Next Article in Journal
Power System Fault Diagnosis Method Based on Deep Reinforcement Learning
Next Article in Special Issue
Mass Transfer Behaviors and Battery Performance of a Ferrocyanide-Based Organic Redox Flow Battery with Different Electrode Shapes
Previous Article in Journal
Reservoir Permeability Calculation under Flow Unit Control
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Principle, Fabrication and Emerging Applications of Nanobottle Motor

1
Beijing Key Laboratory of Multiphase Flow and Heat Transfer for Low Grade Energy Utilization, North China Electric Power University, Beijing 102206, China
2
Department of Microsystems, University of South-Eastern Norway, 3184 Horten, Norway
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(20), 7636; https://doi.org/10.3390/en15207636
Submission received: 24 August 2022 / Revised: 23 September 2022 / Accepted: 10 October 2022 / Published: 16 October 2022

Abstract

:
Micro/nano-motors play an important role in energy, environment, and biomedicines. As a new type of nano-motors, nanobottles attract great attention due to their distinct advantages of a large cavity, high specific surface area, bionic streamline structure, and chemotactic motion. Here, we systematically review the development of nanobottle motors from aspects of propulsion mechanisms, fabrication methods and potential applications. Firstly, three types of propulsive modes are summarized, with focus on chemical propulsion, light driving and magnetic actuation. We then discuss the fabrication methods of nanobottles, including the soft-template-based hydrothermal method and the swelling-inducement and wet-chemistry methods. The potential applications of nanobottle motors are additionally highlighted in energy, environmental, and biomedical fields. Finally, the future challenges and outlooks of nanobottle motors are discussed for the further development of this technology.

1. Introduction

Since the invention of the steam engine, various motors with macro size, shape and composition have brought tremendous changes to industry and daily life. To have a better understanding of our environments, human beings have also started to explore the micro-world. As nanotechnology advances, self-propelled motors with a size small enough to voyage into organisms have attracted extensive research interests. Inspired by biological motors (e.g., cells or certain bacteria) [1], researchers proposed the concept of micro/nano-motors. Typical sizes of the micro/nano-motors are in the range from tens of nanons to dozens of microns [2], commonly referred to as micro-swimmers, micro-machines, and micro-robots. Different from nanoparticles that only perform Brownian motion based on passive diffusion, the micro/nano-motors can convert various external energy into driving force to perform specific mechanical motions as well [3,4,5,6]. Owing to the advantages of microscale, autonomous motion, surface functionalization, adjustability and diverse shapes, micro/nano-motors have become an important tool for scholars to explore the microscopic world [7,8,9,10] and have been used in energy, environment and biomedicine fields [11,12,13,14,15,16].
Different geometries of micro/nano-motors, such as helical, tubular, Janus sphere, nanowire, and other biomimetic shapes have been developed to meet the requirement for practical applications [17,18,19,20,21,22]. Among them, nanobottles have a large cavity, high specific surface area and bionic streamline structure, as well as the characteristics of chemotactic motion of perceived gradient, and adjust direction trend gradient [23,24], showing great potential for energy, environmental, and biomedical applications [25,26,27]. However, the construction of nanobottles is challenged because of difficulties in forming open structures and inadaptability of macroscopic principles. Various strategies have been proposed to manufacture such nanobottles. Early composited nanobottles derived from hydrothermal methods do not have a lid to be opened or closed, which is referred to as inorganic nanobottles [28]. The advanced technologies have fabricated nanobottles with thermo-sensitive phase changed material lid that could only be used once [29]. A few studies have summarized the preparation methods of nanobottles [30,31,32,33], and their fabrication techniques have potential for practical applications.
Although nanobottles are promising for nano-motors, the technology in this area is still at primary stage. Existing studies were mostly aimed at how to form nanobottles, especially in the formation of open structures. The topic of the nanobottle motor has not been systematically studied and summarized. To accelerate the development of nanobottle technology, this work introduces the concept of nanobottle motors from aspects of the propulsion mechanisms, fabrication methods and their emerging applications (Figure 1). The propulsion mechanism is first discussed, including chemical propulsion, light-driving and magnetic actuation. We then summarize the fabrication methods of nanobottles that involve the soft-template-based hydrothermal method and the swelling-inducement and wet-chemistry methods. The emerging applications of nanobottle motors are additionally highlighted in energy, environmental and biomedical fields. Finally, a summary and prospects are given for the future development of this technology. This work provides basic knowledge and encourages researchers to develop new types of nanobottles for diverse applications.

2. Propulsion Mechanisms

The propulsion mechanisms of nanomotors relies on the asymmetric field of chemical products or energy to break the static equilibrium and achieve autonomous motion. The general expression is described by
U = b ψ
where U is the motor velocity, b is the velocity coefficient, and ∇ψ is the gradient of potential function ψ including pressure, potential, concentration, temperature [44]. Based on these factors, nanobottle motors can be classified as three common types of gradient field derived from chemical sources, light or magnetic field.

2.1. Chemical Propulsion

Chemical propulsion is a dominant driving mode due to its advantages of a simple principle, strong power, facile preparation, and common equipment. In such a mode, it requires the introduction of chemical fuels and catalysts into the environment to promote autonomous motion by chemical reaction. Chemical propulsion can be further classified as self-diffusiophoresis, self-electrophoresis and bubble propulsion [45,46,47,48]. Phoretic motor is generally perceived as a separate particle suspended in an infinite fluid containing a diluted dissolved solute. In the circumstances, the solute concentration is described through the species conservation equation:
c i t + · j i = 0
where ci is the species concentration, and ji is the flux that expressed by generalized Nernst–Planck equation including advection (in the dilute limit):
j i = c i u D i c i + c i k B T φ i
where u is the fluid velocity, D is the solute diffusivity, kB is the Boltzmann’s constant, T is the temperature, and φi is a generalized interaction potential that represents overall mutual effect between solute species and moving motors. The three terms on the right side of Equation (3) express the advection, diffusion and interaction caused by great gradient of φ, respectively [49]. Furthermore, at the micro/nano scale, inertia force is ignored, whereas viscous force acts a dominant role. Thus, the micro/nano-motors move in low Reynolds number flows, and the momentum balance of fluid is described by Navier–Stokes Equations:
· u = 0 ,   η 2 u = p + F b
where u is the principal part velocity, η is the fluid viscosity, p is the pressure, and Fb is the generalized body force. Equations (2)–(4) describe the basic system of equations governing self-electrophoresis and self-diffusiophoresis.
Since the current research on nanobottle motors by chemical propulsion only includes self-diffusiophoresis and bubble propulsion, we mainly expound the above two methods described in this section.
Self-diffusiophoresis refers to the chemical reaction on nanobottle surface which forms asymmetrical concentration gradient to propel the motor [50,51]. Since different molecules interact differently with particles, the potential is defined as −∇ψFs, where Fs is the net force experienced through solute molecule. This force can be integrated and eventually converted into a force expressed as Fb = −c∇ψ, which corresponds to the third term of Equation (4) [44]. The chemical catalyst reaction can occur inside the nanobottle, in which the chemical gradient is the maximum near the opening (Figure 2a). Nanobottle motors have the motion with axis of symmetry because of the self-diffusiophoresis. Self-diffusiophoretic force (expressed as Fd = −α∂c) mainly acts on the opening around the motor surface. The bottom diameter of nanobottle is larger than the opening, indicating that the fluid friction (Ff = v/μ) is mainly applied to the bottom of nanobottle (Figure 2b). Hydrophilic nanobottle applies a negative force on the surrounding fluid and form a puller-like flow field (Figure 2c). However, the hydrophobic nanobottle motor exerts a positive force on the surrounding fluid and generates a pusher-like field (Figure 2d) [35].
Bubble propulsion is less affected by electrolyte and has the advantages of a large driving force and high energy utilization rate to sustain movement [53]. The mechanism begins with a chemical reaction between fuels (e.g., H2O2, active metals, CaCO3, etc.) and catalysts to form gas molecules.
2 H 2 O 2 = 2 H 2 O + O 2
With the reaction, gas molecules gradually increase in the cavity. Due to the limitation of cavity, small bubbles are easier to aggregate and produce large bubbles. Under the pressure, large bubbles are ejected from the opening of nanobottle. The reverse driving force formed by bubble separation leads to the motor having the opposite motion (Figure 2e) [52,54]. For example, a nanobottle loaded in Pt nanoparticles can swim in H2O2 solution via O2 detachment (Equation (5)) [55]. Compared with self-diffusiophoresis, bubble propulsion needs a higher reactant concentration to produce bubbles as well as lower surface tension in solution to form bubble molecules. However, it is not sensitive to salt ions and, thus, can be implemented in complex environments.

2.2. Light Driving

As an external stimulus, light is capable of adjusting parameters such as light intensity, light frequency, polarization, propagative direction, as well as precise input tuning energy or select on/off mode to achieve remote control [56]. Light-driven motors are thus considered as a facile propulsion mode, due to the advantages of high controllability, good programmability and easy operation [57,58,59]. The principle of light-driven motors depends on breaking the symmetry of pressure upon light irradiation, leading to the motion of micro/nano-motors [50]. Near-infrared light is often used for propulsion because its characteristics of strong penetration, ideal transmittance, and controllable irradiated area [60,61,62]. Photothermic materials (e.g., precious metals, carbon, etc.) of motors absorb near-infrared light that leads to fluid temperature in cavity rises rapidly and produces local temperature gradient. The temperature gradient increases the internal pressure of the motor and forces the heated fluid to be ejected out from the bottleneck. Hereby, it is the inside bottle, under more pressure than the outside bottle, that creates a net force at the bottom. Nanobottle motors can thus move in the direction of temperature gradient. In addition, the open/close motion of carbonaceous nanobottle motor can be controlled by the near-infrared switch state (Figure 3) [36]. The light-driven thermophoresis force can be expressed as [63]:
F = 9 π d p η 2 k a 2 ρ g T k p T r , t
where dp is the motor diameter, η is the fluid viscosity, ka is the thermal conductivity of fluid, kp is the thermal conductivity of motor, and T (r, t) is the temperature increment as a function of special and time coordinates.

2.3. Magnetic Actuation

Magnetic field features the qualities of high penetration, noninvasive and high controllability [64], so it is an attractive method to control the motion of motors. The mechanism is to exert magnetic force or torque to magnetized objects. The magnetic force (F) and magnetic torque (T) on the magnetic motors can be described as [65]:
F = v M × B
T = v M × B
where v is the volume of magnetic motors, M is the magnetization, and B is the magnetic flux density. We can see that the magnetic force on a magnetic nanobottle motor is zero even in a magnetic field. The magnetic torque relies on the direction of the magnetic dipole moment and magnetic field. The magnetic field must be either time varying (e.g., rotating, oscillating and pulsed magnetic fields) or space varying (e.g., gradient magnetic field) to realize the continuous motion of nanobottle motors [66]. When a magnetic field gradient is imposed, nanobottle motors are subjected a magnetic force that pulls them along the direction dictated by the magnetic field gradient [67]. In such a process, nanobottles can be propelled in the direction of the magnetic field (Figure 4).

