, the solutions remain transparent. In addition, the produced NH3 and CO2 in acetate buffer can be directly detected by a gas sensor, an optoelectronic


analysis equipment that is able to accurately detect low-concentration gases at the ppm level, as shown in Fig. 3d, e. The enzymatic activity of UrNMs in urea solutions in both PBS buffer


(Fig. 3f, S11) and acetate buffer (Fig. S12) was examined. In PBS buffer, the specific enzymatic activity of UrNMs increases from 4.08 ± 0.02 U/mg in 50 mM urea solutions to 4.82 ± 0.41 U/mg


in 300 mM urea solutions. In acetate buffer, the specific enzymatic activity of UrNMs is slightly weaker, with 1.94 ± 0.05 U/mg in 50 mM urea solutions and 3.36 ± 0.45 U/mg in 300 mM urea


solutions. This is because the known optimum pH for urease catalytic activity is around 7 ~ 864. We also examined the enzymatic activity of UrNMs in urine. Figure S13 shows that the


enzymatic activity of urease decreases from 5.25 ± 0.06 U/mg to 3.56 ± 0.15 U/mg in simulated urine65, and to 3.57 ± 0.20 U/mg in real urine. Although there is a significant decrease in


enzymatic activity in urine for 30 min, the activity of UrNMs is still relatively high, verifying the potential of UrNMs for in vivo applications. The above results indicate that the urease


catalysis reaction produces dissolved NH3 and CO2 in PBS buffer and dissolved NH3 and CO2 gas in acetate buffer. Therefore, upon quantifying the upward velocity of the UrNMs particulate, the


results indicate a faster upward movement, from 1.01 ± 0.04 mm/s in PBS buffer to 1.14 ± 0.04 mm/s in acetate buffer, as shown in Fig. S14. These findings validate our assumption that the


chemical products result in a higher density difference between the particulate and the media with fuel, leading to an accelerated movement. VERTICAL CONFINEMENT SHAPES COLLECTIVE BEHAVIOUR


Since buoyancy is the primary force that drives the self-organization of active particulates, we studied the influence of vertical confinement on their collective behaviour. As shown in Fig.


 4, microfluidic chips with three different heights (1.6 mm, 0.5 mm, and 0.25 mm) were designed and filled with urea in the vertically confined chamber. Then active UrNMs were introduced and


entered the chamber from the side by capillary force. In Fig. 4a and video S6, these active UrNMs swarms exhibit collective movement in the chamber of 1.6 mm height. The density maps,


observed from the top, show that the swarms aggregate, coarsen, and change their patterns over time. Particle image velocimetry (PIV) also confirms that the fluid flow is initially faster


when the nanomotors are injected into the chamber. Fig. S15–S17 show the PIV results at 25 s time intervals in confinement with different heights. After 50 s, nanomotors keep moving and


collective behaviour is still transient. After 100 s, the fluid flow keeps a relatively high speed, 1.5 μm/s on average. However, the fluid flow direction remains the same according to the


arrows. As a comparison, without fuel UrNMs sink to the bottom in a confined chamber and expand along the bottom plane, Fig. S18. The convective flow is also weaker than that caused by UrNMs


with fuel, Fig. S19–21. When the vertical confinement is changed to 0.5 mm, the movement of active UrNMs becomes localized. In Fig. 4b, the density map shows that the pattern of UrNMs only


slightly changes over time. The PIV reveals that fluid flow velocity decreases compared to larger height values. After 50 s, the swarms barely move. When the height is further reduced to


0.25 mm, the swarms’ movement is hindered, as displayed by the unchanged shape of swarms over time and the decreased velocity of fluid flow in PIV, Fig. 4c. Active UrNMs in PBS solutions


also show decreased velocity when the chamber height decreases (Fig. S21). However, compared with the active UrNMs in fuel, there are no significant differences. We also analyzed the swarm


dynamics by pixel intensity distribution. A time-lapse sequence of snapshots at 12 s time intervals from video recordings is selected. As shown in Fig. S22, in a 1.6 mm-high chamber, the


pixel intensity of active UrNMs in fuel is broadly distributed in the region of interest (ROI) in the initial 60 seconds, and gradually changes to narrowly distributed in 2 min. However, for


the 0.5 mm-high chamber and the 0.25 mm-high chamber, pixel intensities are monodispersed in the ROI within the time durations. As a comparison, the pixel intensities of active UrNMs in PBS


solutions are highly monodispersed in the three different chambers, Fig. S23. These results indicate that the vertical confinement controls the swarms by affecting fluid convective flows


and provide insight into the buoyancy-driven collective behaviour of nanomotors. COMPUTATIONAL MODELLING SHOWS SIMILARITY WITH EXPERIMENTS Our starting point is two-fluid hydrodynamics66.


One fluid is a solvent with the kinematic viscosity _η_, flow velocity V, solvent pressure _p_, and solvent density _ρ_0. Second fluid is the particulate with the volume density _ρ_,


coarse-grained particulate velocity U, and pressure _P_ = _qρ_, and the factor _q_ depends on the temperature (as for gases). We describe the dynamics by the simplified Navier-Stokes Eq.


(1), coupled to the reaction-advection equation for the concentration of chemical fuel _c_, Eq. (2), and a mass transport equation for the particulate density, Eq. (3): $${\rho


}_{0}\left({\partial }_{t}{{{\bf{v}}}}+{{{\bf{v}}}}\nabla {{{\bf{v}}}}\right)=\eta {\nabla }^{2}{{{\bf{v}}}}-\nabla p-{{{{\bf{z}}}}}_{0}\rho \left(g\alpha -\epsilon c\right)$$ (1)


$${\partial }_{t}c+\nabla \cdot \left({{{\bf{v}}}}c\right)={D}_{c}{\nabla }^{2}c-\gamma \rho c$$ (2) $${\partial }_{t}\rho+\nabla \cdot \left({{{\bf{v}}}}\rho \right)=\left(q{\nabla


}^{2}\rho+\alpha g{\partial }_{z}\rho \right)/{\kappa }_{1}$$ (3) where \({{{{\boldsymbol{z}}}}}_{0}\)_ρєc_ is the volume buoyancy force due to gas generation, \({{{{\boldsymbol{z}}}}}_{0}\)


is the unit vector in the _z_-direction, the gas is produced due to the reaction between fuel _c_ and particulate _ρ_ with the reaction rate _γ_. Other parameters: fuel diffusion _D__c_,


gravity acceleration _g_, relative particulate/solvent density contrast _α_, _є_ is the relative buoyancy coefficient that depends on the density of reaction products, and _κ_1 is the


normalized drag coefficient. The details of model derivation are presented in Supplementary Note 1. Equations (1)–(3) were solved by the finite difference method using Matlab. We considered


a two-dimensional rectangular integration domain (corresponding to the size view) with periodic boundary conditions in the _x_-direction and non-slip conditions in the _z_-direction. The


primary difference with models of enzyme-generated solutal buoyancy mechanisms considered in ref. 67. is that the enzyme distribution is not fixed but dynamically updated by the


reaction-generated flow. When buoyancy is not sufficient to counterbalance the gravity of particulates, like in the cases of high concentration of particles and low concentration of fuel,


the particulate is not able to rise to the top plane and sink to the bottom after seeding, Fig. 5a, left panel. On the contrary, in the cases of low concentration of particles and high


concentration of fuel, particulates rise and spread along the top plane, then descend, experiencing a similar process as in the experiment, Fig. 5a, right panel, and video S7. In


simulations, the volume density _ρ_ changes from 1 to 4, chemical fuel _c_ ranges from 0.6-1.2, and kinematic viscosity _η_ varies from 0.1-1.0 to simulate different concentrations of


particles, fuel, and HA, respectively. In Fig. 5b–d, frames at dimensionless time 2.8 are chosen from computer videos for different parameters. Figure 5b shows that in the same time frame,


particulate with smaller density _ρ_ enters the sinking stage, while particulate with larger _ρ_ is still in the ascending or spreading stage, indicating that lighter particulates move


faster. This observation agrees with the experimental results and can be further verified by Fig. 5e. The mean velocity of particulate during upward movement decreases with the increase of


density _ρ_. In Fig. 5c, particulate settles to the bottom when chemical fuel concentration _c_ is low (_c_ = 0.6). Increasing the _c_ value (_c_ = 0.8) triggers particulate’s upward


movement, yet it settles before reaching the top plane. Only relatively high fuel concentrations force the particulate to go through the three stages, and its upward speed increases with the


increase of _c_ value. In Fig. 5f, the gradual increase of the mean particulate velocity with the fuel concentration from simulations agrees with that observed in the experiments. The


effect of viscosity is shown in Fig. 5d, g. Particulate in lower viscosity media enters the sinking stage earlier than for higher viscosity. Computational modelling confirms that the


increased fuel viscosity slows down the particulate motion. COMPUTATIONAL MODELLING OF THE VERTICAL CONFINEMENT EFFECTS We performed computational modelling of the effect of vertical


confinement on collective behaviour. The details are presented in Supplementary Note 2. The model is derived from Eqs. (1)–(3) by height-averaging using the approach like in ref. 68. The


corresponding two-dimensional equations in the _x-y_ plane are solved by the quasi-spectral method in the periodic square domain using Matlab. Parameter _β_ is proportional to the reaction


rate and parameter _ε_~_h_2, where _h_ is the height of the chamber. We adjust the value of these two control parameters to describe the fluid flow slowdown caused by confinement. In Fig. 


6a, numerical results show that in vertical confinement, particulate moves dynamically and form aggregates in the centre area of the cell. A similar phenomenon has been observed in


experiment, Fig. 4a. However, when the chamber’s height is reduced, the fluid flow slows, and the reaction rate decreases. As a result, particulate movement becomes more localized, and the


shape formed by a particulate remains almost unchanged within the time durations, as shown in Fig. 6b, c and video S8. Furthermore, there is no significant difference between the swarm


dynamics in two highly confined chambers because fluid convection is inhibited by vertical confinement. In conclusion, we investigated the collective behaviour of enzymatic nanomotors from


the side and from the top. We attribute their collective behaviour to buoyancy-induced convection. When introducing a drop of UrNMs, dispersed in PBS buffer, into a fuel medium (high


concentration of urea dissolved in PBS), the UrNMs exhibit directional upward movement due to buoyancy arising from the density difference between the particulate and the fuel medium. UrNMs


decompose urea and generate carbon dioxide and ammonia, with the latter dissolving in water, further reducing the particulate density and enhancing its upward movement. When reaching the


solid-air interface, UrNMs spread along the interface, form an unstable layer of front, and then sink in the form of finger-like aggregates. The process resembles natural bioconvection in


microorganismal suspensions. Particle concentration, fuel concentration, and viscosity are crucial parameters to control enzymatic collective behaviour. Specifically, increasing particle


concentration, decreasing fuel concentration, or increasing viscosity can decrease the density difference between the particulate and the fuel, impeding the initiation of upward movement and


subsequent convection. This phenomenon explains the settlement of nanoparticles to the bottom when observed under inverted microscopy. Furthermore, the movement of UrNMs in vertical


confinement also serves as a demonstration of buoyancy-induced convection. Confinement hinders fluid convection, indicating that the collective behaviour of enzymatic nanomotors requires


vertical spaces to overcome dissipation. While these control factors are essential for understanding collective behaviour, further studies are needed to investigate how to effectively guide


swarm dynamics. Possible strategies could involve combining external fields or exploiting collective chemotaxis behaviour. We performed computational modelling based on the buoyancy-driven


convection mechanisms; the results align well with experimental findings. In computational modelling, particulate ascends due to buoyancy, spreads upon reaching the top, and consequently


descends because of gravity. Consistent with the experimental observations, an increase in particulate density (_ρ_), a decrease in fuel concentration (_c_), or an increase in fuel viscosity