3. Fabrication Methods

Nanobottles refer to hollow colloidal particles with a single opening on the surface. The hollow cavity has high load capacity as warehouse for various cargos and can also be used as a place for reaction to generate driving force. Compared with the hollow particles with entirely sealed shell structure, the hollow particles with holes have large internal and external specific surface areas, which provide more reaction or adsorption sites for functional cargos and channels for the release of functional cargos. The key parameters such as radii and lengths of nanobottle have great impact on their motion, and thus, controllable preparation of nanobottles is critically important for practical application. Nanobottles of different materials (e.g., SiO2, Mg, carbon and precious metals), sizes and shapes have been developed for practical application. The fabrication methods generally include soft-template, swelling-inducement, and wet-chemistry methods. Meanwhile, we briefly list main references of fabrication methods in Table 1.

3.1. Soft-Template Method

Hydrothermal method refers to that in a closed system under high temperature and high pressure to grow crystals in aqueous solution or water vapor. This method has been broadly used in preparation of nanobottles. To control the nanobottle geometry, the hydrothermal method contains hard template and soft template. Soft-template method is carried out through self-assembly between core templates (e.g., organic surfactants, block copolymers, nanoemulsion drops) and precursor molecules to form shape confinement through weak noncovalent bonds (e.g., hydrogen bonds, Van der Waals forces, electrovalent bonds) [68]. Due to the chemical interaction between core templates and precursor molecules, this method relies on self-assembly, nucleation, and growth process, and can be used to control the geometry parameter. These are the major steps to fabricate carbon nanobottle with ribose as the carbon source (as a carbon precursor under hydrothermal condition), oleic acid as the core, and poly(ethylene oxide)-poly(propylene oxide)-poly(ethylene oxide) as surfactant [42]. With temperature increasing, core nanoparticles aggregate and sedimentate to form nanoemulsion. As the reaction is progressed, carbon precursor polymerizes at the emulsion interface, and the polyethylene oxide blocks in surfactant become more hydrophobic to generate swelling of nanoemulsion. The tensile stress is caused by volume expansion of nanoemulsion and eventually leads to polymer shell splitting when the pressure exceeds the critical pressure. At the final stage, continuous filling process leads to the outflow of nanoemulsion to form a new template surface. Precursor continues to aggregate on the surface of the new template to form asymmetric carbon nanobottle (Figure 5).
Soft templates generally in thermodynamics unstable liquid/gas are formed with high deformation ability, which can easily affected by various external parameters (e.g., PH value, additives concentration, solvent and temperature). The soft templates can be easily removed through evaporation or calcination without complicate template removal process, but the uniformity of nanoparticles is often affected. Furthermore, it is necessary to use a large amount of organic substances (e.g., ionic liquids) as reaction medium, which is easy to pollute the environment and not conducive to mass production.

3.2. Swelling-Induced Method

Swelling-induction method relies on swelling polymer microspheres with organic solvent, making organic solvent diffuse from the interior of nanoparticles and then quenching with ethanol to obtain the opening structure [41,69,70,71,72]. The main steps with polystyrene microspheres serving as the starting point [73]. After coating the surface of polystyrene microspheres with SiO2 thin shell, introduce tetrahydrofuran/water emulsion to swell the polystyrene spheres. After absorbing tetrahydrofuran, the swollen polystyrene creates pressure in the shell. When the pressure surpasses a threshold to poke a hole in the shell, the swollen polystyrene will be squeezed out through the opening and reduce the pressure in the shell to ensure only one hole is created. After being quenched with ethanol to from Janus structure (Figure 6a), the swollen extent and the size of Janus particle can be adjusted through increasing the volumetric percentage of solute in tetrahydrofuran/water emulsion (Figure 6b–f). When volume percentage of solvent is less than lower limit, the pressure caused by swelling will not be enough to break the SiO2 shell; while the volume percentage is above the upper limit, dissolution of polystyrene has occurred in shell. Increasing the volume percentage of tetrahydrofuran to 20% causes more polystyrene to protrude through the hole. After quenching through ethanol, the extruded polystyrene spheres have a diameter of 280 nm. When volume percentage of tetrahydrofuran is increased to 30%, the diameter of polystyrene spheres is increased to 333 nm. Further increasing the volume percentage of tetrahydrofuran to 50% would lead to polystyrene being partially dissolved in SiO2 shell. In the pure tetrahydrofuran solution, polystyrene swells to obtain SiO2 nanobottles (Figure 6f).
During swelling induction, the sizes, shapes and structure of nanobottles can be tailored by controlling swell extent through different types or percentage of solvents. In this process, each swollen particle can maintains the spherical shape by quenching the swelling with ethanol. Different from prior methods that rely on Janus particles, this method is based on commercial polystyrene nanospheres, and the hole is automatically punctured through swelling, making it possible to fabricate uniform nanobottles on a large scale.

3.3. Wet-Chemistry Method

Wet chemistry is a general method to synthesize nanobottles by controlling reaction dynamics and thermodynamic parameters to tune the sizes and shapes of nanobottle [74]. Asymmetric Au nanobottle can be prepared with such method (Figure 7a) [40,75]. Firstly, lead acetate and thioacetamide are heat-treated in the solution of cetyltrimethyl ammonium bromide to obtain PbS nanooctahedron with uniform edge lengths and slightly truncated vertex. The PbS nanooctahedron is used as sacrifice template, which can determine the scale parameters of Au nanobottles, and then disperse sulfide nanooctahedron into the growth solution. Au preferentially deposits at one vertex of each octahedron with the resultant core, then grow along four adjacent facets to produce Janus Au/PbS nanostructure. Finally, added weak acids dissolve and remove sulfide components in Au/PbS Janus nanostructures to produce the Au nanobottle. The opening size (d) and the diameter in the direction perpendicular to its symmetrical axis (D) can be adjusted by such a method, which influences the plasma properties of Au nanobottles. Keeping the opening size constant, the cavity volume and overall size of Au nanobottles will expand with the increase in edge length of PbS nanooctahedron. Furthermore, the opening size can be adjusted through controlling the ratio of Au/PbS. With the increase in Au/PbS ratio, the opening size of Au nanobottle first increases, and then it decreases after reaching the maximum, while the overall size and the cavity volume continue to increase in the Au nanobottle (Figure 7b).
In wet-chemistry method, the edge length of PbS nanooctahedron and Au/Pb ratio are crucial factors during Au overgrowth process to adjusting sizes of the Au nanobottles. Moreover, there are three advantages to choosing PbS as sacrificial template for synthesizing Au nanobottles. Firstly, PbS nanocrystals of various shapes can be easily prepared by wet-chemistry method. Secondly, PbS has Fermi level higher than Au, which facilitate the selective deposition of Au on specific surfaces of PbS nanocrystals. Thirdly, PbS is readily dissolved in weak acids.

4. Potential Applications

4.1. Energy

Solar water splitting has been considered as one of the most eco-friendly technologies for clean and renewable energy [76,77,78]. The current solar-to-fuel conversion rate in water splitting is 5–10% or even lower [79]. A lot of efforts have been made in the research of photocatalysts [80]. The photocatalysts usually present themselves in two forms: suspension powders and immobilized film [81]. The absorption and scattering caused by the former result in uneven illumination distribution and, thus, low photon transfer. Photocatalysts film has good photon harvesting properties. However, the internal mass transfer is slow for the immotile film [82,83]. Therefore, it is urgent to design motile photocatalytic carriers. Nano-motors with self-propulsion capability provide an advanced method to solve these problems such as delivery, mass transfer in microcomplex systems. Nano-motors can accelerate the mass transfer rate through the dynamic working mode to improve photocatalytic degradation performance [84,85,86]. Furthermore, light-driven motors can convert light energy into mechanical energy without any pollution, which is quite often applied to photocatalytic systems [87].
Nanobottle motors have large specific surface area and low density, and they can also greatly accelerate the mass transfer rate through motion to improve the photocatalytic degradation. The near infrared light-driven nanobottle motors has been proved to improve the photocatalytic rate [37]. Compared with pure graphitic carbon nitride and TiO2 P25, in the ultraviolet range of 240–280 nm, the samples containing carbon nanobottle motors have strong visible-light absorption ability. The participation of carbon nanobottle motors expands the absorption range in the visible region and improves the photocatalytic activity. By use of trolamine aqueous solution as sacrificial agent, the photocatalytic hydrogen productions are further studied under full spectral light at room temperature. The photocatalytic hydrogen precipitation rate of graphitic carbon nitride is 37.1 μmol mg−1 h−1. An appropriate amount of carbon motors is beneficial to photocatalytic performance. The optimal usage of carbon nanobottle motors is determined to be 10 wt%; the hydrogen precipitation rate of carbon nanobottle/graphitic carbon nitride motors is 78.9 μmol mg−1 h−1, which is 2.1 times higher than that of pure graphitic carbon nitride (Figure 8a). Additionally, the hydrogen production rate of carbon nanobottle/TiO2P25 motors is 158.7 μmol mg−1 h−1, which is 2.3 times larger than that of pristine TiO2 P25 (Figure 8b). Overuse of carbon nanobottles can block part of illumination, leading to a reduction in the photocatalytic efficiency. In addition, the catalytic nanobottle motors containing precious metals are also attractive, due to the catalyst not being easily consumed for continuous motion [43]. To prove the effective motion of nanobottles, several studies have selected solid Au spheres with similar size and the same composition, but with gold-loaded nanobottle motors. The results show that hydrogen production of nanobottle motors is obviously higher than Au spheres. It also confirms that Au spheres exhibit short Brownian motion, and nanobottle motors show long motion trajectory (Figure 8c). UV–Vis data further disclosed that the catalytic efficiency of nanobottle motors was 1.8 times higher than that of the Au sphere. When the content of nanobottle motors was 0.1 wt%, the reduction rate of para-nitrophenol reached 96.5% within 6 min (Figure 8d). These results express the catalytic performance of nanobottle motors is superior to solid spheres because of the super automaticity and high loading capability.
Although nanobottle motors can effectively improve the photocatalytic efficiency and water splitting, there are still some aspects to be improved. Firstly, photocatalytic nanobottle motors have a low speed of motion, with the speed of nanobottle motors ranging from only a few to hundreds of microns per second [34,35]. Secondly, the material of motors driven by infrared light often includes precious metals (e.g., Au, Pt and Ag) as cocatalysts, which makes the fabrication of nanobottle motors a cumbersome and expensive process. Thirdly, light-driven motors generally use the heat generated by photothermic conversion of near-infrared light as a driving force to increase photocatalytic efficiency. This method is generally limited to laboratorial studies. Therefore, future research needs to demonstrate the feasibility of efficient propelled nanobottle motors upon sunlight and their applications in practical production.