(_η_) decreases the mean particulate velocity. Computational modelling also agrees with experimental observations for particulate moving in vertical confinement. By adjusting the parameters


_β_ and _κ_, corresponding to the reaction rate and the chamber height, respectively, the computational model predicts that vertical confinement shapes the swarms by controlling fluid


convection. The buoyancy-driven convective flow enables the collective movement of enzymatic nanomotors and promotes a more homogeneous particle distribution. In a fuel-rich environment,


collective behaviour occurs naturally due to buoyancy and chemical reactions, without requiring external forces. This buoyancy-driven dynamics can be harnessed to design future protocols for


large tissue and organ volumes, such as the bladder and joints. It allows overcoming the limitations of current cancer treatments, including sedimentation and poor dispersion in small


volumes, thereby facilitating mass transport, accumulation, penetration, and effective diffusivity of individual motors. METHODS SYNTHESIS OF MSNPS-NH2 Mesoporous silica nanoparticles


(MSNPs) serving as chassis for urease-propelled nanomotors were synthesized by the sol-gel procedure according to our previous report36. Briefly, a mixture of TEOA (35 g), Milli-Q water (20 


ml), and CTAB (570 mg) was heated to 95 °C under reflux for 30 minutes. TEOS (1.5 ml) was then added dropwise, and the reaction continued for 2 hours. The resulting MSNPs were collected by


centrifugation (2000 × _g_, 5 min) and washed with ethanol, with the process repeated three times. CTAB was removed by refluxing the MSNPs in a methanol (30 ml) and hydrochloric acid (1.8 


ml) mixture at 80 °C for 24 hours. Finally, the MSNPs were collected by centrifugation (2000 × _g_, 5 min), washed in ethanol (three times), and their concentration evaluated by dry


weighing. The surface of MSNPs was then modified for further functionalization. Briefly, 20 mg MSNPs in ethanol 99% (Panreac Applichem cat. no. 131086-1214) and 100 μL


3-aminopropyltriethoxysilane (APTES) 99% (Sigma-Aldrich cat. no. 440140) were mixed and placed in an end-to-end shaker at room temperature for 24 h. The resulting nanoparticles were then


collected and washed in ethanol by centrifugation (2000 × _g_, 5 min) four times to remove residual APTES. The collected MSNPs-NH2 nanoparticles were dried for further use. SYNTHESIS OF


URNMS AND CATNMS The prepared MSNPs-NH2 nanoparticles (2.5 mg) were resuspended in 1 mL PBS 1× (Thermo Fisher Scientific cat. no. 70011-036) and activated with 100 μL GA 25 wt%


(Sigma-Aldrich cat. no. G6257) in an end-to-end shaker for 2.5 hours at room temperature. The activated MSNPs-NH2 were then collected and washed four times in PBS 1× by centrifugation (2000


× _g_, 5 min), then resuspended in 1 mL PBS 1× with 3 mg urease from _Canavalia ensiformis_ (Sigma-Aldrich cat. no. U4002), or in 1 mL PBS 1× with 1 mg catalase from bovine liver


(Sigma-Aldrich cat. no. C40). The mixture reacted at room temperature in an end-to-end shaker overnight. The resulting urease-nanomotors (UrNMs)/catalase-nanomotors (CatNMs) were collected


and washed thrice in PBS 1× by centrifugation (2000 × _g_, 5 min). Keep the supernatant of centrifugation for further quantification of the enzyme linkage. Finally, resuspend the collected


nanomotors in PBS 1× (0.5 mL) and store them in the fridge at 4 °C for future use. DLS MEASUREMENTS OF URNMS Malvern Nanosizer (Zetasizer Nano ZSP) was used to measure the diffusion


coefficient of UrNMs across a range of urea concentrations (0, 50, 100, 150, and 300 mM) and the surface charge of MSNPs, MSNPs-NH2, and UrNMs. We analyzed the diffusion coefficient of UrNMs


(20 μg/mL) at each urea concentration and zeta potential values of each type of nanoparticles (20 μg/mL) with three runs per experiment. Nine measurements per type of particle were


performed to obtain statistically relevant data. URNMS CHARACTERIZATION The synthesized MSNPs were characterized by scanning electron microscope (SEM), revealing a uniform particle size


distribution centred at 450 nm (Fig. S1a, b). Amino groups were then grafted on the surface of MSNPs, facilitating further modification of urease on the MSNPs surface by linking GA molecules


between amino groups. The surface modification process was characterized by zeta potential measurements (Fig. S1c). The introduction of amino groups results in a negative surface charge of


MSNPs reversed from -38.7 ± 4.61 mV to a positive surface charge of 27.77 ± 7.9 mV. The subsequent linking of GA molecules and urease is confirmed by particle surface charge changes due to


the presence of abundant aldehyde groups and carboxyl groups, with negative surface charge reverses to -14.5 ± 9.13 mV and -9.02 ± 4.34 mV for GA molecule and urease, respectively. DLS


measurement indicates that the prepared UrNMs show an enhanced diffusion coefficient in elevated urea concentrations (Fig. S1d). OPTICAL VIDEO RECORDING AND NANOPARTICLE TRACKING The


collective behaviour of UrNMs in a vertically confined space was recorded using a Leica DMi8 microscope equipped with a high-speed cooled charge-coupled device (CCD) camera from Hamamatsu


and a 2.5× objective lens. Two pieces of cover glasses were separated by spacers (Silicone isolators from Grace Bio-Labs) with varying heights: 1.6 mm, 0.5 mm, and 0.25 mm. The confined


space was then filled either with PBS or with a 300 mM urea solution in PBS and positioned under the microscope. A drop of UrNMs or MSNPs (3 μL) was added to the liquid-filled chamber, and


videos (25 fps, 2 min) were recorded. Optical videos from the side view were recorded using either a digital camera (Thorlabs, DCC1240M-GL/Thorlabs, CS165CU) equipped with a lens (FUJINON,