4.2. Environment

Water pollution is one of the major forms of environmental pollution. It is estimated that 1.8 million global deaths in the year 2016 can be attributed to water pollution [88]. The major agents of water pollution are heavy metals, dyes, oil, bacteria, microplastics, etc. [89,90,91,92,93]. It is very urgent to conduct monitoring and remediation of the polluted environment. The conventional techniques of wastewater remediation are precipitation, adsorption, membrane separation and chemical methods [94,95]. These methods have several shortcomings. For example, precipitation leads to the formation of secondary sludge. Membrane separation is an energy-intensive and costly process [96]. Furthermore, most disposal methods translate one pollutant into another, for instance, the use of activated carbon for adsorption, which is then discarded without processing. Chemical methods use harmful reagents such as peroxide, coagulants and flocculants [97]. Therefore, the advanced techniques for water cleaning should adhere to two standards: decontamination capacity and reduce additional pollutants [98,99]. Nanomaterials have a more significant filtration effect as compared with larger particles with the same chemicals. The unique structures of nanoparticles make them strong adsorbents (especially for organics), such as dendritic, spherical, tubs and nanobottle [100,101]. The autonomous motion micro/nano-motors can accelerate the mass transfer rate, so the sewage disposal processes will be significantly superior to traditional remediation [102]. Micro/nano-motors have been used for environmental monitoring and remediation due to the remediation process having a more significant performance and higher efficiency [103,104,105].
The fast motion of nanobottle motors in the solution can actively grab the target molecule or ion to improve the adsorbed and separated efficiency. Additionally, with the large specific surface area of nanobottle motors having a strong affinity for organic pollutants, they can be used in environmental remediation. It is a special way to monitor the target molecules in the environment by observing the motion state of nanobottle motors. As a proof of concept, enzyme-driven nanobottle motors were used to monitor heavy medal ions in waste water, while enzymes were rapidly inactivated in a solution containing heavy metal ions and changed the motion velocity of nanobottle motors. Therefore, the existence of heavy metal ions in the solution can be determined through monitoring the motion velocity of nanobottle motors, and one can even roughly estimate the concentration of heavy metal ions (Figure 9a) [106]. Magnetic mesoporous silica nanobottle motors have a high specific surface area and large porosity, and they can also be used for adsorbing and removing heavy metal ions in polluted water. It was shown that the removal rate of Cu2+ in water solution was 60% in the first 5 min, and then it increased slowly over time, eventually reaching more than 80% after 1 h (Figure 9b). Since the removal of Cu2+ is a surface adsorption process, when the concentration of silica nanobottle motors increased from 0.5 to 3 mg mL−1, the removal rate of Cu2+ increased to more than 90% after 30 min (Figure 9c). In addition, nanobottle motors can also be used for oil removal. A magnetic modulate micromotor was developed with monolayer MnFe2O4 nanobottle motor to remove contaminated oil from water (Figure 9d) [38]. The catalytic decomposition of hydrogen peroxide on the inner surface of MnFe2O4 nanobottle motor generates oxygen, which promotes the nanobottle to sustain autonomous motion. Owing to the hydrophobic long oleic acid chains on the surface of nanobottles, oil droplets can be adsorbed through the interaction between the hydrophobic surface and the oil. The nanobottles can thus capture oil droplets under the action of an external magnetic field and autonomous movement to implement oil–water separation (Figure 9e).
Nanobottle motors have the characteristics of autonomic motion that can used for observing the motion state to realize environmental detection. The autonomic motion of nanobottles also improves the mass transfer of solution to raise the degradation efficiency of pollutants. However, there are still some problems in the research of nanobottle motors in environmental application. First, fuel consumption is a major obstacle in the environmental remediation of nanobottle motors. Furthermore, most reports on sewage treatment only use nanobottle motors to treat the virtual sewage prepared in laboratory, rather than the real sewage from sewage plant. Therefore, the motion and the treatment effect of nanobottle motors are unknown in real polluted environment.

4.3. Biomedicine

Biomedicine is one of the most influential areas of research in human society. Various diseases have been found to threaten human life and health, such as malignant tumor (namely cancer), thrombus, gastroenteritis, etc. [107]. Among them, the death rates associated with malignant tumor are second after cardiovascular diseases, and malignant tumor has become the most concerned of global public healthcare [108]. The current treatment methods of malignant tumor including surgery, radiotherapy, chemotherapy and immunotherapy [109]. Chemotherapy is one of the main cures for cancer. However, this treatment is to passively deliver the drug to the diseased site. The drugs stay on or around the surface, unable to penetrate into deeper areas, and leads to a clinical therapeutic effect that is not ideal. The major constraint is the limited ability of drugs to reach the diseased site, whereby only 0.7% of drug particles can penetrate tumors [110], leading to an insufficient supply of the drug at tumor sites, systemic toxicity, and chemo resistance [111,112,113]. To solve these difficulties, X-ray irradiation, ultrasound, electroporation and magnetic conductivity have been used to deliver drugs to heighten penetration. Nevertheless, these methods can cause inevitable side effects, such as radioactive brain necrosis and low endocrine function. Therefore, it is urgently needed to develop a safe, efficient and targeted drug carrier for clinical treatment. Inspired by the self-propulsion of human cells such as sperm, the self-propelled micro/nano-motors are considered to provide high-efficiency drug delivery for disease treatment. Micro/nano-motors have the characteristics of remote control, high drug loading, autonomous motion and on-demand therapeutic release, which can eliminate the dependence on blood flow and random diffusion, overcome biological barriers to penetrate tissue, and improve the rate of drug target acquisition to minimize the required dose [114,115,116,117]. These advantages of micro/nano-motors promise new prospects for cancer treatment.
The typical example of targeted therapy is hydrogen and chemotherapeutics to synergistically treat the cancer, which are loaded in Mg-based nanobottle motors [39]. Hydrogen is an endogenous gas with physiologic/pathological regulatory functions. Active hydrogen produced by Mg-based nanobottle motors can eliminate overexpressed reactive oxygen species in tumor cells (Figure 10a). CellROX Green reagent is non-fluorescent in reduced status while it combines with DNA and becomes fluorescent after oxidization, which was used to determine the reactive oxygen species level in control group and experimental groups. The results are shown in Figure 10b,c, wherein Mg-based nanobottle motors have the highest reactive oxygen species scavenging ability, which is over 12 times higher than that of passive nanoparticles. Then, they assessed the curative effect of this therapeutic method on colon cancer cell line and breast cancer cell line (Figure 10d,e). When the concentration of Mg–nanobottle-motors–doxorubicin increased from 50 to 100 μg·mL−1, the survival rate of breast cancer cell line decreased sharply from 92.37% to 31.00%, while the viability of Mg-poly (lactic-co-glycolic acid)-doxorubicin decreased slightly from 91.34% to 75.43%. Apart from drug delivery, trapping pathogens is an important therapy. Researchers fabricated an inert parylene nanobottle motor with internal chemoattractant and therapeutic layers, which is capable of self-propulsion through alternating counterclockwise (run) and clockwise rotation (tumble) of flagella [118]. It can attract, trap and destroy pathogens by controlling sequential dissolution and autonomous release of chemical attractants and treatment drugs. The dissolved chemical attractants reduce the propulsion speed, resulting in an increase in the concentration gradient (Figure 10g) [119]. This local concentration distribution leads to chemotactic attraction that causes pathogens to move towards the cavity inlet and converge within the hollow cavity (Figure 10f). The release of fungicides (Ag+) leads to the dissolving of pathogens in the cavity. In contrast, the nanobottle motors consisting of chemical attractants and fungicides have over four times the pathogen killing efficiency (94%) than the microtraps that loaded fungicides alone (20%).
The propulsion mode has been improved via the use of cytotoxic fuel to biocompatible alternative fuels or external field. The motion has been designed from non-directionality and uncontrollable speed to realize precise control of motion. These breakthroughs give nanobottle motors great potential application in biomedicine. Although there has been great progress in biomedicine research by use of nanobottle motors, there are still many challenges from laboratory to clinical treatment. The first point is how to ensure that the motion of nanobottle motors is not affected in the complicated biological environment of the human body. The second is that the biocompatibility and degradability of nanobottle motors need to be improved. Most current research has been conducted at the cellular or animal level, so further studies are needed on circulation metabolism in vivo to ensure the biosafety of nanobottle motors in drug delivery. The third point is that the motion of nanobottle motors in current research is mostly performed by artificial control, so they cannot work independently, and they are vulnerable to environmental forces.

5. Summary and Outlook

Micro/nano-motors are an important research field within nanotechnology, and various geometries have been developed to execute applications. As a new type of motor, the research on nanobottles is still in the initial stage, but it attracts wide attention because of its unique advantages of large cavity, high specific surface area, bionic streamline structure and chemotactic motion. This work summarizes the propulsion mechanisms, fabrication methods and potential applications of nanobottle motors. The propulsion mechanisms include chemical propulsion, light driving and magnetic actuation. Chemical propulsion is divided into bubble propulsion and self-diffusiophoresis. Bubble propulsion generates reverse driving force by the bubble’s separation. The driving force of self-diffusiophoresis and light driving originate from their asymmetric concentration gradient and asymmetric temperature gradient, respectively. Magnetic actuations apply magnetic force or torque to magnetized motors. Three methods including the soft-template-based hydrothermal method and the swelling-inducement and wet-chemistry methods are then summarized for the fabrication of nanobottles. Soft-template-based hydrothermal method is a method of self-assembly between precursor molecules and core templates. Swelling inducement is performed through swelling polymer microspheres with organic solvent and then quenching with ethanol. Wet chemistry tunes the sizes and shapes of nanobottles by controlling reaction dynamics and thermodynamic parameters. Moreover, the potential applications of nanobottle motors in energy, environmental and biomedical fields are additionally discussed in consideration of their structure and motion properties. We constructed a simple table to describe the typical examples of nanobottle motors in three application fields above (Table 2). Overall, nanobottles have irreplaceable advantages for potential applications. However, the current research achievements cannot meet the practical requirements. Here, we list propulsion mechanisms, fabrication methods, potential applications advantages and challenges of nanobottle motors in Table 3. The advantages of a large surface, low density and enhancement of the mass transfer rate give nanobottle motors great photocatalytic performance. The fast motion and strong affinity for organic pollutants of nanobottle motors makes them useful for water cleaning. Nanobottles motors promise new prospects for cancer treatment due to the characteristics of an asymmetrical cavity structure, autonomic motion and high impermeability. Although the research on nanobottle motors has seen huge efforts, much of this field remains to be explored. Future research may aim at the speed increase, fuel consumption, biocompatibility and intelligence of nanobottles.
(1)
In energy applications, motor speed is crucial for photocatalytic efficiency. The current propulsion methods of nanobottles use a single driving source. Multiple-fields propulsion may provide an effective way to increase the speed of nanobottles. The first is choosing one propulsion method (such as light driving) as the main role, and the other one (e.g., magnetic-field pulsion) as an auxiliary means. The other way is to place the nanobottle in the coupling field to achieve better propulsion. In addition, the prices and properties (e.g., light absorption capacity, surface reaction kinetics) of nanobottle motors materials should be comprehensive and researched in depth, so as to obtain fabrication materials with low cost and high solar energy conversion efficiency. On the other hand, machine learning can design, predict and screen materials without experimentation, which can effectively save time and materials costs [122,123,124,125]. Furthermore, it is important to strengthen the research on catalytic efficiency in sunlight to promote the application of nanobottle motors in practical industrial production.
(2)
In terms of environment applications, the hug hurdle is fuel consumption. Improving the motion efficiency as well as precise controlling the structure and performance of nanobottle motors may overcome the problems. Another significant attention is development of eco-friendly nanobottles or design of appropriate recycling methods to avoid secondly pollution in the process of environment monitoring and remediation. The last is the need to be further tested in practical applications (such as the sewage from sewage plant) by using eco-friendly fuels.
(3)
For biomedical applications, toxic chemical fuels are strictly forbidden due to the constraints of the human environment. Combination cell derivatives with biocompatible materials could be a possible way to improve the biocompatibility of nanobottles. Moreover, strengthening the research will be focused on how the blood constituents affect the motion of nanobottle motors in the blood stream, as well as through changing the driving source and designing innovative materials with nanobottle motors to improve the mobility in complex physiological environments. Finally, the intellectualization of multiple nanobottles is expected to be achieved for targeted therapy, imaging and diagnosing the location of the disease in vivo, and cleaning of a tumor or microthrombus.