HF35HA-1S) or a Leica DFC3000G camera equipped with a 10×/0.3 objective lens. A 22 × 1.6 × 8 mm (length × width × height) chamber was prepared by separating two pieces of cover glasses with


spacers. A drop of UrNMs or MSNPs (3 μL) was added to the liquid-filled chamber and videos (15 fps, 2 min) were recorded. Then these videos were analyzed using a home-designed programme in


Python27. The motion profiles of urease-powered nanomotors were analyzed in a 50 mM urea solution with varying ionic strengths, achieved by adding different concentrations of NaCl (0.05,


0.5, 5, 25, 50 mM). Videos were recorded at 25 fps for 30 seconds using an inverted optical microscope (Leica DMi8) equipped with a Hamamatsu digital camera (C11440) and a 63× water


immersion objective. For each experiment, a drop of urea solution with the specified NaCl concentration was placed on a glass slide, followed by the addition of 3 µL of UrNMs. The glass


slide was then covered with a coverslip. The acquired videos were analyzed using a custom-designed Python software. The mean square displacement (MSD) can be calculated from the extracted


trajectories by the equation MSD(Δt) = 4_D_Δt, where _D_ is the diffusion coefficient. BICINCHONINIC ACID (BCA) ASSAY The amount of urease or catalase linked onto the MSNPs surface was


quantified by BCA analysis (Thermo Fisher Scientific cat. no. 23227), table S2. Bovine serum albumin (BSA) was used as the standard for quantifying the concentration of protein


concentrations. First, a series of BSA concentrations 2000, 1500, 1000, 750, 500, 250, 125, 25, 0 μg/mL were prepared, and BSA solutions and the as-prepared supernatant of UrNMs or CatNMs


samples were added separately into a 96-multi-well plate, 25 μL for each well. Then 200 μL working reagent (light sensitive), made with 50 parts of reagent A and 1 part of reagent B, was


added to each well that has been used, either with the BSA standards or the sample. Next, shake the plate for 30 s to mix the solutions, and incubate the reaction for 30 min at 37 °C.


Afterward, the absorbance of both BSA solutions and the samples after the reaction was measured at a wavelength of 562 nm. By comparing the protein quantity remaining in the supernatant to


the initial amount of protein added and the standard concentrations of BSA, the amount of attached enzyme can be quantified. ENZYMATIC ACTIVITY MEASUREMENT Urease activity was detected


before and after being linked on the surface of MSNPs and was compared in PBS and acetate buffer. 0.025 mM Phenol red (Sigma-Aldrich cat. no. 114529) was added to different concentrations of


urea solutions (0, 50, 100, 200, and 300 mM). The solvent of these solutions could be either PBS or acetate buffer. 2 μL PBS solution or UrNMs or urease (130 μg/mL) were added in a 96-well


plate separately, followed by the addition of urea solutions. The 96-well plate was immediately placed in a multimode microplate reader (BioTek Synergy HTX). The absorbance changes of phenol


red were measured in real time at a wavelength of 560 nm for 60 min for measurements in PBS buffer and 100 min for acetate buffer. The incubation was performed at room temperature with a


minimum measurement interval of 30 seconds, and the sample was orbitally shaken at a minimal frequency. ESTIMATION OF SPECIFIC ENZYMATIC ACTIVITY The specific enzymatic activity was


determined by calculating the slope of the enzymatic activity curve. According to the Beer-Lambert law, $$A={klc}$$ where _A_ is the absorbance, _k_ is the molar attenuation coefficient of


phenol red, and _l_ is the path length of 0.5 cm, the concentration changes _c_ of phenol red per minute were computed. Subsequently, the specific enzymatic activity was derived based on the


quantity of enzyme utilized. GAS DETECTION The generated CO2 and NH3 were identified using a gas detector (Dräger X-am 7000). In a glass bottle filled with urea dissolved in PBS (10 mL, 200


 mM) or acetate buffer (10 mL, 200 mM), UrNMs (2.5 mg) were added, and the caps were securely fastened to prevent gas release. After 30 min, the bottle caps were removed, and the probe of


the gas detector was placed over the solutions to record the generated gases. VIDEOS ANALYSIS To investigate the dynamics of swarms over time, the recorded videos were analyzed by pixel


intensity distribution and density maps. For pixel intensity distribution, snapshots of videos were captured at 12-second intervals. Subsequently, a region of interest (ROI) measuring 300


pixels by 300 pixels was selected, and the pixel intensity distribution within the ROI was analyzed using ImageJ software. To perform density map analysis, the videos were initially


processed to remove the background using ImageJ software. Then 40-second segments were extracted from these videos. The cumulative pixel intensity of these segments, consisting of 1000


frames each, was computed and visualized using the turbo colormap. PARTICLE IMAGE VELOCIMETRY (PIV) The PIV of recorded videos was conducted by a custom Python code based on the OpenPIV


library. The consecutive frames of videos within desired time intervals were extracted and then loaded into the code OpenPIV, with an interrogation window size of 32 × 32 pixels


(width×height), an overlap of 16 × 16 pixels (horizontal×vertical), and a frame rate of 3.33 fps. The results were then reloaded into the Python code to adjust the arrow size and display


particle velocities in colour bars. DATA AVAILABILITY All data supporting the findings are available within the article and the Supplementary Information. The raw data generated in this


study have been deposited in the Figshare database https://doi.org/10.6084/m9.figshare.27134523.v169. CODE AVAILABILITY The custom scripts used for computational analysis in this study are


available on GitHub at https://github.com/SC357/Convective_Dynamics70. The code is provided under the MIT License. Additional information can be found in the repository’s README file and can


also be requested from the corresponding author. REFERENCES * Wang, H. & Pumera, M. Coordinated behaviors of artificial micro/nanomachines: From mutual interactions to interactions with


the environment. _Chem. Soc. Rev._ 49, 3211–3230 (2020). Article  PubMed  Google Scholar  * Shim, G., Devenport, D. & Cohen, D. J. Overriding native cell coordination enhances external


programming of collective cell migration. _Proc. Natl Acad. Sci. USA._ 118, e2101352118 (2021). Article  PubMed  PubMed Central  CAS  Google Scholar  * Peleg, O., Peters, J. M., Salcedo, M.


K. & Mahadevan, L. Collective mechanical adaptation of honeybee swarms. _Nat. Phys._ 14, 1193–1198 (2018). Article  CAS  Google Scholar  * Zitterbart, D. P., Wienecke, B., Butler, J. P.


& Fabry, B. Coordinated movements prevent jamming in an emperor penguin huddle. _PLoS One_ 6, e20260 (2011). Article  ADS  PubMed  PubMed Central  CAS  Google Scholar  * Yang, M. et al.


Swarming magnetic nanorobots bio-interfaced by heparinoid-polymer brushes for in vivo safe synergistic thrombolysis. _Sci. Adv._ 9, eadk7251 (2023). Article  PubMed  PubMed Central  CAS 


Google Scholar  * Yu, J. et al. Active generation and magnetic actuation of microrobotic swarms in bio-fluids. _Nat. Commun._ 10, 5631–5642 (2019). Article  ADS  PubMed  PubMed Central  CAS


  Google Scholar  * Xie, H. et al. Reconfigurable magnetic microrobot swarm: multimode transformation, locomotion, and manipulation. _Sci. Robot._ 4, eaav8006 (2019). Article  PubMed  Google


Scholar  * Bente, K. et al. Selective actuation and tomographic imaging of swarming magnetite nanoparticles. _ACS Appl. Mater. Interfaces_ 4, 6752–6759 (2021). CAS  Google Scholar  * Chen,


M. et al. Programmable dynamic shapes with a swarm of light-powered colloidal motors. _Angew. Chem. Int. Ed._ 60, 16674–16679 (2021). Article  ADS  CAS  Google Scholar  * Aubret, A.,


Youssef, M., Sacanna, S. & Palacci, J. Targeted assembly and synchronization of self-spinning microgears. _Nat. Phy._ 14, 1114–1118 (2018). Article  CAS  Google Scholar  * Xu, T. et al.


Reversible swarming and separation of self-propelled chemically powered nanomotors under acoustic fields. _J. Am. Chem. Soc._ 137, 2163–2166 (2015). Article  PubMed  CAS  Google Scholar  *


Tang, S. et al. Structure-dependent optical modulation of propulsion and collective behavior of acoustic/light-driven hybrid microbowls. _Adv. Funct. Mater._ 29, 1809003–1809009 (2019).


Article  ADS  Google Scholar  * Leunissen, M. E., Vutukuri, H. R. & Van Blaaderen, A. Directing colloidal self-assembly with biaxial electric fields. _Adv. Mater._ 21, 3116–3120 (2009).


Article  CAS  Google Scholar  * Ma, F., Wang, S., Wu, D. T. & Wu, N. Electric-field-induced assembly and propulsion of chiral colloidal clusters. _Proc. Natl Acad. Sci. USA._ 112,


6307–6312 (2015). Article  ADS  PubMed  PubMed Central  CAS  Google Scholar  * Duan, W., Liu, R. & Sen, A. Transition between collective behaviors of micromotors in response to different


stimuli. _J. Am. Chem. Soc._ 135, 1280–1283 (2013). Article  PubMed  CAS  Google Scholar  * Gentile, K., Somasundar, A., Bhide, A. & Sen, A. Chemically powered synthetic “living”


systems. _Chem_ 6, 2174–2185 (2020). Article  CAS  Google Scholar  * Ziepke, A., Maryshev, I., Aranson, I. S. & Frey, E. Multi-scale organization in communicating active matter. _Nat.


Commun_. 13, 6727–6736 (2022). * Zhang, F. et al. Extremophile-based biohybrid micromotors for biomedical operations in harsh acidic environments. _Sci. Adv._ 8, eade645 (2022). Article  ADS


  Google Scholar  * Chen, H. et al. An engineered bacteria-hybrid microrobot with the magnetothermal bioswitch for remotely collective perception and imaging-guided cancer treatment. _ACS


Nano_ 16, 6118–6133 (2022). Article  PubMed  CAS  Google Scholar  * Zhuang, J. & Sitti, M. Chemotaxis of bio-hybrid multiple bacteria-driven microswimmers. _Sci. Rep._ 6, 32135–32144


(2016). Article  ADS  PubMed  PubMed Central  CAS  Google Scholar  * Akter, M. et al. Cooperative cargo transportation by a swarm of molecular machines. _Sci. Robot._ 7, eabm0677 (2022).


Article  PubMed  CAS  Google Scholar  * Keya, J. J. et al. DNA-assisted swarm control in a biomolecular motor system. _Nat. Commun._ 9, 453–460 (2018). Article  ADS  PubMed  PubMed Central 


Google Scholar  * Chen, S. et al. Dual-source powered nanomotor with integrated functions for cancer photo-theranostics. _Biomaterials_ 288, 121744–121753 (2022). Article  PubMed  CAS 


Google Scholar  * Chen, S. et al. Active nanomotors surpass passive nanomedicines: current progress and challenges. _J. Mater. Chem. B_ 10, 7099–7107 (2022). Article  PubMed  CAS  Google


Scholar  * Ma, X. et al. Enzyme-powered hollow mesoporous Janus nanomotors. _Nano Lett._ 15, 7043–7050 (2015). Article  ADS  PubMed  Google Scholar  * Ma, X., Hortelao, A. C., Miguel-López,


A. & Sánchez, S. Bubble-free propulsion of ultrasmall tubular nanojets powered by biocatalytic reactions. _J. Am. Chem. Soc._ 138, 13782–13785 (2016). Article  PubMed  PubMed Central 