Author Contributions

Q.L.: writing—original draft. L.W.: resources preparation and editing. K.W.: review and editing. T.W.: supervision and editing. G.L.: conceptualization, supervision and writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of China (Grant No. 52076077; Grant No. 51876059) and the Fundamental Research Funds for the Central Universities (Grant No. 2022JG001; Grant No. 2021MS009).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, Z.; Tu, Y.; Chen, Y.; Peng, F. Emerging micro/nanomotor-based platforms for biomedical therapy. Adv. Intell. Syst. 2020, 2, 1900081. [Google Scholar] [CrossRef]
  2. Safdar, M.; Khan, S.U.; Janis, J. Progress toward catalytic micro-and nanomotors for biomedical and environmental applications. Adv. Mater. 2018, 30, e1703660. [Google Scholar] [CrossRef] [Green Version]
  3. Wang, J.; Xiong, Z.; Liu, M.; Li, X.M.; Zheng, J.; Zhan, X.; Ding, W.; Chen, J.; Li, X.; Li, X.D.; et al. Rational design of reversible redox shuttle for highly efficient light-driven microswimmer. ACS Nano 2020, 14, 3272–3280. [Google Scholar] [CrossRef]
  4. Zhou, H.; Mayorga-Martinez, C.C.; Pane, S.; Zhang, L.; Pumera, M. Magnetically driven micro and nanorobots. Chem. Rev. 2021, 121, 4999–5041. [Google Scholar] [CrossRef] [PubMed]
  5. Park, S.; Finkelman, L.; Yossifon, G. Enhanced cargo loading of electrically powered metallo-dielectric pollen bearing multiple dielectrophoretic traps. J. Colloid Interface Sci. 2021, 588, 611–618. [Google Scholar] [CrossRef]
  6. Venugopalan, P.L.; Esteban-Fernandez de Avila, B.; Pal, M.; Ghosh, A.; Wang, J. Fantastic voyage of nanomotors into the cell. ACS Nano. 2020, 14, 9423–9439. [Google Scholar] [CrossRef] [PubMed]
  7. Soto, F.; Wang, J.; Ahmed, R.; Demirci, U. Medical micro/nanorobots in precision medicine. Adv. Sci. 2020, 7, 2002203. [Google Scholar] [CrossRef]
  8. Xu, D.; Wang, Y.; Liang, C.; You, Y.; Sanchez, S.; Ma, X. Self-propelled micro/nanomotors for on-demand biomedical cargo transportation. Small 2020, 16, e1902464. [Google Scholar] [CrossRef]
  9. Moo, J.G.S.; Mayorga-Martinez, C.C.; Wang, H.; Teo, W.Z.; Tan, B.H.; Luong, T.D.; Gonzalez-Avila, S.R.; Ohl, C.-D.; Pumera, M. Bjerknes forces in motion: Long-range translational motion and chiral directionality switching in bubble-propelled micromotors via an ultrasonic pathway. Adv. Funct. Mater. 2018, 28, 1702618. [Google Scholar] [CrossRef]
  10. Kim, K.; Guo, J.; Liang, Z.; Fan, D. Artificial micro/nanomachines for bioapplications: Biochemical delivery and diagnostic sensing. Adv. Funct. Mater. 2018, 28, 1705867. [Google Scholar] [CrossRef]
  11. Gao, W.; de Avila, B.E.; Zhang, L.; Wang, J. Targeting and isolation of cancer cells using micro/nanomotors. Adv. Drug Deliv. Rev. 2018, 125, 94–101. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, J.; Mou, F.; Wu, Z.; Tang, S.; Xie, H.; You, M.; Liang, X.; Xu, L.; Guan, J. Simple-structured micromotors based on inherent asymmetry in crystalline phases: Design, large-scale preparation, and environmental application. ACS Appl. Mater. Interfaces. 2019, 11, 16639–16646. [Google Scholar] [CrossRef]
  13. Vilela, D.; Parmar, J.; Zeng, Y.; Zhao, Y.; Sánchez, S. Graphene-based microbots for toxic heavy metal removal and recovery from water. Nano Lett. 2016, 16, 2860–2866. [Google Scholar] [CrossRef] [Green Version]
  14. Vasantha Ramachandran, R.; Bhat, R.; Kumar Saini, D.; Ghosh, A. Theragnostic nanomotors: Successes and upcoming challenges. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2021, 13, e1736. [Google Scholar] [CrossRef]
  15. Karshalev, E.; Esteban-Fernandez de Avila, B.; Wang, J. Micromotors for “chemistry-on-the-fly”. J. Am. Chem. Soc. 2018, 140, 3810–3820. [Google Scholar] [CrossRef]
  16. Choi, H.; Hwang, B.W.; Park, K.M.; Kim, K.S.; Hahn, S.K. Degradable nanomotors using platinum deposited complex of calcium carbonate and hyaluronate nanogels for targeted drug delivery. Part Part Syst Charact. 2019, 37, 1900418. [Google Scholar] [CrossRef]
  17. Zhao, F.; Rong, W.; Wang, L.; Sun, L. Magnetic actuated shape-memory helical microswimmers with programmable pecovery behaviors. J. Bionic. Eng. 2021, 18, 799–811. [Google Scholar] [CrossRef]
  18. Lu, X.; Shen, H.; Wei, Y.; Ge, H.; Wang, J.; Peng, H.; Liu, W. Ultrafast growth and locomotion of dandelion-like microswarms with tubular micromotors. Small 2020, 16, e2003678. [Google Scholar] [CrossRef] [PubMed]
  19. Ji, Y.; Lin, X.; Zhang, H.; Wu, Y.; Li, J.; He, Q. Thermoresponsive polymer brush modulation on the direction of motion of phoretically driven Janus micromotors. Angew. Chem. Int. Ed. Engl. 2019, 58, 4184–4188. [Google Scholar] [CrossRef]
  20. Liang, Z.; Teal, D.; Fan, D.E. Light programmable micro/nanomotors with optically tunable in-phase electric polarization. Nat. Commun. 2019, 10, 5275. [Google Scholar] [CrossRef]
  21. Yao, J.; Ma, Y.; Liu, J.; Liu, S.; Pan, J. Janus-like boronate affinity magnetic molecularly imprinted nanobottles for specific adsorption and fast separation of luteolin. Chem. Eng. J. 2019, 356, 436–444. [Google Scholar] [CrossRef]
  22. Dai, J.; Cheng, X.; Li, X.; Wang, Z.; Wang, Y.; Zheng, J.; Liu, J.; Chen, J.; Wu, C.; Tang, J. Solution-synthesized multifunctional Janus nanotree microswimmer. Adv. Funct. Mater. 2021, 31, 2106204. [Google Scholar] [CrossRef]
  23. Xiong, F.; Han, Y.; Wang, S.; Li, G.; Qin, T.; Chen, Y.; Chu, F. Preparation and formation mechanism of renewable lignin hollow nanospheres with a single hole by self-assembly. ACS Sustain. Chem. Eng. 2017, 5, 2273–2281. [Google Scholar] [CrossRef]
  24. Zhou, C.; Gao, C.; Wu, Y.; Si, T.; Yang, M.; He, Q. Torque-driven orientation motion of chemotactic colloidal motors. Angew. Chem. Int. Ed. Engl. 2022, 61, e202116013. [Google Scholar]
  25. Lou, Z.; Wang, Y.; Yang, Y.; Wang, Y.; Qin, C.; Liang, R.; Chen, X.; Ye, Z.; Zhu, L. Carbon sphere template derived hollow nanostructure for photocatalysis and gas sensing. Nanomaterials 2020, 10, 378. [Google Scholar] [CrossRef] [Green Version]
  26. Qiu, J.; Huo, D.; Xia, Y. Phase-change materials for controlled release and related applications. Adv. Mater. 2020, 32, e2000660. [Google Scholar] [CrossRef]
  27. Ou, J.; Liu, K.; Jiang, J.; Wilson, D.A.; Liu, L.; Wang, F.; Wang, S.; Tu, Y.; Peng, F. Micro-/nanomotors toward biomedical applications: The recent progress in biocompatibility. Small 2020, 16, e1906184. [Google Scholar] [CrossRef]
  28. Zhang, G.; Yu, Y.; Chen, X.; Han, Y.; Di, Y.; Yang, B.; Xiao, F.; Shen, J. Silica nanobottles templated from functional polymer spheres. J. Colloid Interf Sci. 2003, 263, 467–472. [Google Scholar] [CrossRef]
  29. Hyun, D.C.; Lu, P.; Choi, S.I.; Jeong, U.; Xia, Y. Microscale polymer bottles corked with a phase-change material for temperature-controlled release. Angew. Chem. Int. Ed. Engl. 2013, 52, 10468–10471. [Google Scholar] [CrossRef] [Green Version]
  30. Qiu, J.; Xu, J.; Xia, Y. Nanobottles for controlled release and drug delivery. Adv. Healthc. Mater. 2021, 10, e2000587. [Google Scholar] [CrossRef]
  31. Qiu, J.; Camargo, P.H.C.; Jeong, U.; Xia, Y. Synthesis, transformation, and utilization of monodispersed colloidal spheres. Acc. Chem. Res. 2019, 52, 3475–3487. [Google Scholar] [CrossRef]
  32. Si, Y.; Chen, M.; Wu, L. Syntheses and biomedical applications of hollow micro-/nano-spheres with large-through-holes. Chem. Soc. Rev. 2016, 45, 690–714. [Google Scholar] [CrossRef]
  33. Zou, H.; Shang, K. Synthetic strategies for hollow particles with open holes on their surfaces. Mater. Chem. Front. 2021, 5, 3765–3787. [Google Scholar] [CrossRef]
  34. Liu, M.; Zhao, K. Engineering active micro and nanomotors. Micromachines 2021, 12, 687. [Google Scholar] [CrossRef]
  35. Wrede, P.; Medina-Sanchez, M.; Fomin, V.M.; Schmidt, O.G. Switching propulsion mechanisms of tubular catalytic micromotors. Small 2021, 17, e2006449. [Google Scholar] [CrossRef]
  36. Shelke, Y.; Srinivasan, N.R.; Thampi, S.P.; Mani, E. Transition from linear to circular motion in active spherical-cap colloids. Langmuir 2019, 35, 4718–4725. [Google Scholar] [CrossRef]
  37. Kong, L.; Mayorga-Martinez, C.C.; Guan, J.; Pumera, M. Photocatalytic micromotors activated by UV to visible light for environmental remediation, micropumps, reversible assembly, transportation, and biomimicry. Small 2020, 16, e1903179. [Google Scholar] [CrossRef]
  38. Suematsu, N.J.; Nakata, S. Evolution of self-propelled objects: From the viewpoint of nonlinear science. Chem. Eur. J. 2018, 24, 6308–6324. [Google Scholar] [CrossRef]
  39. Moran, J.L.; Posner, J.D. Phoretic self-propulsion. Annu. Rev. Fluid. Mech. 2017, 49, 511–540. [Google Scholar] [CrossRef]
  40. Wang, Y.; Tu, Y.; Peng, F. The energy conversion behind micro-and nanomotors. Micromachines 2021, 12, 222. [Google Scholar] [CrossRef]
  41. Zhou, C.; Zhang, H.P.; Tang, J.; Wang, W. Photochemically powered AgCl Janus micromotors as a model system to understand Ionic self-diffusiophoresis. Langmuir 2018, 34, 3289–3295. [Google Scholar] [CrossRef]
  42. Gao, C.; Zhou, C.; Lin, Z.; Yang, M.; He, Q. Surface wettability-directed propulsion of glucose-powered nanoflask motors. ACS Nano. 2019, 13, 12758–12766. [Google Scholar] [CrossRef] [PubMed]
  43. Hayakawa, M.; Onoe, H.; Nagai, K.H.; Takinoue, M. Influence of asymmetry and driving forces on the propulsion of bubble-propelled catalytic micromotors. Micromachines 2016, 7, 229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Wu, Y.; Song, Z.; Deng, G.; Jiang, K.; Wang, H.; Zhang, X.; Han, H. Gastric acid powered nanomotors release antibiotics for in vivo treatment of helicobacter pylori infection. Small 2021, 17, e2006877. [Google Scholar] [CrossRef] [PubMed]
  45. Nourhani, A.; Karshalev, E.; Soto, F.; Wang, J. Multigear bubble propulsion of transient micromotors. Research 2020, 2020, 7823615. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Li, L.; Chen, B.; Guo, M.; Yang, Q.; Zhang, Y.; Zhang, M. Platinum Janus nanoparticles as peroxidase mimics for catalytic immunosorbent assay. ACS Appl. Nano Mater. 2022, 5, 1397–1407. [Google Scholar] [CrossRef]
  47. Xu, L.; Mou, F.; Gong, H.; Luo, M.; Guan, J. Light-driven micro/nanomotors: From fundamentals to applications. Chem. Soc. Rev. 2017, 46, 6905–6926. [Google Scholar] [CrossRef] [PubMed]
  48. Zhan, X.; Zheng, J.; Zhao, Y.; Zhu, B.; Cheng, R.; Wang, J.; Liu, J.; Tang, J.; Tang, J. From strong dichroic nanomotor to polarotactic microswimmer. Adv. Mater. 2019, 31, e1903329. [Google Scholar] [CrossRef]
  49. Eskandarloo, H.; Kierulf, A.; Abbaspourrad, A. Light-harvesting synthetic nano- and micromotors: A review. Nanoscale 2017, 9, 12218–12230. [Google Scholar] [CrossRef]
  50. Dong, R.; Cai, Y.; Yang, Y.; Gao, W.; Ren, B. Photocatalytic micro/nanomotors: From construction to applications. Acc. Chem. Res. 2018, 51, 1940–1947. [Google Scholar] [CrossRef] [Green Version]
  51. Li, Z.; Wang, D.; Gao, H.; Cao, H.; Zhao, Y.; Miao, Z.; Yang, Z.; He, W. Application of double click to prepare near-infrared absorbing dye for photo-thermal tuning of cholesteric liquid crystal. React. Funct. Polym. 2020, 151, 104549. [Google Scholar] [CrossRef]
  52. Yaguchi, M.; Jia, X.; Schlesinger, R.; Jiang, X.; Ataka, K.; Heberle, J. Near-infrared activation of sensory rhodopsin II mediated by NIR-to-blue upconversion nanoparticles. Front. Mol. Biosci. 2021, 8, 782688. [Google Scholar] [CrossRef]
  53. Tang, Y.; Wang, G. NIR light-responsive nanocarriers for controlled release. J. Photoch. Photobio. C. 2021, 47, 100420. [Google Scholar] [CrossRef]
  54. Xuan, M.; Mestre, R.; Gao, C.; Zhou, C.; He, Q.; Sanchez, S. Noncontinuous super-diffusive dynamics of a light-activated nanobottle motor. Angew. Chem. Int. Ed. Engl. 2018, 57, 6838–6842. [Google Scholar] [CrossRef]
  55. Wu, Z.; Si, T.; Gao, W.; Lin, X.; Wang, J.; He, Q. Superfast near-infrared light-driven polymer multilayer rockets. Small 2016, 12, 577–582. [Google Scholar] [CrossRef]
  56. Ge, M.; Xu, D.; Chen, Z.; Wei, C.; Zhang, Y.; Yang, C.; Chen, Y.; Lin, H.; Shi, J. Magnetostrictive-piezoelectric-triggered nanocatalytic tumor therapy. Nano Lett. 2021, 21, 6764–6772. [Google Scholar] [CrossRef]
  57. Peyer, K.E.; Zhang, L.; Nelson, B.J. Bio-inspired magnetic swimming microrobots for biomedical applications. Nanoscale 2013, 5, 1259–1272. [Google Scholar] [CrossRef]
  58. Wang, B.; Kostarelos, K.; Nelson, B.J.; Zhang, L. Trends in micro-/nanorobotics: Materials development, actuation, localization, and system integration for biomedical applications. Adv. Mater. 2021, 33, e2002047. [Google Scholar] [CrossRef]
  59. Fernández-Medina, M.; Ramos-Docampo, M.A.; Hovorka, O.; Salgueiriño, V.; Städler, B. Recent advances in nano-and micromotors. Adv. Funct. Mater. 2020, 30, 1908283. [Google Scholar] [CrossRef]
  60. Li, Y.; Shi, Q.; Luo, Y.; Chu, G.; Zou, H.; Zhang, L.; Sun, B. Hydrothermal controllable synthesis of hollow carbon particles: Reaction-growth mechanism. Chem. Eng. Sci. 2020, 225, 115787. [Google Scholar] [CrossRef]
  61. Chen, C.; Wang, H.; Han, C.; Deng, J.; Wang, J.; Li, M.; Tang, M.; Jin, H.; Wang, Y. Asymmetric flasklike hollow carbonaceous nanoparticles fabricated by the synergistic interaction between soft Tetmplate and biomass. J. Am. Chem. Soc. 2017, 139, 2657–2663. [Google Scholar] [CrossRef]
  62. Yi, D.; Zhang, Q.; Liu, Y.; Song, J.; Tang, Y.; Caruso, F.; Wang, Y. Synthesis of chemically asymmetric silica nanobottles and their application for cargo loading and as nanoreactors and nanomotors. Angew. Chem. Int. Ed. Engl. 2016, 55, 14733–14737. [Google Scholar] [CrossRef] [Green Version]
  63. Zheng, L.; Chen, Q.; Chen, B.; Lin, J. Swelling synthesis and modification of Janus composite particles containing natural urushiol. Mater. Lett. 2014, 120, 271–274. [Google Scholar] [CrossRef]
  64. Chen, S.R.; Bobrin, V.A.; Jia, Z.; Monteiro, M.J. Temperature-directed formation of anisotropic kettlebell and tadpole nanostructures in the absence of a swelling-induced solvent. Angew. Chem. Int. Ed. Engl. 2022, 61, e202113974. [Google Scholar]
  65. Liu, Y.; Wang, J.; Shao, Y.; Deng, R.; Zhu, J.; Yang, Z. Recent advances in scalable synthesis and performance of Janus polymer/inorganic nanocomposites. Prog. Mater. Sci. 2022, 124, 100888. [Google Scholar] [CrossRef]
  66. Qiu, J.; Shi, Y.; Xia, Y. Polydopamine nanobottles with photothermal capability for controlled release and related applications. Adv. Mater. 2021, 33, e2104729. [Google Scholar] [CrossRef]
  67. Qiu, J.; Chen, Z.; Chi, M.; Xia, Y. Swelling-induced symmetry breaking: A versatile approach to the scalable production of colloidal particles with a Janus structure. Angew. Chem. Int. Ed. Engl. 2021, 60, 12980–12984. [Google Scholar] [CrossRef]
  68. Xu, L.; Liu, D.; Chen, D.; Liu, H.; Yang, J. Size and shape controlled synthesis of rhodium nanoparticles. Heliyon 2019, 5, e01165. [Google Scholar] [CrossRef] [Green Version]
  69. Zhang, H.; Chen, J.; Li, N.; Jiang, R.; Zhu, X.M.; Wang, J. Au nanobottles with synthetically tunable overall and opening sizes for chemo-photothermal combined therapy. ACS Appl. Mater. Interfaces. 2019, 11, 5353–5363. [Google Scholar] [CrossRef]
  70. Jiang, R.; Qin, F.; Liu, Y.; Ling, X.Y.; Guo, J.; Tang, M.; Cheng, S.; Wang, J. Colloidal gold nanocups with orientation-dependent plasmonic properties. Adv. Mater. 2016, 28, 6322–6331. [Google Scholar] [CrossRef]
  71. Jeong, S.; Song, J.; Lee, S. Photoelectrochemical device designs toward practical solar water splitting: A review on the recent progress of BiVO4 and BiFeO3 photoanodes. Appl. Sci. 2018, 8, 1388. [Google Scholar] [CrossRef] [Green Version]
  72. Zhang, F.Q.; Hu, Y.; Sun, R.N.; Fu, H.; Peng, K.Q. Gold-sensitized silicon/ZnO core/shell nanowire array for solar water splitting. Front. Chem. 2019, 7, 206. [Google Scholar] [CrossRef] [PubMed]
  73. Li, Y.; Tsang, S.C.E. Recent progress and strategies for enhancing photocatalytic water splitting. Mater. Today Sustain. 2020, 9, 100032. [Google Scholar] [CrossRef]
  74. Wang, Z.; Li, C.; Domen, K. Recent developments in heterogeneous photocatalysts for solar-driven overall water splitting. Chem. Soc. Rev. 2019, 48, 2109–2125. [Google Scholar] [CrossRef] [PubMed]
  75. Humayun, M.; Ullah, H.; Usman, M.; Habibi-Yangjeh, A.; Tahir, A.A.; Wang, C.; Luo, W. Perovskite-type lanthanum ferrite based photocatalysts: Preparation, properties, and applications. J. Energy Chem. 2022, 66, 314–338. [Google Scholar] [CrossRef]
  76. Wang, W.; Huang, X.; Lai, M.; Lu, C. RGO/TiO2 nanosheets immobilized on magnetically actuated artificial cilia film: A new mode for efficient photocatalytic reaction. RSC Adv. 2017, 7, 10517–10523. [Google Scholar] [CrossRef] [Green Version]
  77. Peng, F.; Zhou, Q.; Lu, C.; Ni, Y.; Kou, J.; Xu, Z. Construction of (001) facets exposed ZnO nanosheets on magnetically driven cilia film for highly active photocatalysis. Appl. Surf. Sci. 2017, 394, 115–124. [Google Scholar] [CrossRef]
  78. Peng, F.; Ni, Y.; Zhou, Q.; Lu, C.; Kou, J.; Xu, Z. Design of inner-motile ZnO@TiO2 mushroom arrays on magnetic cilia film with enhanced photocatalytic performance. J. Photoch. Photobio. A 2017, 332, 150–157. [Google Scholar] [CrossRef]
  79. Wu, Y.; Dong, R.; Zhang, Q.; Ren, B. Dye-enhanced self-electrophoretic propulsion of light-driven TiO2-Au Janus micromotors. Nanomicro Lett. 2017, 9, 30. [Google Scholar] [CrossRef] [Green Version]