CAS  Google Scholar  * Arqué, X. et al. Intrinsic enzymatic properties modulate the self-propulsion of micromotors. _Nat. Commun._ 10, 2826–2837 (2019). Article  ADS  PubMed  PubMed Central


  Google Scholar  * Ma, X. & Sanchez, S. A bio-catalytically driven Janus mesoporous silica cluster motor with magnetic guidance. _Chem. Commun._ 51, 5467–5470 (2015). Article  CAS 


Google Scholar  * Hortelão, A. C., Patiño, T., Perez-Jiménez, A., Blanco, À. & Sánchez, S. Enzyme-powered nanobots enhance anticancer drug delivery. _Adv. Funct. Mater._ 28,


1705086–1705095 (2018). Article  Google Scholar  * Llopis-Lorente, A. et al. Enzyme-powered gated mesoporous silica nanomotors for on-command intracellular payload delivery. _ACS Nano_ 13,


12171–12183 (2019). Article  PubMed  CAS  Google Scholar  * Hortelao, A. C., Carrascosa, R., Murillo-Cremaes, N., Patino, T. & Sánchez, S. Targeting 3D bladder cancer spheroids with


urease-powered nanomotors. _ACS Nano_ 13, 429–439 (2019). Article  PubMed  CAS  Google Scholar  * Tang, S. et al. Enzyme-powered Janus platelet cell robots for active and targeted drug


delivery. _Sci. Robot._ 5, eaba6137 (2020). Article  PubMed  Google Scholar  * Choi, H., Cho, S. H. & Hahn, S. K. Urease-powered polydopamine nanomotors for intravesical therapy of


bladder diseases. _ACS Nano_ 14, 6683–6692 (2020). Article  PubMed  CAS  Google Scholar  * Patino, T. et al. Self-sensing enzyme-powered micromotors equipped with pH-responsive DNA


nanoswitches. _Nano Lett._ 19, 3440–3447 (2019). Article  ADS  PubMed  CAS  Google Scholar  * Liu, X. et al. Urease-powered micromotors with spatially selective distribution of enzymes for


capturing and sensing exosomes. _ACS Nano_ 17, 24343–24354 (2023). Article  PubMed  CAS  Google Scholar  * Hortelao, A. C. et al. Swarming behavior and in vivo monitoring of enzymatic


nanomotors within the bladder. _Sci. Robot._ 6, eabd2823 (2021). Article  PubMed  Google Scholar  * Simó, C. et al. Urease-powered nanobots for radionuclide bladder cancer therapy. _Nat.


Nanotechnol._ 19, 554–564 (2024). Article  ADS  PubMed  PubMed Central  Google Scholar  * Serra-Casablancas, M. et al. Catalase-powered nanobots for overcoming the mucus barrier. _ACS Nano_


18, 16701–16714 (2024). Article  PubMed  PubMed Central  CAS  Google Scholar  * Ruiz-González, N. et al. Swarms of enzyme-powered nanomotors enhance the diffusion of macromolecules in


viscous media. _Small_ 20, 2309387–2309403 (2024). Article  Google Scholar  * Ramos-Docampo, M. A., Wang, N., Pendlmayr, S. & Städler, B. Self-propelled collagenase-powered


nano/micromotors. _ACS Appl. Nano Mater._ 5, 14622–14629 (2022). Article  CAS  Google Scholar  * Ramos-Docampo, M. A. et al. Microswimmers with heat delivery capacity for 3D cell spheroid


penetration. _ACS Nano_ 13, 12192–12205 (2019). Article  PubMed  CAS  Google Scholar  * Fraire, J. C. et al. Light-triggered mechanical disruption of extracellular barriers by swarms of


enzyme-powered nanomotors for enhanced delivery. _ACS Nano_ 17, 7180–7193 (2023). Article  PubMed  PubMed Central  CAS  Google Scholar  * Kessler, J. O. & Hill, N. A. The growth of


bioconvection patterns in a uniform suspension of gyrotactic micro-organisms. _J. Fluid Mech._ 195, 223–237 (1988). Article  ADS  MathSciNet  PubMed  Google Scholar  * Bodenschatz, E.,


Pesch, W. & Ahlers, G. Recent developments in Rayleigh-Bénard Convection. _Annu. Rev. Fluid Mech._ 32, 709–778 (2000). Article  ADS  Google Scholar  * Bees, M. A. Advances in


bioconvection. _Annu. Rev. Fluid Mech._ 52, 449–476 (2020). Article  ADS  MathSciNet  Google Scholar  * Metcalfe, A. M. & Pedley, T. J. Falling plumes in bacterial bioconvection. _J.


Fluid Mech._ 445, 121–149 (2001). Article  ADS  Google Scholar  * Zhang, J. et al. Light-powered, fuel-Free oscillation, migration, and reversible manipulation of multiple cargo types by


micromotor swarms. _ACS Nano_ 17, 251–262 (2023). Article  PubMed  CAS  Google Scholar  * Sun, M. et al. Bioinspired self-assembled colloidal collectives drifting in three dimensions


underwater. _Sci. Adv._ 9, eadj4201 (2023). Article  ADS  PubMed  PubMed Central  Google Scholar  * Kumar, B. V. V. S. P., Patil, A. J. & Mann, S. Enzyme-powered motility in buoyant


organoclay/DNA protocells. _Nat. Chem._ 10, 1154–1163 (2018). Article  ADS  PubMed  CAS  Google Scholar  * Song, J., Shklyaev, O. E., Sapre, A., Balazs, A. C. & Sen, A. Self‐propelling


macroscale sheets powered by enzyme pumps. _Angew. Chem. Int. Ed._ 63, e202311556 (2023). Article  Google Scholar  * Sapre, A. et al. Enzyme catalysis causes fluid flow, motility, and


directional transport on supported lipid bilayers. _ACS Appl. Mater. Interfaces_ 16, 9380–9387 (2024). Article  PubMed  CAS  Google Scholar  * Patiño, T. et al. Influence of enzyme quantity


and distribution on the self-propulsion of non-Janus urease-powered micromotors. _J. Am. Chem. Soc._ 140, 7896–7903 (2018). Article  PubMed  Google Scholar  * Patiño, T., Llacer-Wintle, J.,


Pujals, S., Albertazzi, L. & Sánchez, S. Unveiling protein corona formation around self-propelled enzyme nanomotors by nanoscopy. _Nanoscale_ 16, 2904–2912 (2023). Article  Google


Scholar  * De Corato, M. et al. Self-propulsion of active colloids via ion release: theory and experiments. _Phys. Rev. Lett._ 124, 108001–108006 (2020). Article  ADS  PubMed  Google Scholar


  * De Corato, M., Pagonabarraga, I., Abdelmohsen, L. K. E. A., Sánchez, S. & Arroyo, M. Spontaneous polarization and locomotion of an active particle with surface-mobile enzymes. _Phys.