  80. Dong, Y.; Yi, C.; Yang, S.; Wang, J.; Chen, P.; Liu, X.; Du, W.; Wang, S.; Liu, B.F. A substrate-free graphene oxide-based micromotor for rapid adsorption of antibiotics. Nanoscale 2019, 11, 4562–4570. [Google Scholar] [CrossRef]
  81. Mou, F.; Kong, L.; Chen, C.; Chen, Z.; Xu, L.; Guan, J. Light-controlled propulsion, aggregation and separation of water-fuelled TiO2/Pt Janus submicromotors and their “on-the-fly” photocatalytic activities. Nanoscale 2016, 8, 4976–4983. [Google Scholar] [CrossRef]
  82. Dong, R.; Hu, Y.; Wu, Y.; Gao, W.; Ren, B.; Wang, Q.; Cai, Y. Visible-light-driven BiOI-based Janus micromotor in pure water. J. Am. Chem. Soc. 2017, 139, 1722–1725. [Google Scholar] [CrossRef]
  83. Dong, T.; Sun, M.; Hu, K.; Wang, Q.; Lu, C.; Kou, J. Self-propelled jet carbon micromotor enhanced photocatalytic performance for water splitting. Int. J. Hydrogen. Energ. 2021, 46, 17187–17196. [Google Scholar] [CrossRef]
  84. Yan, M.; Ma, D.; Qiu, B.; Liu, T.; Xie, L.; Zeng, J.; Liang, K.; Xin, H.; Lian, Z.; Jiang, L.; et al. Superassembled hierarchical asymmetric magnetic mesoporous nanorobots driven by smart confined catalytic degradation. Chem. Eur. J. 2022, 28, e202200307. [Google Scholar] [CrossRef]
  85. Zhou, C.; Gao, C.; Lin, Z.; Wang, D.; Li, Y.; Yuan, Y.; Zhu, B.; He, Q. Autonomous motion of bubble-powered carbonaceous nanoflask motors. Langmuir 2020, 36, 7039–7045. [Google Scholar] [CrossRef]
  86. Kumar, L.; Ragunathan, V.; Chugh, M.; Bharadvaja, N. Nanomaterials for remediation of contaminants: A review. Environ Chem Lett. 2021, 19, 3139–3163. [Google Scholar] [CrossRef]
  87. Egbueri, J.C.; Unigwe, C.O. Understanding the extent of heavy metal pollution in drinking water supplies from umunya, nigeria: An indexical and statistical assessment. Anal. Lett. 2020, 53, 2122–2144. [Google Scholar] [CrossRef]
  88. Saravanan, A.; Senthil Kumar, P.; Jeevanantham, S.; Karishma, S.; Tajsabreen, B.; Yaashikaa, P.R.; Reshma, B. Effective water/wastewater treatment methodologies for toxic pollutants removal: Processes and applications towards sustainable development. Chemosphere 2021, 280, 130595. [Google Scholar] [CrossRef]
  89. Worthington, M.J.H.; Shearer, C.J.; Esdaile, L.J.; Campbell, J.A.; Gibson, C.T.; Legg, S.K.; Yin, Y.; Lundquist, N.A.; Gascooke, J.R.; Albuquerque, I.S.; et al. Sustainable polysulfides for oil spill remediation: Repurposing industrial waste for environmental benefit. Adv. Sustain. Syst. 2018, 2, 1800024. [Google Scholar] [CrossRef] [Green Version]
  90. Ji, M.; Liu, Z.; Sun, K.; Li, Z.; Fan, X.; Li, Q. Bacteriophages in water pollution control: Advantages and limitations. Front. Environ. Sci. Eng. 2021, 15, 84. [Google Scholar] [CrossRef]
  91. Bhatt, P.; Pathak, V.M.; Bagheri, A.R.; Bilal, M. Microplastic contaminants in the aqueous environment, fate, toxicity consequences, and remediation strategies. Environ. Res. 2021, 200, 111762. [Google Scholar] [CrossRef]
  92. Verma, B.; Balomajumder, C. Surface modification of one-dimensional carbon nanotubes: A review for the management of heavy metals in wastewater. Environ. Technol. Innov. 2020, 17, 100596. [Google Scholar] [CrossRef]
  93. Khin, M.M.; Nair, A.S.; Babu, V.J.; Murugan, R.; Ramakrishna, S. A review on nanomaterials for environmental remediation. Energ. Environ. Sci. 2012, 5, 8075. [Google Scholar] [CrossRef]
  94. Verma, B.; Gumfekar, S.P.; Sabapathy, M. A critical review on micro-and nanomotors: Application towards wastewater treatment. Can. J. Chem. Eng. 2021, 100, 540–558. [Google Scholar] [CrossRef]
  95. Peng, H.; Guo, J. Removal of chromium from wastewater by membrane filtration, chemical precipitation, ion exchange, adsorption electrocoagulation, electrochemical reduction, electrodialysis, electrodeionization, photocatalysis and nanotechnology: A review. Environ. Chem Lett. 2020, 18, 2055–2068. [Google Scholar] [CrossRef]
  96. Soler, L.; Sanchez, S. Catalytic nanomotors for environmental monitoring and water remediation. Nanoscale 2014, 6, 7175–7182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Sun, M.; Chen, W.; Fan, X.; Tian, C.; Sun, L.; Xie, H. Cooperative recyclable magnetic microsubmarines for oil and microplastics removal from water. Appl. Mater. Today 2020, 20, 100682. [Google Scholar] [CrossRef]
  98. Taha, M.R.; Mobasser, S. Adsorption of DDT and PCB by nanomaterials from residual soil. PLoS ONE. 2015, 10, e0144071. [Google Scholar] [CrossRef]
  99. Taha, M.R.; Mobasser, S. Adsorption of DDT from contaminated soil using carbon nanotubes. Soil Sediment Contam. 2014, 23, 703–714. [Google Scholar] [CrossRef]
  100. Yuan, K.; Asunción-Nadal, V.d.l.; Li, Y.; Jurado-Sánchez, B.; Escarpa, A. Graphdiyne tubular micromotors: Electrosynthesis, characterization and self-propelled capabilities. Appl. Mater. Today 2020, 20, 100743. [Google Scholar] [CrossRef]
  101. Safdar, M.; Simmchen, J.; Jänis, J. Light-driven micro-and nanomotors for environmental remediation. Environ Sci. Nano. 2017, 4, 1602–1616. [Google Scholar] [CrossRef]
  102. Ying, Y.; Pumera, M. Micro/nanomotors for water purification. Chem. Eur. J. 2019, 25, 106–121. [Google Scholar] [CrossRef]
  103. Ye, H.; Wang, Y.; Xu, D.; Liu, X.; Liu, S.; Ma, X. Design and fabrication of micro/nano-motors for environmental and sensing applications. Appl. Mater. Today 2021, 23, 101007. [Google Scholar] [CrossRef]
  104. Qiu, B.; Xie, L.; Zeng, J.; Liu, T.; Yan, M.; Zhou, S.; Liang, Q.; Tang, J.; Liang, K.; Kong, B. Interfacially super-assembled asymmetric and H2O2 sensitive multilayer-sandwich nagnetic nesoporous silica nanomotors for detecting and Removing heavy metal ions. Adv. Funct. Mater. 2021, 31, 2010694. [Google Scholar] [CrossRef]
  105. Mou, F.; Pan, D.; Chen, C.; Gao, Y.; Xu, L.; Guan, J. Magnetically modulated pot-like MnFe2O4 micromotors: Nanoparticle assembly fabrication and their capability for direct oil removal. Adv. Funct. Mater. 2015, 25, 6173–6181. [Google Scholar] [CrossRef]
  106. Lv, J.; Xing, Y.; Xu, T.; Zhang, X.; Du, X. Advanced micro/nanomotors for enhanced bioadhesion and tissue penetration. Appl. Mater. Today 2021, 23, 101034. [Google Scholar] [CrossRef]
  107. Edis, Z.; Wang, J.; Waqas, M.K.; Ijaz, M.; Ijaz, M. Nanocarriers-mediated drug delivery systems for anticancer agents: An Overview and perspectives. Int. J. Nanomed. 2021, 16, 1313–1330. [Google Scholar] [CrossRef]
  108. Saeki, H.; Sohda, M.; Sakai, M.; Sano, A.; Shirabe, K. Role of surgery in multidisciplinary treatment strategies for locally advanced esophageal squamous cell carcinoma. Ann. Gastroenterol. Surg. 2020, 4, 490–497. [Google Scholar] [CrossRef]
  109. Wilhelm, S.; Tavares, A.J.; Dai, Q.; Ohta, S.; Audet, J.; Dvorak, H.F.; Chan, W.C.W. Analysis of nanoparticle delivery to tumours. Nat. Rev. Mater. 2016, 1, 16014. [Google Scholar] [CrossRef]
  110. Ma, Z.; Jia, X.; Bai, J.; Ruan, Y.; Wang, C.; Li, J.; Zhang, M.; Jiang, X. MnO2Gatekeeper: An intelligent and O2-evolving shell for preventing premature release of high cargo payload core, overcoming tumor hypoxia, and acidic H2O2-sensitive MRI. Adv. Funct. Mater. 2017, 27, 1604258. [Google Scholar] [CrossRef]
  111. Rizvi, S.A.A.; Saleh, A.M. Applications of nanoparticle systems in drug delivery technology. Saudi Pharm. J. 2018, 26, 64–70. [Google Scholar] [CrossRef] [PubMed]
  112. Martinelli, C.; Pucci, C.; Ciofani, G. Nanostructured carriers as innovative tools for cancer diagnosis and therapy. APL Bioeng. 2019, 3, 011502. [Google Scholar] [CrossRef] [PubMed]
  113. Lin, R.; Yu, W.; Chen, X.; Gao, H. Self-propelled micro/nanomotors for tumor targeting delivery and terapy. Adv. Healthc. Mater. 2021, 10, e2001212. [Google Scholar]
  114. Agrahari, V.; Agrahari, V.; Chou, M.L.; Chew, C.H.; Noll, J.; Burnouf, T. Intelligent micro-/nanorobots as drug and cell carrier devices for biomedical therapeutic advancement: Promising development opportunities and translational challenges. Biomaterials 2020, 260, 120163. [Google Scholar] [CrossRef]
  115. Reinisova, L.; Hermanova, S.; Pumera, M. Micro/nanomachines: What is needed for them to become a real force in cancer therapy? Nanoscale 2019, 11, 6519–6532. [Google Scholar] [CrossRef]
  116. Lim, S.-h.; Sohn, S.W.; Lee, H.; Choi, D.; Jang, E.; Kim, M.; Lee, J.; Park, S. Analysis and evaluation of path planning algorithms for autonomous driving of electromagnetically actuated microrobot. Int. J. Control. Autom. 2020, 18, 2943–2954. [Google Scholar] [CrossRef]
  117. Liu, K.; Ou, J.; Wang, S.; Gao, J.; Liu, L.; Ye, Y.; Wilson, D.A.; Hu, Y.; Peng, F.; Tu, Y. Magnesium-based micromotors for enhanced active and synergistic hydrogen chemotherapy. Appl. Mater. Today 2020, 20, 100694. [Google Scholar] [CrossRef]