Rev. Fluids_ 5, 122001–122011 (2020). Article  ADS  Google Scholar  * Arqué, X. et al. Ionic species affect the self-propulsion of urease-powered micromotors. _Research_ 2020, 1–14 (2020).


* Simmchen, J. et al. Topographical pathways guide chemical microswimmers. _Nat. Commun._ 7, 10598–10606 (2016). Article  ADS  PubMed  PubMed Central  CAS  Google Scholar  * Katuri, J.,


Uspal, W. E., Popescu, M. N. & Sánchez, S. Inferring non-equilibrium interactions from tracer response near confined active Janus particles. _Sci. Adv._ 7, eabd0719 (2021). Article  ADS


  PubMed  PubMed Central  CAS  Google Scholar  * Katuri, J., Caballero, D., Voituriez, R., Samitier, J. & Sanchez, S. Directed flow of micromotors through alignment interactions with


micropatterned ratchets. _ACS Nano_ 12, 7282–7291 (2018). Article  PubMed  CAS  Google Scholar  * Feng, Y. et al. Self-adaptive enzyme-powered micromotors with switchable propulsion


mechanism and motion directionality. _Appl. Phys. Rev._ 8, 011406–011413 (2021). Article  ADS  CAS  Google Scholar  * Hales, J. M. & Drewes, D. R. Solubility of ammonia in water at low


concentrations. _Atmos. Environ._ 13, 1133–1147 (1979). Article  ADS  CAS  Google Scholar  * Gottwald, F. et al. Physical properties of ammonia solutions. Ammonium nitrate-ammonia-water and


urea-ammonia-water. _Ind. Eng. Chem._ 44, 910–913 (1952). Article  Google Scholar  * Wang, X., Conway, W., Burns, R., McCann, N. & Maeder, M. _Phys. Chem. A_ 114, 1734–1740 (2010).


Article  CAS  Google Scholar  * Estiu, G. & Merz, K. M. Catalyzed decomposition of urea. Molecular dynamics simulations of the binding of urea to urease. _Biochemistry_ 45, 4429–4443


(2006). Article  PubMed  CAS  Google Scholar  * Sarigul, N., Korkmaz, F. & Kurultak, İ. A new artificial urine protocol to better imitate human urine. _Sci. Rep._ 9, 20159–20169 (2019).


Article  ADS  PubMed  PubMed Central  CAS  Google Scholar  * Ishii, M. & Mishima, K. Two-fluid model and hydrodynamic constitutive relations. _Nucl. Eng. Des._ 82, 107–126 (1984).


Article  CAS  Google Scholar  * Valdez, L., Shum, H., Ortiz-Rivera, I., Balazs, A. C. & Sen, A. Solutal and thermal buoyancy effects in self-powered phosphatase micropumps. _Soft Matter_


13, 2800–2807 (2017). Article  ADS  PubMed  CAS  Google Scholar  * Aranson, I. S. & Sapozhnikov, M. V. Theory of pattern formation of metallic microparticles in poorly conducting


liquids. _Phys. Rev. Lett._ 92, 234301–234304 (2004). Article  ADS  PubMed  CAS  Google Scholar  * Chen, S. et al. Collective buoyancy-driven dynamics of enzymatic nanomotors. _Figshare_,


https://doi.org/10.6084/m9.figshare.27134523.v1 (2024). * Chen, S. et al. Collective buoyancy-driven dynamics of enzymatic nanomotors. _Github_, https://doi.org/10.5281/zenodo.13842706


(2024). Download references ACKNOWLEDGEMENTS The research leading to these results has received funding from the grants PID2021-128417OB-I00 and PDC2022-133753-I00 funded by MCIN/AEI/


10.13039/501100011033 and, by “ERDF A way of making Europe” and European Union Next Generation EU, (Bots4BB and BOJOS projects) (S.S.). This project has also received funding from the


European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program (grant agreement No 866348, iNanoSwarms) (S.S.). The IBEC team wishes to thank the


CERCA programme of the Generalitat de Catalunya, the Secretaria d’Universitats i Recerca del Departament d’Empresa i Coneixement de la Generalitat de Catalunya through the project 2021 SGR


01606, and the “Centro de Excelencia Severo Ochoa”, funded by Agencia Estatal de Investigación (CEX2018-000789-S). We thank Santiago Marco Colás and Eduardo Caballero Saldivar for their


technical support. S.S. and S.C. acknowledge the Predoctoral AGAUR-FI Joan Oró grant (2023 FI-1 00654) funded by “Secretaria d’Universitats i Recerca del Departament de Recerca i


Universitats de la Generalitat de Catalunya” and by European Social Fund Plus. The research of I.S.A. was supported by the NSF award PHY-2140010. AUTHOR INFORMATION AUTHORS AND AFFILIATIONS


* Institute for Bioengineering of Catalonia (IBEC), The Barcelona Institute for Science and Technology (BIST), Baldiri i Reixac 10-12, Barcelona, 08028, Spain Shuqin Chen, Xander Peetroons, 


Anna C. Bakenecker, Florencia Lezcano & Samuel Sánchez * Faculty of Physics, University of Barcelona, Martí i Franquès 1, Barcelona, Spain Shuqin Chen * Department of Biomedical


Engineering, The Pennsylvania State University, University Park, PA, 16802, USA Igor S. Aranson * Department of Chemistry, The Pennsylvania State University, University Park, PA, 16802, USA


Igor S. Aranson * Department of Mathematics, The Pennsylvania State University, University Park, PA, 16802, USA Igor S. Aranson * Catalan Institute for Research and Advanced Studies (ICREA),


Passeig Lluís Companys 23, Barcelona, Spain Samuel Sánchez Authors * Shuqin Chen View author publications You can also search for this author inPubMed Google Scholar * Xander Peetroons View


author publications You can also search for this author inPubMed Google Scholar * Anna C. Bakenecker View author publications You can also search for this author inPubMed Google Scholar *


Florencia Lezcano View author publications You can also search for this author inPubMed Google Scholar * Igor S. Aranson View author publications You can also search for this author inPubMed


 Google Scholar * Samuel Sánchez View author publications You can also search for this author inPubMed Google Scholar CONTRIBUTIONS S.C. synthesized and characterized the nanomotors,


performed the motion and velocity analysis, analyzed the data, and wrote the paper. X.P. developed Python scripts for the upward velocity analysis and developed Matlab scripts for the


computational modelling. A.C.B. designed the side view setup and contributed to the simulation work. F.L. analyzed the centre of mass tracking and contributed to the data processing. I.S.A.


conceived the initial idea, developed the computational modelling, and edited the paper. S.S. conceived the initial idea, supervised the project, and edited the paper. CORRESPONDING AUTHORS


Correspondence to Igor S. Aranson or Samuel Sánchez. ETHICS DECLARATIONS COMPETING INTERESTS The authors declare no competing interests. PEER REVIEW PEER REVIEW INFORMATION _Nature


Communications_ thanks Jianguo Guan, Akira Kakugo and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available. ADDITIONAL


INFORMATION PUBLISHER’S NOTE Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. SUPPLEMENTARY INFORMATION SUPPLEMENTARY


INFORMATION DESCRIPTION OF ADDITIONAL SUPPLEMENTARY FILES SUPPORTING MOVIE 1 THE UNIVERSALITY OF THE MECHANISM SUPPORTING MOVIE 2 CONTROL FACTORS INFLUENCE ENZYMATIC COLLECTIVE BEHAVIOUR


SUPPORTING MOVIE 3 CONTROL FACTORS INFLUENCE THE MOTION OF MSNPS SUPPORTING MOVIE 4 THE PH CHANGE OF THE COLLECTIVE DYNAMICS SUPPORTING MOVIE 5 BUBBLES PRODUCED BY ENZYMATIC CATALYSIS


REACTION SUPPORTING MOVIE 6 VERTICAL CONFINEMENT SHAPES COLLECTIVE BEHAVIOUR SUPPORTING MOVIE 7 COMPUTATIONAL MODELLING OF THE CONTROL FACTORS SUPPORTING MOVIE 8 COMPUTATIONAL MODELLING OF


THE VERTICAL CONFINEMENT EFFECTS TRANSPARENT PEER REVIEW FILE RIGHTS AND PERMISSIONS OPEN ACCESS This article is licensed under a Creative Commons Attribution 4.0 International License,


which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link


to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless


indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or


exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. Reprints


and permissions ABOUT THIS ARTICLE CITE THIS ARTICLE Chen, S., Peetroons, X., Bakenecker, A.C. _et al._ Collective buoyancy-driven dynamics in swarming enzymatic nanomotors. _Nat Commun_ 15,


9315 (2024). https://doi.org/10.1038/s41467-024-53664-w Download citation * Received: 29 February 2024 * Accepted: 15 October 2024 * Published: 29 October 2024 * DOI:


https://doi.org/10.1038/s41467-024-53664-w SHARE THIS ARTICLE Anyone you share the following link with will be able to read this content: Get shareable link Sorry, a shareable link is not


currently available for this article. Copy to clipboard Provided by the Springer Nature SharedIt content-sharing initiative


Trending News

Everything you should know about herpes gladiatorum

Herpes gladiatorum, also known as mat herpes, is a common skin condition caused by herpes simplex virus type 1 (HSV-1). ...

National trust keeps colonial art to encourage debate

The National Trust plans to keep the majority of “distressing” objects linked to slavery and colonial oppression on disp...

Urgent dairy recall: are your milk and cream products affected?

NP AND DJ DARWIN LTD WHOLE MILK Pack Size - All Use by - All dates up to and including 29 November 2019 The products are...

Climate change will affect global water availability through compounding changes in seasonal precipitation and evaporation

ABSTRACT Both seasonal and annual mean precipitation and evaporation influence patterns of water availability impacting ...

Climbers rescue dog after it spent a month at the top of pico de orizaba

A dog that followed a group of hikers up the Pico de Orizaba volcano and stayed on the peak for almost a month has been ...

Latests News

Collective buoyancy-driven dynamics in swarming enzymatic nanomotors

ABSTRACT Enzymatic nanomotors harvest kinetic energy through the catalysis of chemical fuels. When a drop containing nan...

Biofilm formation in the lung contributes to virulence and drug tolerance of mycobacterium tuberculosis

ABSTRACT Tuberculosis is a chronic disease that displays several features commonly associated with biofilm-associated in...

Murray's major qf streak ends with us open loss

Andy Murray watched a 130 mph ace zoom by to create a two-set deficit at the U.S. Open, and then sat in his changeover c...

Critic's pick: steve coogan in 'philomena'

Much has been made of Judi Dench’s finely nuanced turn in director Stephen Frears’ excellent new drama, “Philomena.” But...

Publisher correction: death by inflammation: drug makers chase the master controller

Correction to: _Nature Biotechnology_ https://doi.org/10.1038/s41587-019-0023-4, published online 4 February 2019. In th...

Top