  118. Soto, F.; Kupor, D.; Lopez-Ramirez, M.A.; Wei, F.; Karshalev, E.; Tang, S.; Tehrani, F.; Wang, J. Onion-like multifunctional microtrap vehicles for attraction-trapping-destruction of biological threats. Angew. Chem. Int. Ed. Engl. 2020, 59, 3480–3485. [Google Scholar] [CrossRef]
  119. Rahman, M.A.; Cheng, J.; Wang, Z.; Ohta, A.T. Cooperative micromanipulation using the independent actuation of fifty microrobots in parallel. Sci. Rep. 2017, 7, 3278. [Google Scholar] [CrossRef] [Green Version]
  120. Mahmood, A.; Irfan, A.; Wang, J.-L. Machine learning and molecular dynamics simulation-assisted evolutionary design and discovery pipeline to screen efficient small molecule acceptors for PTB7-Th-based organic solar cells with over 15% efficiency. J. Mater. Chem. A. 2022, 10, 4170–4180. [Google Scholar] [CrossRef]
  121. Mahmood, A.; Irfan, A.; Wang, J.-L. Machine learning for organic photovoltaic polymers: A minireview. Chin. J Polym Sci. 2022, 40, 870–876. [Google Scholar] [CrossRef]
  122. Mahmood, A.; Wang, J.-L. A time and resource efficient machine learning assisted design of non-fullerene small molecule acceptors for P3HT-based organic solar cells and green solvent selection. J. Mater. Chem. A. 2021, 9, 15684–15695. [Google Scholar] [CrossRef]
  123. Mahmood, A.; Irfan, A.; Wang, J.L. Developing efficient small molecule acceptors with sp(2) -hybridized nitrogen at different positions by density functional theory calculations, molecular dynamics simulations and machine learning. Chem. 2022, 28, e202103712. [Google Scholar] [CrossRef]
  124. Wei, X.; Beltran-Gastelum, M.; Karshalev, E.; Esteban-Fernandez de Avila, B.; Zhou, J.; Ran, D.; Angsantikul, P.; Fang, R.H.; Wang, J.; Zhang, L. Biomimetic micromotor enables active delivery of antigens for oral vaccination. Nano Lett. 2019, 19, 1914–1921. [Google Scholar] [CrossRef]
  125. Qiu, J.; Huo, D.; Xue, J.; Zhu, G.; Liu, H.; Xia, Y. Encapsulation of a phase-change material in nanocapsules with a well-defined hole in the wall for the controlled release of drugs. Angew. Chem. Int. Engl. 2019, 58, 10606–10611. [Google Scholar] [CrossRef]
Figure 1. Nanobottles and the propulsion mechanisms, fabrication methods and potential applications [34,35,36,37,38,39,40,41,42,43]. Reprinted with permission from Ref. [34]. Copyright 2019, copyright American Chemical Society. Reprinted with permission from Ref. [35]. Copyright 2018, copyright Wiley-VCH GmbH. Reprinted with permission from Ref. [36]. Copyright 2017, copyright American Chemical Society. Reprinted with permission from Ref. [37]. Copyright 2021, copyright Wiley-VCH GmbH. Reprinted with permission from Ref. [38]. Copyright 2019, copyright American Chemical Society. Reprinted with permission from Ref. [39]. Copyright 2021, copyright Elsevier Ltd. Reprinted with permission from Ref. [40]. Copyright 2022, copyright Wiley-VCH GmbH. Reprinted with permission from Ref. [41]. Copyright 2020, copyright American Chemical Society. Reprinted with permission from Ref. [42]. Copyright 2015, copyright Wiley-VCH GmbH. Reprinted with permission from Ref. [43]. Copyright 2020, copyright Elsevier Ltd.
Figure 1. Nanobottles and the propulsion mechanisms, fabrication methods and potential applications [34,35,36,37,38,39,40,41,42,43]. Reprinted with permission from Ref. [34]. Copyright 2019, copyright American Chemical Society. Reprinted with permission from Ref. [35]. Copyright 2018, copyright Wiley-VCH GmbH. Reprinted with permission from Ref. [36]. Copyright 2017, copyright American Chemical Society. Reprinted with permission from Ref. [37]. Copyright 2021, copyright Wiley-VCH GmbH. Reprinted with permission from Ref. [38]. Copyright 2019, copyright American Chemical Society. Reprinted with permission from Ref. [39]. Copyright 2021, copyright Elsevier Ltd. Reprinted with permission from Ref. [40]. Copyright 2022, copyright Wiley-VCH GmbH. Reprinted with permission from Ref. [41]. Copyright 2020, copyright American Chemical Society. Reprinted with permission from Ref. [42]. Copyright 2015, copyright Wiley-VCH GmbH. Reprinted with permission from Ref. [43]. Copyright 2020, copyright Elsevier Ltd.
Energies 15 07636 g001
Figure 2. (a) The concentration distribution of products caused by catalytic reaction. (b) The driving force (Fd) and friction (Ff) of hydrophilic carbonaceous nanobottle motor and hydrophobic carbonaceous nanobottle motor exerted by the chemical gradient. (c) Puller-like flow field around the hydrophilic carbonaceous nanobottle motor. (d) Pusher-like field around the hydrophobic carbonaceous nanobottle motor, where the red arrow represent the flow velocity and green arrow express motion direction. Reprinted with permission from Ref. [35]. Copyright 2019, copyright American Chemical Society. (e) Bubble generation inside the SiO2 nanobottle motor. (f) Photos of nanobottle moving in H2O2 solution. The scale bar is 20 μm. Reprinted with permission from Ref. [52]. Copyright 2021, copyright Wiley-VCH GmbH.
Figure 2. (a) The concentration distribution of products caused by catalytic reaction. (b) The driving force (Fd) and friction (Ff) of hydrophilic carbonaceous nanobottle motor and hydrophobic carbonaceous nanobottle motor exerted by the chemical gradient. (c) Puller-like flow field around the hydrophilic carbonaceous nanobottle motor. (d) Pusher-like field around the hydrophobic carbonaceous nanobottle motor, where the red arrow represent the flow velocity and green arrow express motion direction. Reprinted with permission from Ref. [35]. Copyright 2019, copyright American Chemical Society. (e) Bubble generation inside the SiO2 nanobottle motor. (f) Photos of nanobottle moving in H2O2 solution. The scale bar is 20 μm. Reprinted with permission from Ref. [52]. Copyright 2021, copyright Wiley-VCH GmbH.
Energies 15 07636 g002
Figure 3. (a) The temperature gradient around carbonaceous nanobottle. (b) The fluid velocity distribution of heated fluid in nanobottle. (c) Pressure distribution caused by the change in internal pressure. At (d) 0 Wcm−2 and (e) 1.36 Wcm−2, the trajectories diagram of carbonaceous nanobottle motor in near-infrared light are shown. (f) The image of carbonaceous nanobottle motor motion, through the switching on/off of the near-infrared light, white arrows are the direction of directional motion. The scale bar is 10 μm. Reprinted with permission from Ref. [36]. Copyright 2018, copyright Wiley-VCH GmbH.
Figure 3. (a) The temperature gradient around carbonaceous nanobottle. (b) The fluid velocity distribution of heated fluid in nanobottle. (c) Pressure distribution caused by the change in internal pressure. At (d) 0 Wcm−2 and (e) 1.36 Wcm−2, the trajectories diagram of carbonaceous nanobottle motor in near-infrared light are shown. (f) The image of carbonaceous nanobottle motor motion, through the switching on/off of the near-infrared light, white arrows are the direction of directional motion. The scale bar is 10 μm. Reprinted with permission from Ref. [36]. Copyright 2018, copyright Wiley-VCH GmbH.
Energies 15 07636 g003
Figure 4. (a) The directional motion trajectories motors under the action of magnets. (bd) The motion direction of nanobottle motors is consistent with the magnets. The scale bar is 20 μm. Reprinted with permission from Ref. [43]. Copyright 2022, copyright Wiley-VCH GmbH.
Figure 4. (a) The directional motion trajectories motors under the action of magnets. (bd) The motion direction of nanobottle motors is consistent with the magnets. The scale bar is 20 μm. Reprinted with permission from Ref. [43]. Copyright 2022, copyright Wiley-VCH GmbH.
Energies 15 07636 g004
Figure 5. (a) Fabrication process of carbon nanobottle motor. (b) SEM image and (c) TEM image of Carbon nanobottles. The scale bar is 500 nm. Reprinted with permission from Ref. [42]. Copyright 2017, copyright American Chemical Society.
Figure 5. (a) Fabrication process of carbon nanobottle motor. (b) SEM image and (c) TEM image of Carbon nanobottles. The scale bar is 500 nm. Reprinted with permission from Ref. [42]. Copyright 2017, copyright American Chemical Society.
Energies 15 07636 g005
Figure 6. (a) The fabrication steps of nanobottle. (b) TEM image of polystyrene particles coated with SiO2. (ce) TEM images of Janus structure obtained through swelling with 0, 20, 30 and 50% tetrahydrofuran/water mixture. (f) SiO2 nanobottles obtained by dissolution of polystyrene particles with pure tetrahydrofuran. The scale bar is 500 nm. Reprinted with permission from Ref. [73]. Copyright 2021, copyright Wiley-VCH GmbH.
Figure 6. (a) The fabrication steps of nanobottle. (b) TEM image of polystyrene particles coated with SiO2. (ce) TEM images of Janus structure obtained through swelling with 0, 20, 30 and 50% tetrahydrofuran/water mixture. (f) SiO2 nanobottles obtained by dissolution of polystyrene particles with pure tetrahydrofuran. The scale bar is 500 nm. Reprinted with permission from Ref. [73]. Copyright 2021, copyright Wiley-VCH GmbH.
Energies 15 07636 g006
Figure 7. (a) The growth process of Au nanobottle. (b) The change in opening size (d) and the diameter (D) of Au nanobottle. The scale bar is 500 nm. Reprinted with permission from Ref. [40]. Copyright 2019, copyright American Chemical Society.
Figure 7. (a) The growth process of Au nanobottle. (b) The change in opening size (d) and the diameter (D) of Au nanobottle. The scale bar is 500 nm. Reprinted with permission from Ref. [40]. Copyright 2019, copyright American Chemical Society.
Energies 15 07636 g007
Figure 8. (a,b) Photocatalytic hydrogen production activities of with or without nanobottle motors. Reprinted with permission from Ref. [37]. Copyright 2021, copyright Elsevier Ltd. (c) Motion Trajectories and (d) reduction rate of Au-nanobottle motors, Au sphere. Reprinted with permission from Ref. [43]. Copyright 2022, copyright Wiley-VCH GmbH.
Figure 8. (a,b) Photocatalytic hydrogen production activities of with or without nanobottle motors. Reprinted with permission from Ref. [37]. Copyright 2021, copyright Elsevier Ltd. (c) Motion Trajectories and (d) reduction rate of Au-nanobottle motors, Au sphere. Reprinted with permission from Ref. [43]. Copyright 2022, copyright Wiley-VCH GmbH.
Energies 15 07636 g008
Figure 9. (a) Motion velocity of nanobottle motors changes with time in the solution at different concentrations of heavy metal ions. (b) Removal rate of mesoporous silica nanobottle motors in Cu2+ solution. (c) Removal rate of Cu+2 at concentration of catalase-propelled magnetic mesoporous silica nanobottle (Cat-MMNB) motors of 0.5, 1, 2, 3 mg mL−1. Reprinted with permission from Ref. [106]. Copyright 2021, copyright Wiley-VCH GmbH. (d) Nanobottle motor absorbs oil droplets through hydrophobic surface. (e) Nanobottle motor approaches, captures and delivers oil droplets (1 and 2) under magnetic guidance and autonomous motion. The scale bar is 100 μm. Reprinted with permission from Ref. [38]. Copyright 2015, copyright Wiley-VCH GmbH.
Figure 9. (a) Motion velocity of nanobottle motors changes with time in the solution at different concentrations of heavy metal ions. (b) Removal rate of mesoporous silica nanobottle motors in Cu2+ solution. (c) Removal rate of Cu+2 at concentration of catalase-propelled magnetic mesoporous silica nanobottle (Cat-MMNB) motors of 0.5, 1, 2, 3 mg mL−1. Reprinted with permission from Ref. [106]. Copyright 2021, copyright Wiley-VCH GmbH. (d) Nanobottle motor absorbs oil droplets through hydrophobic surface. (e) Nanobottle motor approaches, captures and delivers oil droplets (1 and 2) under magnetic guidance and autonomous motion. The scale bar is 100 μm. Reprinted with permission from Ref. [38]. Copyright 2015, copyright Wiley-VCH GmbH.
Energies 15 07636 g009
Figure 10. (a) Nanobottle motors scavenge intracellular reactive oxygen species. (b,c) CellROX detect reactive oxygen species in the cell. (d,e) Cell survival rate of cancer cell line with five kinds of treatments. The scale bare is 20 μm. Reprinted with permission from Ref. [39]. Copyright 2020, copyright Elsevier Ltd. (f) Chemoattractant release and pathogen capture in the hollow cavity. (g) The chemotactic of pathogen (green) around nanobottle motor (blue). The scale bar is 15 mm. Reprinted with permission from Ref. [118]. Copyright 2019, copyright Wiley-VCH GmbH.
Figure 10. (a) Nanobottle motors scavenge intracellular reactive oxygen species. (b,c) CellROX detect reactive oxygen species in the cell. (d,e) Cell survival rate of cancer cell line with five kinds of treatments. The scale bare is 20 μm. Reprinted with permission from Ref. [39]. Copyright 2020, copyright Elsevier Ltd. (f) Chemoattractant release and pathogen capture in the hollow cavity. (g) The chemotactic of pathogen (green) around nanobottle motor (blue). The scale bar is 15 mm. Reprinted with permission from Ref. [118]. Copyright 2019, copyright Wiley-VCH GmbH.
Energies 15 07636 g010
Table 1. The main references of nanobottles growth.
Table 1. The main references of nanobottles growth.
MethodMaterialsVariablesBasic Steps
Soft-
template
Ribose
Oleic acid
Poly (ethylene oxide)-poly (propylene oxide)-poly (ethylene oxide)
Temperature
Reaction time
  • Produce nanoemulsion
  • Nanoemulsion swelling, the shell cracked by tensile force
  • Nanoemulsion flows out, forming the neck
Swelling-
inducement
Polystyrene microspheres
SiO2
Ethanol
Pressure
Volumetric percentage
of solute in
tetrahydrofuran/water
  • After absorbing tetrahydrofuran, the swollen microsphere pokes a hole in the shell
  • The swollen polystyrene is squeezed through opening
  • Ethanol quenching
Wet-chemistry
synthetic
Acetate
Thioacetamide
Cetyltrimethyl ammonium bromid
Au
PbS
Weak acids
Au/PbS ratio
  • Obtain the edge PbS nanooctahedron
  • Au deposition, formation of Au/PbS nanostructure
  • Adding weak acids to dissolve and remove sulfide
Table 2. Some typical applications of nanobottle motors.
Table 2. Some typical applications of nanobottle motors.
FieldRepresent ExamplesMaterials InvolvedPrinciple
EnergySolar water splittingCarbonLarge specific surface area, low density, accelerates the mass transfer rate through motion [37]
PhotocatalyticSilica, Au, Fe3O4Introduces chemically active ingredients, accelerates the mass transfer rate through motion [43]
EnvironmentEnvironment monitoringMnFe2O4The motion of nanobottle motors in the solution can grab the target molecule or ion [38]
Environment RemediationMagnetic silicaLarge specific surface area of nanobottle motors has affinity for organic pollutants [106]
BiomedicineTargeted therapyMgGreat performance of drug loading, tissue penetration and biocompatibility [39]
Trapping pathogensMgThe Mg core is depleted, the exposed inner surface dissolves, releasing chemoattractant [118]
Drug deliveryMgTiO2Bubble propulsion [120]
Photothermal therapy (PTT)AuPhotothermal conversion [40]
Treat gastroenteric diseasesSilicaThe reaction of CaO2 with gastric juice neutralizes gastric acid [52]
Drug loading and controlled releasePhase-change material (PCM)Controlling the temperature causes the phase-change material blocking the hole to change the state of the phase [29,121]
Table 3. The propulsion mechanisms, fabrication methods and potential applications of nanobottles as well as their major advantages and limitations.
Table 3. The propulsion mechanisms, fabrication methods and potential applications of nanobottles as well as their major advantages and limitations.
PropulsionMechanismFabrication MethodFeaturePotential ApplicationsAdvantagesChallenges
Chemical propulsionBubbles separated
from motor
Soft-template methodSelf-assembly between precursor molecules and core templates.
The fabrication makes it easy to control morphology.
Solar water splitting/Photocatalytic
  • Large surface area
  • Low density
  • Enhances the mass transfer
  • Low speed of motion
  • Light utilization is difficult
  • Requires photocatalysts
Asymmetrical
concentration gradient
Light drivingAsymmetric
temperature gradient
Swelling-
inducement method
Swelling polymer microspheres with organic solvent and then quenching with ethanol.
Opening structure is obtained through diffusion of emulsion solvent from inside of nanoparticle.
Environment monitoring/
Remediation
  • Fast motion in solution
  • Strong affinity for organic pollutants
  • Fuel consumption
  • The motion and post-treatment of nanobottles have unknown effect on real environment
Magnetic actuationApply magnetic force or torque to magnetized motors.Wet-chemistry synthetic methodControlling reaction dynamics and thermodynamic parameters to control the sizes and shapes of noble metal nanobottles.
A facile and robust method.
Targeted therapy/Trapping pathogens
  • Asymmetrical cavity structure
  • Autonomic motion
  • High impermeability
  • Achieves functional nanobottles in complex environment
  • Biocompatibility/degradability
  • Difficult to work independently
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Q.; Wang, L.; Wang, K.; Wang, T.; Liu, G. Principle, Fabrication and Emerging Applications of Nanobottle Motor. Energies 2022, 15, 7636. https://doi.org/10.3390/en15207636

AMA Style

Liu Q, Wang L, Wang K, Wang T, Liu G. Principle, Fabrication and Emerging Applications of Nanobottle Motor. Energies. 2022; 15(20):7636. https://doi.org/10.3390/en15207636

Chicago/Turabian Style

Liu, Qingyuan, Lin Wang, Kaiying Wang, Tianhu Wang, and Guohua Liu. 2022. "Principle, Fabrication and Emerging Applications of Nanobottle Motor" Energies 15, no. 20: 7636. https://doi.org/10.3390/en15207636

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop