Next Article in Journal
Research on Aerodynamic Force/Thrust Vector Combined Trajectory Optimization Method for Hypersonic Drones Based on Deep Reinforcement Learning
Previous Article in Journal
Agentic AI for Real-Time Adaptive PID Control of a Servo Motor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Performance Comparison of Mechanical and Ferrofluidic Micropumps: Structural and Operational Perspectives

1
Key Laboratory of Fluid and Power Machinery, Ministry of Education, Xihua University, Chengdu 610039, China
2
Engineer School, Qinghai Institute of Technology, Xining 810016, China
*
Author to whom correspondence should be addressed.
Actuators 2025, 14(9), 460; https://doi.org/10.3390/act14090460
Submission received: 14 August 2025 / Revised: 13 September 2025 / Accepted: 17 September 2025 / Published: 20 September 2025
(This article belongs to the Section Miniaturized and Micro Actuators)

Abstract

Since the successful implementation of microfluidic technology in biomedical applications, research on micropumps—the central component of these systems—has gained significant momentum. Benefiting from advancements in pump materials and corresponding fabrication methods, micropumps have evolved from structurally complex mechanical designs to simpler non-mechanical configurations. This paper reviews well-developed mechanical micropumps, discussing their diaphragms, pump chambers, materials, and other aspects to outline their developmental trajectory and current applications, while also highlighting their limitations. After identifying the shortcomings of traditional micropumps, we introduce the concept of ferrofluid-based micropumps, emphasizing their structural simplicity, self-sealing capability, and recoverability. Previous research on ferrofluidic micropumps is summarized, demonstrating their superior performance in certain aspects. Finally, we provide an outlook on their potential applications in biomedicine and specialized fields.

1. Introduction

In biological [1,2,3,4], medical [5,6,7,8], and even space exploration [9,10] applications, precise control of minute fluid volumes—termed microfluidics [11,12,13,14]—has gained increasing attention. This technology encompasses fluid transport [15], separation [16,17], and mixing [18], exemplified by the following: (1) delivering biological samples to miniaturized analysis systems [19,20], (2) isolating magnetic beads from cells [21,22,23], and (3) conducting novel mixing reactions in microreactors [24,25]. While passive mechanisms like surface tension [26], gravity [27,28], or the Marangoni effect [29,30] can manipulate microfluids, they inherently lack control over flow volume, direction, and process termination—an unacceptable limitation for microfluidic systems requiring microliter-scale precision (e.g., insulin delivery [31,32]). This critical need for active actuation has driven the development of diverse micropumps [33,34,35,36] based on various working principles, each with distinct advantages and limitations. In recent years, microfluidics has been researched and applied in the biomedical and digital fields, as shown in Figure 1.
This review examines mechanical micropumps in microfluidics, analyzes their applications and shortcomings, and introduces an innovative alternative: ferrofluidic micropumps with superior performance characteristics.
In the 1980s, micropumps first found medical application when Sefton et al. [41] pioneered their use in controlled insulin delivery systems, maintaining diabetic patients’ blood glucose levels throughout the day while reducing needle frequency. For implantable micropumps, high volumetric flow rates are unnecessary (e.g., insulin requirements typically remain below 1 mL/day [42,43]), but precise dosing is critical. These systems must also generate sufficient back pressure to overcome substantial physiological pressures (approximately 25 kPa [44]). Furthermore, reliability, cost-effectiveness, and biocompatibility constitute essential evaluation metrics for implantable micropump designs [45,46,47]. Figure 2 shows some implantable microfluidic devices.
In recent years, miniaturized chemical and biological analysis systems [53,54,55] have garnered significant attention. Compared with conventional analytical platforms, these microsystems offer distinct advantages including reduced sample/reagent consumption [56], accelerated analysis rates, and minimized human-induced errors [57,58].
Micropumps represent ideal fluid-handling components for such systems, where their precise pumping capability ensures accurate reagent proportioning. However, widespread adoption remains hindered by fabrication complexity and high costs, leading researchers to prefer low-cost alternatives such as manual pipetting [59], external pneumatic actuation [60], or induced electroosmotic flow [61,62]. Current research has proposed several disposable micropump designs: Qi et al. [63] developed a reusable-magnet disposable electromagnetic micropump by decoupling expensive magnets from low-cost pump bodies; Meng et al. [64] created an economical PDMS-based micropump that uses medical-grade tape and filter paper to detect sickle cell detection, as shown in Figure 3 below; Ohashi et al. [65] engineered a sweat-collection chip incorporating disposable micropump components. Abate et al. [66] proposed a method to produce monodisperse emulsions using only microfluidic devices and manual syringes, eliminating the need for bulky pumps, control computers, and other additional equipment.
To address diverse application scenarios, over 200 novel micropump designs have been reported since their initial development. For systematic research, scholars typically classify micropumps into two categories based on the presence of moving parts [67]: (1) Mechanical micropumps containing oscillating diaphragms or pistons [68] as moving components, and (2) non-mechanical micropumps that convert various forms of non-mechanical energy into kinetic energy [69] without relying on moving parts. Mechanical micropumps, also termed displacement pumps, operate by displacing fluid through movable boundaries typically provided by pistons or elastic membranes [70], such as piezoelectric micropumps [71] and electrostatic micropumps [72]. In contrast, non-mechanical micropumps utilize field effects (e.g., electromagnetic [73] or thermal fields [74]) to induce ionization or phase changes in working fluids, thereby generating pressure and flow. Notably, we highlight an innovative design—the ferrofluidic micropump [75]—which leverages its unique liquid material (ferrofluid [76]) to achieve pumping through coordinated interaction with external magnetic fields.
To demonstrate the superior operational principles and application potential of ferrofluidic micropumps, we conduct a systematic comparison with mature mechanical micropump technologies. Our evaluation begins by classifying mechanical micropumps according to their actuator types, followed by detailed examinations of their core components, including pump bodies, diaphragms, and valves. Through this analysis, we identify inherent limitations of mechanical designs that highlight the advantages of ferrofluid-based alternatives. For standardized performance comparison, we employ two key metrics: maximum flow rate Qmax and maximum pressure difference ΔPmax, supplemented with dimensional parameters of pump chambers and diaphragms to assess spatial efficiency. Operational characteristics such as driving voltage U and working frequency f are analyzed for their impact on power consumption and operational safety. Notably, evaluation parameters for ferrofluidic micropumps shift focus from electrical inputs to magnetic field strength due to their fundamentally different actuation mechanism. The study presents three comprehensive comparison tables: one summarizing performance data of mechanical micropumps reported in the past decade, and another detailing geometric parameters (length, divergence angle, etc.) and materials of nozzle/diffuser elements used for flow rectification. Finally, we will select some mechanical micropumps and ferrofluidic micropumps and compare their flow rate and back pressure data to evaluate the performance of the two types of micropumps under conditions of comparable power consumption and safety levels. Special attention is given to the analysis of unidirectional valves and currently prevalent diffuser structures.

2. Mechanical Micropumps and Their Compositions

The current literature indicates that mechanical micropumps constitute the predominant category [77,78]. These pumps rely on internal movable boundaries or surfaces, typically implemented through either pistons or diaphragms [79]. While piston-based designs face challenges in miniaturized sealing technology, limiting their practical application, most contemporary mechanical micropumps employ deformable diaphragms as the dynamic boundary. The diaphragm’s oscillation modulates the pump chamber volume, enabling liquid transport through controlled pressure variations. A standard mechanical micropump comprises three key components: (1) a pump chamber (housing the diaphragm/piston), (2) an actuator, and (3) paired check valves (inlet and outlet), with the operational sequence illustrated in Figure 4. The actuator generates a driving force transmitted to the diaphragm, inducing pneumatic pressure changes that draw fluid into the chamber. This working principle exemplifies the fundamental mechanism of mechanical micropumps—coordinated interaction between the actuator and movable boundary to cyclically alter chamber volume (pressure), thereby producing controlled fluid discharge at specified flow rates.
It is important to note that micropump diaphragms are not always single-layer. For example, the multi-layer micropump developed by Nguyen et al. [80] contains three separate diaphragms, each driven by individual piezoelectric actuators, making it particularly suitable for peristaltic pumping. The structure is composed of four bonded material layers: the upper PDMS plate contains one rectangular channel and two inlet/outlet seals, while the lower PDMS plate integrates three rectangular micro-lightweight piezoelectric composite actuators, with an outer PMMA cover. During pumping, the actuators sequentially act on the three independent diaphragms, causing the fluid to move through the gaps to the outlet. When the first diaphragm is acted upon by the actuator’s force, fluid begins to enter; the system then applies force to the second diaphragm while removing the force from the first diaphragm, allowing the first diaphragm to return to its original position, creating a sealing effect that prevents backflow. Subsequently, the third diaphragm is forced to open the channel while the second diaphragm returns to its original state, pushing the fluid in the pump chamber out to the outlet. Finally, all diaphragms reset, completing the pumping process.

2.1. Actuator

Given the structural variations arising from different actuation principles, we categorize mechanical micropumps based on their actuator types. The predominant actuation methods include the following: piezoelectric [81,82,83], electrostatic [84], electromagnetic [85], pneumatic [86,87,88]. Current research indicates that piezoelectric actuators dominate the micropump field, representing the most extensively studied configuration.
The actuator applies force to the diaphragm, causing its deformation or displacement. Piezoelectric actuators operate based on the piezoelectric effect, where an applied voltage induces deformation in piezoelectric crystals (such as quartz and barium titanate-based ceramics), thereby deforming the attached diaphragm and altering the pump chamber volume. Electrostatic actuators essentially function as capacitors composed of two parallel electrodes, which exert force on the pump diaphragm positioned between the plates according to Coulomb’s law, resulting in its deformation. Electromagnetic actuators utilize magnetic effects, where external magnets apply force to either a pump diaphragm embedded with magnetic materials or a permanent magnet piston, inducing vibration. Pneumatic micropumps do not require specialized actuators.
Piezoelectric actuation, as the earliest driving principle applied in the field of micropumps, is characterized by a large driving force, long diaphragm stroke, and fast mechanical response. Currently, commercial piezoelectric materials mainly include lead zirconate titanate (PZT), polyvinylidene fluoride (PVDF), and aluminum nitride (AlN). Among these, PZT materials are most commonly used due to their high electromechanical coupling coefficients, as evidenced by their adoption in studies by Munas et al. [89], Kan et al. [90], and Johnston et al. [91] using PZT-based piezoelectric actuators. Electrostatic actuators consist of conductive and insulating layers. The conductive layer typically employs metal thin films (aluminum, gold, copper) or doped semiconductors (polycrystalline silicon, graphene), while the insulating layer prevents short-circuiting between electrodes.
The fundamental structure of electromagnetic actuators resembles that of electromagnets, consisting of coils and magnetic core materials (e.g., Fe-Si alloys, Mn-Zn ferrites). The force exerted on the internal diaphragm can be expressed as
F = B 2 S 2 μ 0
where F represents the actuator force on the diaphragm (unit: newton, N); B denotes the magnetic flux density generated by the actuator (unit: tesla, T); S is the effective actuator area (unit: square meter, m2); μ0 is the vacuum permeability (μ0 = 4π×10−7 T·m/A). The magnetic flux density B can be approximated by
B = μ 0 n I
where n is the number of coil turns per unit length; I is the current through the coil (unit: ampere, A).
B c N I L
where I represents current (unit: A), and L denotes the characteristic length of the actuator coil. This indicates that the magnetic field strength is directly proportional to the number of coil turns and current, while inversely proportional to the actuator’s dimensions. Theoretically, achieving sufficient actuation force requires increasing coil turns and current while minimizing inter-coil gaps. However, excessive coil turns may lead to device bulkiness, and high current can cause Joule heating that risks actuator burnout. Current electrostatic actuators typically employ 50–200 coil turns. For particularly miniaturized MEMS micropumps, fewer turns are required—for instance, Amrani et al. [92] utilized a 12-turn coil to reduce device size.
The actuation mechanisms for mechanical micropumps are not limited to the four types mentioned above but also include other principles such as phase change actuation and shape memory alloy (SMA) actuation. Figure 5 summarizes the currently dominant actuator technologies in the field.
The author created a table (Table 1) summarizing the advantages, disadvantages, and typical applications of several types of mechanical micropumps shown in Figure 5.

2.2. Diaphragm

The diaphragm serves the critical function of altering the pump chamber volume. As previously mentioned, flexible membranes are typically employed to provide movable boundaries. These diaphragms are generally positioned either at the periphery [93] or center [94] of the pump chamber and connected to fixed boundaries. The diaphragm thickness directly impacts both the micropump’s driving force [95] and durability [96]. Standard diaphragm thicknesses range from 5 μm to 500 μm. For instance, Lin et al. [97] developed a valveless pneumatic micropump with a 300 μm thick diaphragm.
The diaphragm of the piezoelectric micropump is bonded to the actuator and placed on a metal substrate. The arrangement of diaphragms in electrostatic micropumps is divided into two types: the monolithic long diaphragm (5 mm scale, 60 μm thickness) fabricated by Machauf et al. [98] in 2005, and the series of diaphragms (eight pieces, each 2.5 μm wide with 1.0 μm spacing) made by Lee et al. [99] in 2012. A single diaphragm can cause alternating changes in the pump chamber volume through its own vibration, whereas multiple diaphragms are arranged parallel to each other with equal spacing and rely on the sequential action of an external actuator to move the diaphragms in turn, thereby opening the flow channel. After the fluid passes through, the diaphragms immediately return to their original state to seal the flow channel and prevent backflow. Additionally, it should be noted that the diaphragm in an electrostatic micropump is located in the middle of the pump chamber, so the number and thickness of the diaphragms affect the smoothness of fluid flow and the effective volume of the pump chamber, respectively. For electromagnetic micropumps, the diaphragm must be able to respond to external magnetic forces. Therefore, magnetic materials (Fe3O4) need to be uniformly mixed into the conventional diaphragm material or a soft magnetic metal layer must be coated. In 2019, Tahmasebipour et al. [100] used Fe3O4-doped PDMS material to fabricate a diaphragm, which reduced the diaphragm recovery time while maintaining biocompatibility.
In theory, the diaphragms of micropumps can vibrate bidirectionally, which has spurred a new research direction: bidirectional pumping. Bidirectional pumping helps reduce the vacuum time during micropump operation and improves pumping efficiency. Lin et al. [101] proposed a three-electrode electrostatic soft pump capable of bidirectional pumping for robotic applications. Nicola et al. [102] developed a dual-enzyme micropump by embedding β-glucosidase and urease into a patch, demonstrating inward flow (0.95 ± 0.37 μm·s−1, 20 mM) in a cellobiose environment and outward flow (1.46 ± 0.47 μm·s−1, 20 mM) in the presence of urea.
For pneumatic micropumps, a diaphragm is not mandatory. Pneumatic micropumps can rely on gas pressure acting on the diaphragm [103,104] to induce compression or expansion. Alternatively, they may utilize direct gas-driven liquid flow without requiring a physical diaphragm to separate the gas and liquid phases. For instance, Ni et al. [105] designed a diaphragm-less pneumatic micropump incorporating a pair of thin check-valve baffles as a substitute for the pump diaphragm. The diaphragm-based approach is widely used in biomedical applications, primarily because it prevents gas–liquid contact within the pump, thereby avoiding drug contamination [106]. The diaphragm-less design, on the other hand, eliminates the need for complex diaphragm fabrication, significantly reducing manufacturing difficulty.
However, it is interesting to note that the field of droplet microfluidics, which is closely related to pneumatic micropumps, heavily relies on the use of membrane structures. Micrometer-sized droplets serve as ideal reactors for researchers [107], and these highly valuable droplets require the driving force provided by pneumatic micropumps and microfluidic chips with varying channel dimensions equipped with membrane valves [108]. These membrane structures can be categorized into single-layer valves and multi-layer valves, capable of performing functions such as droplet generation, sorting, guidance, and trapping. Abate et al. [109] developed a single-layer membrane valve capable of continuously regulating flow rates. This valve combines the simplicity and ease of fabrication of stamped microfluidic devices with the precise control characteristics of membrane valves. In 2010, Abate et al. [110] further utilized high-speed single-layer membrane valves to control flow rates and perform sorting functions, even extending their applicability to live cells. Agnihotri et al. [111] employed a large number of single-layer microvalves to create a highly biocompatible droplet splitting and merging system. This system was designed to synthesize droplets of varying concentrations and types, enabling them to carry biological samples. Figure 6 illustrates the operational workflows of several chips based on pneumatic micropumps, which are renowned for their uniquely complex chip architectures and extensive utilization of valves.

2.3. Valves

Valves are responsible for the rectification of microfluid flow [112]. To suppress backflow, researchers have introduced various types of valves. In the early stages of micropump development, Spencer employed active flap valves made of piezoelectric bimorphs with dimensions of 0.4 mm × 4 mm × 20 mm. However, the opening and closing of active valves require external energy input, which is detrimental to system integration and hinders the automation of micropumps and their systems. Currently, most micropumps utilize passive valves, such as the ball valve used by Liu et al. [113], the flap valve employed by Cesmeci et al. [114], and the umbrella valve adopted by Jie et al. [115]. These valves demonstrate effective rectification in micropumps and do not require additional energy for actuation. Microvalves are typically comparable in size to the micropump body and are often fabricated as an integrated unit. Meanwhile, in droplet-based systems, valves find more extensive applications, particularly membrane valves. For instance, the multi-layer valve system developed by Grover et al. [116] delivers a flow rate of up to 380 nL/s at a driving pressure of 30 kPa and reliably seals against fluid pressures of up to 75 kPa, providing a simple and robust integrated solution for fluid manipulation in complex glass-based microfluidic and electrophoretic analysis devices. Raveshi et al. [117] utilized a microfluidic system based on single-layer valves to achieve selective droplet splitting, enabling parent droplets to move freely on the chip while ensuring accurate positioning and expulsion of smaller daughter droplets.
Ongoing development and research on valves, particularly for pneumatically driven microfluidic components, continues to advance. In 2011, Roy et al. [118] reported a rapid fabrication technique and demonstrated its application in creating and assembling pneumatic valves in multi-layer microfluidic devices entirely made of thermoplastics. Unlike conventional normally open valves (NO valves), Mohan et al. [119] proposed a pneumatically actuated normally closed valve (NC valve), which enhances the portability of microvalve systems.
However, compared to other components of the pump, microvalves demand higher fabrication precision and more complex manufacturing processes due to their open–close functionality, posing certain challenges in production and application. Similarly to diaphragms, many microvalves exhibit extremely thin structures and are prone to irreversible plastic deformation or even failure due to long-term cyclic vibration and repeated actuation, ultimately leading to pump malfunction.
To address these issues, researchers have adopted tapered diffusers to replace microvalves in developing valveless micropump designs. These diffusers operate on Bernoulli’s principle, utilizing asymmetric inlet/outlet widths to regulate internal fluid pressure, thereby preferentially directing microfluidic flow from narrow to wide sections while increasing flow velocity. The diffusers reduce device complexity and manufacturing difficulty by eliminating movable microvalve components, allowing integration of rectifying elements with the micropump body, which simplifies fabrication and minimizes leakage risks. Without moving parts, the micropump’s lifespan is no longer limited by microvalve durability, significantly extending the system’s operational lifetime. It should be noted that diffuser dimensions critically influence internal flow velocity, requiring computational fluid dynamics analysis to determine optimal diffuser geometry for achieving stable flow conditions in the micropump system. Key design parameters include tube length, expansion angle, and wide/narrow section dimensions, with specific geometric parameters from various researchers’ micropump diffusers presented in Table 2.

2.4. Pump Chamber and Pump Body Materials

The shape of pump chambers varies, with circular [133] and square [134] designs being the most common. Electrostatic micropumps typically adopt a square chamber due to the arrangement of internal electrode plates. In the 1980s, high-strength materials such as stainless steel [71] and low-carbon steel [134] were initially used for pump bodies. However, due to their poor corrosion resistance and limited machining precision (micron-level), metal-based pump structures gradually fell out of favor. To achieve a smaller size and higher integration [135], researchers turned to silicon and glass [136,137,138], which offered superior machining accuracy. With advancements in fabrication technology, transparent polymer materials—such as polymethyl methacrylate (PMMA) [139,140] and polydimethylsiloxane (PDMS) [141,142]—have gained prominence. These materials are favored not only for their excellent biocompatibility [143] but also for their low manufacturing cost [144]. Notably, their transparency enables real-time observation of fluid dynamics inside the pump, providing deeper insights into pumping mechanisms. This also facilitates optical research methods, such as particle tracking velocimetry (PTV), further advancing micropump studies. However, a critical limitation of these polymers is their proneness to deformation at elevated temperatures (around 80 °C). Thus, environmental temperature control is essential during operation. Additionally, if electromagnetic coils are integrated, excessive Joule heating must be avoided to prevent plastic deformation, which could lead to structural failure.

2.5. Fabrication Techniques

The development of micropumps has been closely tied to advancements in pump body machining technologies. In the early stages, due to the maturity of mechanical machining, metal-based micropumps progressed more rapidly. For instance, W. J. Spencer [145] utilized stainless steel as the pump body material.
With the refinement of higher-precision photolithography and lower-cost injection molding processes, non-metallic micropumps emerged and eventually became mainstream. Today, the field of micropump manufacturing has incorporated technologies such as CNC machining [146], CO2 laser micromachining, and 3D printing [147,148,149]. A schematic diagram of these Machining Methods is shown in Figure 7. These methods enable the fabrication of complex structures that are difficult to achieve using conventional techniques like photolithography, etching, or injection molding. This diversification in manufacturing processes has expanded design possibilities, allowing for more intricate and optimized micropump architectures.
In the current micropump manufacturing industry, few researchers have explored monolithic fabrication processes. Instead, most adopt a multi-layer fabrication [136] and assembly approach. For example, Li et al. [150] successfully integrated a three-layer complex structure into a single micropump chip by rapidly injecting liquid metal into blind-end microchannels. Ni et al. [151] developed a five-layer micropump structure consisting of a PDMS substrate, a thin PDMS functional layer, a 150 μm thick PDMS flexible membrane, a glass plate with through-holes, and a PDMS top layer for pneumatic actuation.
The multi-layer approach enhances manufacturing flexibility, enabling precise fabrication and assembly of differently shaped layers [152], while also improving overall machining accuracy [153]. However, post-assembly bonding and sealing remain major challenges. Figure 8 illustrates a micropump chip fabricated using this multi-layer assembly technique.

2.6. External Equipment of Micropump

External devices provide the driving force for micropumps. For piezoelectric and electrostatic micropumps, the primary external equipment is a power supply. For driving PZT, the most widely used actuator material, a driving voltage of 50–200 V [155,156] is required. Electrostatic micropumps usually demand even higher driving voltages, ranging from 200 to 1000 V [157]. To ensure safe operation, series resistors (to limit current) and TVS diodes (for electrostatic discharge protection) must be incorporated into the circuit. The high operating voltages lead to complex and costly driving electronics [158], making voltage reduction a key research goal for both piezoelectric and electrostatic micropumps [159].
Electromagnetic micropumps require a controllable power supply to drive an electromagnet, which actuates an internal magnetic diaphragm.
For pneumatic micropumps, a high-pressure gas source is needed to deliver chemically stable gas. Industrial systems typically use oil-free air compressors, while laboratory setups rely on miniaturized high-pressure gas cylinders. To ensure safe operation within the micropump’s pressure limits, a pressure-reducing valve must be integrated. Additionally, biomedical applications often require an extra filtration system, leading to increased complexity and hindering miniaturization efforts.
To address these challenges, researchers have explored alternative gas generation methods. For example, Ochoa et al. [160] demonstrated a compact pneumatic micropump system by filling the pump chamber with a yeast solution and drugs, which produced low-pressure actuation gas through fermentation, effectively reducing the system’s overall size.

2.7. Applications of Mechanical Micropumps

The earliest mechanical micropumps were applied in the medical field, including drug delivery systems [161,162,163,164], in vitro diagnostics [165], and mechanistic studies of artificial organs [166] (such as the liver). Examples include a miniature pulmonary artery cell culture chip using mechanical micropumps as the power source [167] and the highly reliable integrated microfluidic plate for cell culture-based multi-organ microphysiological systems developed by Al-Hilal et al. [168]. Indeed, micropumps are not far removed from our daily lives. The printers we commonly use today leverage the advantages of piezoelectric micropumps to precisely control ink dosage, delivering superior performance compared to earlier printers that relied on micro-heaters to generate bubbles (incidentally, this “bubble jet” technology can also be classified as a type of thermal bubble micropump, which similarly falls under the category of micropumps).
As shown in Figure 9, the author presents schematic diagrams and physical images of (a,b) pulmonary artery cells and (c) a multi-organ cell culture chip.
In the chemical field, micropumps are used for quantitative delivery and addition of reagents, such as experimental testing of new reactions [170] and etching solution dispensing in the semiconductor industry [171]. In environmental and energy applications, there are micropump-based portable water quality detectors [172,173] and fuel cells [174]; in scientific research, applications include droplet generation [175] and organ-on-a-chip culture [176,177]. In aerospace, microfluidic technology can integrate the functions of a traditional laboratory into a micron-scale chip. Utilizing suitable mechanical micropumps enables unmanned space experiments, overcoming limitations such as restricted space environments, high experimental costs, and reliance on manual operations. Additionally, it allows for long-term experiment operation and remote data recording [178].
In the future, microfluidic devices primarily based on mechanical micropumps may trend toward miniaturization and integration. However, it is regrettable that although most existing cases are exemplary in laboratory-scale miniaturization techniques, they remain essentially independent functional modules—analogous to individual components in an integrated circuit—rather than truly integrated systems. Thorsen et al. [179] pioneered the convergence of microfluidics and electronic circuits by using a chip as a comparator and enzymes as mediators. Through bacterial colonies producing fluorescent outputs analogous to signal outputs in electronic circuits, they successfully achieved genuine integration of a microfluidic system. Using this as a foundational clue, continued research into the integration pathway of micropumps will undoubtedly unlock new applications for such technology.

2.8. Failure of Mechanical Micropumps

The failure of mechanical micropumps can be divided into three types: material fatigue, drive system failure, and fluid compatibility. The diaphragm of mechanical micropumps may undergo highly irreversible deformation or fracture due to long-term repeated vibration or overloaded operation, and even increasing diaphragm thickness and limiting voltage [92] cannot completely prevent diaphragm failure. Drive system failure manifests as actuator failure, such as piezoelectric ceramic cracking or electromagnetic coil burnout, which can be avoided by limiting voltage [180] and enhancing heat transfer [181]. Additionally, pumping unsuitable fluids such as corrosive substances [182], high-viscosity solutions [158], and gas-containing liquids [183] may damage the micropump and cause failure.
All three types of damage can lead to micropump failure. Material fatigue is inherent to the working principle of mechanical micropumps and can only be mitigated, not completely avoided. This demonstrates that mechanical micropumps have unavoidable and irreparable design limitations, resulting in certain application shortcomings, including short lifespan, high susceptibility to damage, and non-repairability.

2.9. Challenges of Mechanical Micropumps

Compared to the ferrofluidic micropump to be introduced next, mechanical micropumps have certain disadvantages, which are mainly concentrated in three aspects: high manufacturing difficulty, reliance on high driving voltage or current, and short lifespan. The author will provide a brief analysis below.
First, the manufacturing difficulty of mechanical micropumps is reflected in two parts: the processing of tiny movable components and the assembly of various parts. (1) The fabrication of miniature components, such as thin films or microelectrodes, requires these extremely small components to be manufactured with uniform thickness, which poses a challenge to manufacturing technology. (2) The assembly of movable components is another point that differs from ferrofluidic micropumps. Most mechanical micropumps can only function properly after all components are precisely installed; otherwise, the pump will fail. The challenges associated with assembly at this micron scale are immense.
Second, the reliance of mechanical micropumps on high driving voltage or high current stems from their own driving principles: Mechanical micropumps that rely on high driving voltage are mainly piezoelectric and electrostatic micropumps. The former is because the deformation of the piezoelectric film is proportional to the electric field strength, and only a sufficient voltage-provided electric field can drive the extremely thin film; the latter requires sufficient electrostatic force to operate, but due to the micron-sized air gap between the electrodes, a high voltage is needed. Micropumps known for high current mainly involve thermal bubble micropumps and electromagnetic micropumps. The former requires significant Joule heat to heat the medium fluid, while the latter relies on high current to generate a large Ampere force (Ampere force is proportional to current intensity). High driving voltage or high current makes it difficult to ensure safety during their use. In contrast, the ferrofluidic micropump only requires a power supply with a voltage ranging from a few volts to over ten volts, and its maximum current generally does not exceed 0.5 A. Moreover, this power supply does not directly act on the micropump chip but drives the permanent magnet to indirectly act on the ferrofluid inside the chip.
Finally, due to long-term exposure to high voltage or current, as well as wear and tear or fatigue caused by repeated vibration of internal movable components, mechanical micropumps are more prone to failure compared to ferrofluidic micropumps. If any single component fails, the entire mechanical micropump’s lifespan ends, and it is difficult to repair.
In summary, the development of mechanical micropumps can focus on reducing manufacturing and assembly difficulties and achieving low-voltage, low-current driving. Additionally, using materials that are less prone to fatigue and fracture to make mechanical micropump components may also become a solution to the micropump’s pain points.
For a clearer understanding, we have compiled relevant mechanical micropump images in Figure 10, and some representative working data is shown in Table 3.

3. Composition and Development Status of Ferrofluidic Micropumps

In the classification of micropumps, ferrofluid belongs to non-mechanical micropumps as it contains no movable mechanical components. However, it can also be considered a piston pump since the internal ferrofluid functions as a “piston”. Compared to various mechanical micropumps mentioned above, ferrofluidic micropumps demonstrate distinct advantages due to their unique composition—ferrofluid—including non-contact operation, repairability, self-sealing capability, long lifespan, and simple structure.

3.1. Ferrofluid

Ferrofluid is a special type of liquid, consisting of a stable colloidal suspension where magnetic particles are uniformly dispersed in a liquid carrier medium, appearing as a black colloidal substance. It was invented by Stephen in 1963 [197]. Normally, ferrofluid does not exhibit magnetism, but under a magnetic field, it demonstrates ferromagnetic properties [198] and aggregates toward areas of stronger magnetic field intensity. When subjected to a sufficiently strong permanent magnet, the ferrofluid coalesces into a compact liquid mass, forming a piston-like structure within microchannels. By moving the external magnetic source, this piston can be precisely controlled. A typical ferrofluidic micropump system [199] comprises the pump body, ferrofluid, external magnetic source, motor, and power supply. Unlike mechanical micropumps that require high voltage and energy consumption, the power supply in ferrofluidic micropumps only provides energy for the motor, with these motors typically operating at rated voltages below 20 V [200], endowing ferrofluidic micropumps with superior safety and energy efficiency characteristics. Figure 11 shows a physical image of the magnetic fluid.

3.2. Actuating Magnets

Early ferrofluidic micropumps employed electromagnets [201] with adjustable magnetic fields and no moving mechanical parts as external driving magnetic sources. For example, Gusenbauer et al.’s micropump [79] utilized ten electromagnets with 200-turn coils driven by a 1–6 A current. Compared to permanent magnets as alternative magnetic sources, electromagnets offer superior controllability by enabling magnetic field activation and intensity adjustment through current variation. However, electromagnets fail to provide the required magnetic field strength [202] (typically around 0.3 T for micropump operation) as theoretically expected, while also presenting Joule heating challenges [203]. Researchers have progressively adopted permanent magnets (NdFeB, SmCo, etc.) with diameters in the range of 6–10 mm [204], which can generate magnetic fields up to 1 T while maintaining compact dimensions comparable to buttons and exhibiting no heat generation. Magnetic field intensity adjustment can be achieved by modifying the size and quantity of permanent magnets. Furthermore, to fully utilize strong magnetic fields, the distance between external magnetic sources and the micropump body should be minimized (maintained below 1 mm) [205].

3.3. Valves and Sealing Means

Unlike other micropumps, ferrofluidic micropumps achieve self-sealing [206]. In addition to the actuating magnetic source, conventional ferrofluidic micropumps incorporate a stationary sealing magnetic source positioned between the inlet and outlet. This attracts part of the ferrofluid within the microchannel to form a seal [207]. The pumped liquid must be immiscible with the ferrofluid (achieved by ensuring incompatibility between the ferrofluid carrier liquid and the pumped liquid), while also exhibiting poorer wettability to the pump body than the ferrofluid. These conditions prevent backflow of the pumped liquid to the inlet. Consequently, ferrofluidic micropumps require no valves to suppress backflow. However, some researchers have incorporated check structures to accommodate specific designs. For instance, in 2018, Liu et al. [208] employed a C-shaped baffle with a 7.1 mm inner diameter and a 7.5 mm outer diameter in a ferrofluidic micropump; Zhu et al. [209] created a fish-mouth baffle to coordinate with kerosene-based ferrofluid in the micropump; and Tang et al. [210] fabricated a butterfly valve for the micropump using polyimide (PI) slices, movable magnets, and elastic silicone diaphragms.

3.4. Pump Body and Manufacturing

3.4.1. Classification of Pump Body Shapes

Current ferrofluidic micropump designs can be categorized into three types: annular, circular, and linear configurations. The annular pump design originated from Hatch et al.’s [205] pioneering work on ferrofluidic micropumps incorporating stationary magnetic sources for self-sealing functionality, representing the initial proposed micropump architecture (shown in Figure 12). Distinct from conventional micropumps, this design eliminates reliance on adhesives or sealants by utilizing the inherent properties of ferrofluid to achieve sealing.

3.4.2. Fabrication of Ferrofluidic Micropumps

To facilitate the observation of ferrofluid motion in micropumps, current mainstream pump body materials are predominantly transparent, including PMMA, PDMS, and glass. Important considerations include the following: PMMA’s poor heat resistance requires avoidance in environments at 80 °C or higher; while glass exhibits fragility and poor machinability with high processing costs, its chemical inertness makes it perfectly suitable for chemical microreactors [211,212,213].
Currently available fabrication methods primarily include the following options: ① Soft lithography, once introduced by Unger et al. [214] in the year 2000 in the journal Science, enables the creation of complex microchannels and is suitable for soft materials like PDMS, as demonstrated by Liu et al.’s [208] PDMS micropump with C-shaped baffles. ② MEMS processing employs lithography or etching to produce precise microchannels in rigid materials such as glass and silicon, exemplified by silicon-based pump bodies fabricated by Hatch et al. [205] in 2001 and Hartshorne et al. [215] in 2004. ③ Three-dimensional printing is exclusively applicable to resins and can construct complex micropump structures (e.g., miniature diffusers), though material limitations result in surface roughness—Gusenbauer et al. [79] utilized this approach in 2018, employing Stratasys J750 polymer to manufacture electromagnetically driven micropumps.

3.5. Pump Chamber and Piston of Ferrofluidic Micropumps

During the pumping process, the ferrofluidic micropump chamber is divided into inlet and outlet chambers by boundaries formed by the piston and stationary ferrofluid [76]. When the moving magnetic source aligns with the sealing magnetic source, the piston merges with the stationary ferrofluid into a single mass, eliminating the separation between the inlet and outlet chambers. As the moving magnetic source departs, the piston reforms and draws in the pumped liquid from the inlet, creating the inlet chamber between the inlet and the piston. When the pumped liquid reaches the outlet area, it stops and waits for the piston to remerge with the stationary sealing ferrofluid. Upon reforming, the area between the piston and the outlet becomes the outlet chamber, where the pumped liquid is finally expelled to complete the pumping cycle. During continuous operation, both inlet and outlet chambers coexist simultaneously, separated by the piston, with their volumes constantly changing.

3.6. Failure and Restart of Ferrofluidic Micropumps

Ferrofluidic micropumps contain no movable mechanical components, and their pumping failures can be categorized into three types: ① Piston disintegration: Due to excessive bubble presence [216] (caused by high-speed rotation or bubbles in the pumped liquid) or ferrofluid loss, the ferrofluid can no longer form a stable, compact piston that fills the channel cross-section, rendering the micropump non-functional. ② External actuation magnetic source failure, primarily occurring in electromagnets, manifested as excessive Joule heat generation and coil burnout. ③ Miscibility between the pumped liquid and ferrofluid or backflow through the boundary layer, resulting from incompatibility between the pumped liquid and selected ferrofluid type, requiring the selection of ferrofluid materials that are immiscible with the pumped liquid and exhibit higher affinity to the channel wall material.
For ferrofluidic micropumps, types ② and ③ can be avoided through proper selection of magnetic sources and ferrofluid types, while the first point indicates that a theoretically infinite lifespan is achievable by operating within appropriate speed ranges and timely ferrofluid replenishment. Below are the pumping performance data for selected ferrofluidic micropumps. Table 4 shows geometric parameters and performance characteristics of representative ferrofluidic micropumps.

3.7. Application Fields

Due to their self-sealing and non-contact characteristics, ferrofluidic micropumps and their fundamental principles demonstrate significant application potential and value across multiple domains including biomedical applications [225,226,227], point-of-care diagnostics [228,229,230], bioanalysis [231,232,233], microelectronic cooling [234,235], as well as microgravity and extreme-environment applications [122,123]. For instance, Chang et al. [218] enhanced a ferrofluidic micropump structure by incorporating four 14 mm × 7 mm annular microchannels as reaction zones, developing a medium-flow-rate microchip for immunoassays. Ahn et al. [77] fabricated a semicircular ferrofluidic micropump device using silicon for contact-free biological sample handling. Peng et al. [124] successfully addressed 10 nl level microfluid manipulation in sealed glass wafers using ferrofluidic micropump principles, with important implications for electronics, biomedicine, and chemical synthesis. Park et al. [125] integrated magnetic beads with micropumps to create a miniature permanent magnet-driven bead extraction system, enabling precise bead retention and release operations for DNA extraction processes. Hamilton et al. [126] developed magnetically actuated pumps and valves for controlling low-magnetic swimmers, with this remote manipulation technology inspiring micropump researchers to implement targeted fluid and sample control in microdomains through ferrofluid-based remote actuation in microchips, potentially enabling precise cellular and fluid manipulation at the microscale. Based on microfluidic simulation analysis, Zhou et al. [236] proposed a semi-Y-shaped labyrinth seal to reduce leakage.
Although ferrofluidic micropumps do offer several unique advantages, they also suffer from certain inherent limitations—much like the fundamental drawbacks of most mechanical micropumps—which have hindered their widespread adoption. The key issues are outlined below: 1. Biocompatibility constraints: ferrofluid materials are generally incompatible with most biological substances and can contaminate pumped bio-samples. This significantly restricts their application in biomedical and diagnostic fields. 2. Long-term loss of ferrofluid: During extended operation, the ferrofluid may gradually be carried away by the working fluid, leading to volumetric loss and eventual failure of the pump. 3. Evaporation of carrier liquid: If a ferrofluidic micropump system remains idle for prolonged periods, the base carrier liquid in the ferrofluid may evaporate, resulting in loss of function.
Despite these challenges, researchers have begun exploring strategies to facilitate the practical implementation and real-world translation of ferrofluidic micropump technologies. In 2025, Li and his team completed a series of studies, including investigations into the driving mechanisms of ferrofluidic micropumps under annular microscale constraints [199], analyses of dynamic morphological transformations in rotating and stationary ferrofluids under varying magnetic field polarities and mass fractions [216], and research on bubble formation mechanisms within ferrofluidic micropumps [237]. These studies collectively advanced the practical application of ferrofluidic micropump technology.

3.8. Performance Comparison

Obviously, for micropumps, flow rate and back pressure are two extremely important parameters. To provide a more intuitive comparison of the performance between mechanical micropumps and ferrofluidic micropumps, the author presents a table comparing their flow rates and back pressures, along with evaluations based on their parameter ranges. Due to space constraints, the following discussion only addresses the micropumps included in the table and cannot fully cover all types or their precise upper and lower limits.
As shown in Table 5, the flow rate of current ferrofluidic micropumps (i.e., FM) can range from tens of ml/min to thousands of ml/min, while the back pressure generally fluctuates around 1 kPa, with a maximum of 1.35 kPa. In contrast, piezoelectric (PE) micropumps have a lower flow rate compared to ferrofluidic micropumps but can overcome a higher maximum back pressure. Moreover, there appears to be an inverse relationship between their flow rate and back pressure, requiring users to make a trade-off between the two values. Electromagnetic (EM) micropumps have a slightly narrower range of back pressure compared to piezoelectric micropumps but overlap significantly with the back pressure range of ferrofluidic micropumps. However, their flow rate only fluctuates within a few hundred nanoliters, failing to reach even 1 milliliter. Pneumatic (PM) micropumps exhibit an exceptionally high back pressure, reaching an astonishing 68.95 kPa, which is a remarkable figure, though their flow rate is relatively small. Data on electrostatic (ES) micropumps regarding back pressure is temporarily unavailable in the table, but their extremely low flow rate is already noteworthy. Limited data is available for thermal bubble (TPM) micropumps, but it can be observed that their back pressure is higher than that of conventional ferrofluidic micropumps, while their flow rate is relatively low. From this, it can be concluded that another advantage of ferrofluidic micropumps lies in their high flow rate, enabling them to handle most microfluidic pumping tasks.

4. Conclusions

In the above content, we have broadly summarized the application areas of mechanical micropumps and classified them according to their driving principles. Subsequently, we focused on describing the structures and working principles of four specific types of micropumps: piezoelectric micropumps, known for their fast response speed and high control precision; electrostatic micropumps, which are easily miniaturized and integrated; electromagnetic micropumps, recognized for their high output force and strong driving capability; pneumatic micropumps, valued for their simple structure and bio-sample compatibility. To enrich the discussion, we referenced the work of pioneers and colleagues in academia who have studied mechanical micropumps. We also introduced the novel ferrofluidic micropump, elaborating on its advantages, including a simple structure, no moving parts, long lifespan, non-contact manipulation, and low driving voltage. Performance parameters from numerous peer research findings were provided, revealing that ferrofluidic micropumps exhibit outstanding pumping performance in the field of micropumps. However, challenges hindering the practical application of ferrofluidic micropumps were also addressed, such as incompatibility with certain samples and the tendency of magnetic fluids to evaporate. These issues are expected to be resolved in future research.
Overall, micropumps play significant roles in biomedical, scientific research, and industrial applications. With their precise fluid volume control capabilities, they hold great importance in accurate drug delivery, implantable micropumps, and microfluidic chips as core research components. Different micropump principles suit various working scenarios, demonstrating advantages of miniaturization, high integrability, and automation. As primary micropump components, mechanical micropumps have matured considerably, giving rise to various actuation principles including piezoelectric, electrostatic, electromagnetic, and pneumatic types. These have found substantial applications supporting cell culture, bioanalysis, aerospace, and other fields. However, due to inherent actuation principles, mechanical micropumps face challenges like high fabrication difficulty, large driving voltages, and short lifespans that prove extremely difficult to resolve. Ferrofluidic micropumps, as a novel non-mechanical alternative, leverage their non-contact operation, self-sealing capability, long lifespan, and restartability to enable targeted biological sample manipulation, remote detection, and leak-free microelectronic cooling in vivo or in extreme environments. Compared to traditional mechanical micropumps, they demonstrate clearer advantages by maintaining pumping efficiency while reducing failure risks, offering higher environmental safety, and representing a micropump technology worthy of focused development.

Author Contributions

Z.L. planned the project and supervised all aspects of the research. X.Z. performed all the data collection with support from B.H., analyzed data with support from Q.G., and Z.L. served as supervisors of this project, discussing all aspects of the research. X.Z. wrote the manuscript, and all authors contributed to revising it. Z.Q. is responsible for arranging the references. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 52379092), the Qinghai Talent Program (Grant No.2024YF0900003GX), and the Qinghai Provincial Science and Technology Program Project (Grant No.2025ZJ958M).

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
PDMSPolydimethylsiloxane
PMMAPolymethyl methacrylate
PZTLead zirconate titanate
MEMSMicro-Electro-Mechanical Systems
SMAShape memory alloy
EAPElectroactive polymer
CNCComputer numerical control

References

  1. Mujtaba, J.; Liu, J.; Dey, K.K.; Li, T.; Chakraborty, R.; Xu, K. Micro-Bio-Chemo-Mechanical-Systems: Micromotors, Microfluidics, and Nanozymes for Biomedical Applications. Adv. Mater. 2021, 33, 2007465. [Google Scholar] [CrossRef] [PubMed]
  2. Kant, K. Microfluidic Bio-Sensors and Their Applications. Biosensors 2023, 13, 843. [Google Scholar] [CrossRef]
  3. Filippi, M.; Buchner, T.; Yasa, O.; Weirich, S.; Katzschmann, R.K. Microfluidic Tissue Engineering and Bio-Actuation. Adv. Mater. 2022, 34, e2108427. [Google Scholar] [CrossRef] [PubMed]
  4. Peng, C.; Ju, Y.S. Finger-Powered Electro-Digital-Microfluidics. Methods Mol. Biol. 2017, 1572, 293–311. [Google Scholar]
  5. Kimura, H.; Sakai, Y.; Fujii, T. Organ/body-on-a-chip based on microfluidic technology for drug discovery. Drug Metab. Pharmacokinet. 2018, 33, 43–48. [Google Scholar] [CrossRef] [PubMed]
  6. Tien, J.; Dance, Y.W. Microfluidic Biomaterials. Adv. Healthc. Mater. 2021, 10, e2001028. [Google Scholar] [CrossRef]
  7. Mao, K.; Min, X.; Zhang, H.; Zhang, K.; Cao, H.; Guo, Y. Paper-based microfluidics for rapid diagnostics and drug delivery. J. Control Release 2020, 322, 187–199. [Google Scholar] [CrossRef]
  8. Maurya, R.; Gohil, N.; Bhattacharjee, G.; Alzahrani, K.J.; Ramakrishna, S.; Singh, V. Microfluidics device for drug discovery, screening and delivery. Prog. Mol. Biol. Transl. Sci. 2022, 187, 335–346. [Google Scholar]
  9. Kuang, S.; Singh, N.M.; Wu, Y.; Shen, Y.; Ren, W.; Tu, L. Role of microfluidics in accelerating new space missions. Biomicrofluidics 2022, 16, 021503. [Google Scholar] [CrossRef]
  10. Yeung, C.K.; Koenig, P.; Countryman, S.; Thummel, K.E.; Himmelfarb, J.; Kelly, E.J. Tissue Chips in Space-Challenges and Opportunities. Clin. Transl. Sci. 2020, 13, 8–10. [Google Scholar] [CrossRef]
  11. Sun, H.; Hu, N.; Wang, J. Application of microfluidic technology in antibody screening. Biotechnol. J. 2022, 17, e2100623. [Google Scholar] [CrossRef]
  12. He, S.; Joseph, N.; Feng, S.; Jellicoe, M.; Raston, C.L. Application of microfluidic technology in food processing. Food Funct. 2020, 11, 5726–5737. [Google Scholar] [CrossRef]
  13. Wang, Z.; Zhang, Y.; Li, Z.; Wang, H.; Li, N.; Deng, Y. Microfluidic Brain-on-a-Chip: From Key Technology to System Integration and Application. Small 2023, 19, e2304427. [Google Scholar] [CrossRef]
  14. Shi, Y.; Shao, X. Development of microfluidic technology in reproductive researches. Se Pu 2019, 37, 925–931. [Google Scholar] [CrossRef] [PubMed]
  15. Liu, Y.; Zeng, W.; Chen, R.; Ge, Y.; Zi, L.; Yang, J. Tactic movement of microalgae and its application in targeted transport: A review. Sheng Wu Gong Cheng Xue Bao 2022, 38, 578–591. [Google Scholar]
  16. Srisom, K.; Tittabutr, P.; Teaumroong, N.; Lapwong, Y.; Phatthanakun, R.; Sirivisoot, S. New method for arbuscular mycorrhizal fungus spore separation using a microfluidic device based on manual temporary flow diversion. Mycorrhiza 2020, 30, 789–796. [Google Scholar] [CrossRef]
  17. Cha, H.; Fallahi, H.; Dai, Y.; Yuan, D.; An, H.; Nguyen, N.T. Multiphysics microfluidics for cell manipulation and separation: A review. Lab Chip 2022, 22, 423–444. [Google Scholar] [CrossRef]
  18. El-Mayta, R.; Padilla, M.S.; Billingsley, M.M.; Han, X.; Mitchell, M.J. Testing the In Vitro and In Vivo Efficiency of mRNA-Lipid Nanoparticles Formulated by Microfluidic Mixing. J. Vis. Exp. 2023, e64810. [Google Scholar]
  19. Manshadi, M.K.D.; Mohammadi, M.; Monfared, L.K.; Sanati-Nezhad, A. Manipulation of micro- and nanoparticles in viscoelastic fluid flows within microfluid systems. Biotechnol. Bioeng. 2020, 117, 580–592. [Google Scholar] [CrossRef] [PubMed]
  20. Ávila, B.E.; Gao, W.; Karshalev, E.; Zhang, L.; Wang, J. Cell-Like Micromotors. Acc. Chem. Res. 2018, 51, 1901–1910. [Google Scholar]
  21. Yadav, A.S.; Galogahi, F.M.; Vashi, A.; Tran, D.T.; Kijanka, G.S.; Cha, H. Synthesis and active manipulation of magnetic liquid beads. Biomed. Microdevices 2024, 26, 24. [Google Scholar] [CrossRef]
  22. Descamps, L.; Audry, M.C.; Howard, J.; Mekkaoui, S.; Albin, C.; Barthelemy, D. Self-Assembled Permanent Micro-Magnets in a Polymer-Based Microfluidic Device for Magnetic Cell Sorting. Cells 2021, 10, 1734. [Google Scholar] [CrossRef]
  23. Lv, X.; Geng, Z.; Fan, Z.; Wang, S.; Pei, W.; Chen, H. An integrated method for cell isolation and migration on a chip. Sci. Rep. 2017, 7, 8963. [Google Scholar] [CrossRef]
  24. Ghazimirsaeed, E.; Madadelahi, M.; Dizani, M.; Shamloo, A. Secondary Flows, Mixing, and Chemical Reaction Analysis of Droplet-Based Flow inside Serpentine Microchannels with Different Cross Sections. Langmuir 2021, 37, 5118–5130. [Google Scholar] [CrossRef]
  25. Liu, M.; Li, N.; Cui, S.; Li, G.; Yang, F. Biochemical Reaction Acceleration by Electrokinetic Mixing in a Microfluidic Chip. J. Phys. Chem. Lett. 2022, 13, 5633–5637. [Google Scholar] [CrossRef]
  26. Hazra, S.; Mitra, S.; Sen, A.K. Migration and Spreading of Droplets across a Fluid-Fluid Interface in Microfluidic Coflow. Langmuir 2022, 38, 9660–9668. [Google Scholar] [CrossRef]
  27. Reis, N.M.; Needs, S.H.; Jegouic, S.M.; Gill, K.K.; Sirivisoot, S.; Howard, S. Gravity-Driven Microfluidic Siphons: Fluidic Characterization and Application to Quantitative Immunoassays. ACS Sens. 2021, 6, 4338–4348. [Google Scholar] [CrossRef] [PubMed]
  28. Kitenbergs, G.; Tatuļčenkovs, A.; Puķina, L.; Cēbers, A. Gravity effects on mixing with magnetic micro-convection in microfluidics. Eur. Phys. J. E Soft Matter 2018, 41, 138. [Google Scholar] [CrossRef] [PubMed]
  29. Pajic-Lijakovic, I.; Milivojevic, M. Marangoni effect and cell spreading. Eur. Biophys. J. 2022, 51, 419–429. [Google Scholar] [CrossRef] [PubMed]
  30. Li, D.; Guo, F.; Cui, Z.; Zhou, J.; Zhai, Y.; Du, Y. Controllable and Continuous Hollow Fiber Swimmers Based on the Marangoni Effect. ACS Appl. Mater. Interfaces 2020, 12, 53503–53509. [Google Scholar] [CrossRef]
  31. Glieberman, A.L.; Pope, B.D.; Zimmerman, J.F.; Liu, Q.; Ferrier, J.P.; Kenty, J.H.R. Synchronized stimulation and continuous insulin sensing in a microfluidic human Islet on a Chip designed for scalable manufacturing. Lab Chip 2019, 19, 2993–3010. [Google Scholar] [CrossRef] [PubMed]
  32. Payne, F.W.; Ledden, B.; Lamps, G. Capabilities of Next-Generation Patch Pump: Improved Precision, Instant Occlusion Detection, and Dual-Hormone Therapy. J. Diabetes Sci. Technol. 2019, 13, 49–54. [Google Scholar] [CrossRef]
  33. Mohith, S.; Karanth, P.N.; Kulkarni, S.M. Recent trends in mechanical micropumps and their applications: A review. Mechatronics 2019, 60, 34–55. [Google Scholar] [CrossRef]
  34. Nepomnyashchy, A. Droplet on a liquid substrate: Wetting, dewetting, dynamics, instabilities. Curr. Opin. Colloid Interface Sci. 2021, 51, 101398. [Google Scholar] [CrossRef]
  35. Salari, A.; Navi, M.; Lijnse, T.; Dalton, C. AC Electrothermal Effect in Microfluidics: A Review. Micromachines 2019, 10, 762. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, Y.N.; Fu, L.M. Micropumps and biomedical applications—A review. Microelectron. Eng. 2018, 195, 121–138. [Google Scholar] [CrossRef]
  37. Chen, P.; Chen, C.; Liu, Y.H.; Du, W.; Feng, X.J.; Liu, B.F. Fully integrated nucleic acid pretreatment, amplification, and detection on a paper chip for identifying EGFR mutations in lung cancer cells. Sens. Actuators B: Chem. 2019, 283, 472–477. [Google Scholar] [CrossRef]
  38. Yang, Y.T.; Ho, T.Y. Conquering the Tyranny of Number With Digital Microfluidics. Front. Chem. 2021, 9, 676365. [Google Scholar] [CrossRef]
  39. Farzbod, A.; Moon, H. Integration of reconfigurable potentiometric electrochemical sensors into a digital microfluidic platform. Biosens. Bioelectron. 2018, 106, 37–42. [Google Scholar] [CrossRef]
  40. Alistar, M.; Gaudenz, U. OpenDrop: An integrated do-it-yourself platform for personal use of biochips. Bioengineering 2017, 4, 45. [Google Scholar] [CrossRef]
  41. Sefton, M.V.; Lusher, H.M.; Firth, S.R.; Waher, M.U. Controlled release micropump for insulin administration. Ann. Biomed. Eng. 1979, 7, 329–343. [Google Scholar] [CrossRef] [PubMed]
  42. Senesh, G.; Bushi, D.; Neta, A.; Yodfat, O. Compatibility of insulin Lispro, Aspart, and Glulisine with the Solo MicroPump, a novel miniature insulin pump. J. Diabetes Sci. Technol. 2010, 4, 104–110. [Google Scholar] [CrossRef] [PubMed]
  43. Rimon, M.T.I.; Hasan, M.W.; Hassan, M.F.; Cesmeci, S. Advancements in Insulin Pumps: A Comprehensive Exploration of Insulin Pump Systems, Technologies, and Future Directions. Pharmaceutics 2024, 16, 944. [Google Scholar] [CrossRef]
  44. Cobo, A.; Sheybani, R.; Tu, H.; Meng, E. A Wireless Implantable Micropump for Chronic Drug Infusion Against Cancer. Sens. Actuators A Phys. 2016, 239, 18–25. [Google Scholar] [CrossRef]
  45. Kar, A.; Ahamad, N.; Dewani, M.; Awasthi, L.; Patil, R.; Banerjee, R. Wearable and implantable devices for drug delivery: Applications and challenges. Biomaterials 2022, 283, 121435. [Google Scholar] [CrossRef]
  46. Meng, E.; Hoang, T. MEMS-enabled implantable drug infusion pumps for laboratory animal research, preclinical, and clinical applications. Adv. Drug Deliv. Rev. 2012, 64, 1628–1638. [Google Scholar] [CrossRef]
  47. Dolete, G.; Chircov, C.; Motelica, L.; Ficai, D.; Oprea, O.C.; Gheorghe, M. Magneto-Mechanically Triggered Thick Films for Drug Delivery Micropumps. Nanomaterials 2022, 12, 3598. [Google Scholar] [CrossRef]
  48. Ito, T.; Ota, T.; Kono, R.; Miyaoka, Y.; Ishibashi, H.; Komori, M. Pump-Free Microfluidic Hemofiltration Device. Micromachines 2021, 12, 992. [Google Scholar] [CrossRef]
  49. Mu, X.; Chen, F.D.; Dang, K.M.; Brunk, M.G.K.; Li, J.; Wahn, H. Implantable photonic neural probes with 3D-printed microfluidics and applications touncaging. Front. Neurosci. 2023, 17, 1213265. [Google Scholar] [CrossRef]
  50. Lee, S.H.; Ahn, J.W.; Cho, Y.C.; Kim, S.N.; Lee, C.; Ku, G.W. Wirelessly Controlled Implantable System for On-demand and Pulsatile Insulin Administration. Sci. Rep. 2019, 9, 5009. [Google Scholar] [CrossRef] [PubMed]
  51. Attiguppe, A.P.; Chatterjee, D.; DasGupta, A. A Novel Integrated Transdermal Drug Delivery System with Micropump and Microneedle Made from Polymers. Micromachines 2022, 14, 71. [Google Scholar] [CrossRef]
  52. Liu, Y.; Yang, L.; Cui, Y. A wearable, rapidly manufacturable, stability-enhancing microneedle patch for closed-loop diabetes management. Microsyst. Nanoeng. 2024, 10, 112. [Google Scholar] [CrossRef] [PubMed]
  53. Shao, J.; Wu, L.; Wu, J.; Zheng, Y.; Zhao, H.; Jin, Q. Integrated microfluidic chip for endothelial cells culture and analysis exposed to a pulsatile and oscillatory shear stress. Lab Chip 2009, 9, 3118–3125. [Google Scholar] [CrossRef] [PubMed]
  54. Kumaresan, P.; Yang, C.J.; Cronier, S.A.; Blazej, R.G.; Mathies, R.A. High-throughput single copy DNA amplification and cell analysis in engineered nanoliter droplets. Anal. Chem. 2008, 80, 3522–3529. [Google Scholar] [CrossRef]
  55. Gao, T.; McNeill, J.M.; Oliver, V.A.; Xiao, L.; Mallouk, T.E. Geometric and Scaling Effects in the Speed of Catalytic Enzyme Micropumps. ACS Appl. Mater. Interfaces 2022, 14, 39515–39523. [Google Scholar] [CrossRef]
  56. Rothbauer, M.; Wartmann, D.; Charwat, V.; Ertl, P. Recent advances and future applications of microfluidic live-cell microarrays. Biotechnol. Adv. 2015, 33, 948–961. [Google Scholar] [CrossRef] [PubMed]
  57. Ou, X.; Chen, P.; Huang, X.; Li, S.; Liu, B.F. Microfluidic chip electrophoresis for biochemical analysis. J. Sep. Sci. 2020, 43, 258–270. [Google Scholar] [CrossRef]
  58. Wei, Z.; Fan, P.; Jiao, Y.; Wang, Y.; Huang, Y.; Liu, Z. Integrated microfluidic chip for on-line proteome analysis with combination of denaturing and rapid digestion of protein. Anal. Chim. Acta 2020, 1102, 1–10. [Google Scholar] [CrossRef]
  59. Guan, X.L.; Chang, D.P.S.; Mok, Z.X.; Lee, B. Assessing variations in manual pipetting: An under-investigated requirement of good laboratory practice. J. Mass Spectrom. Adv. Clin. Lab. 2023, 30, 25–29. [Google Scholar] [CrossRef]
  60. Martin, J.S.; Borges, A.R.; Beck, D.T. Peripheral conduit and resistance artery function are improved following a single, 1-h bout of peristaltic pulse external pneumatic compression. Eur. J. Appl. Physiol. 2015, 115, 2019–2029. [Google Scholar] [CrossRef]
  61. Li, Z.; Lowe, J.P.; Fletcher, P.J.; Carta, M.; McKeown, N.B.; Marken, F. Tuning and Coupling Irreversible Electroosmotic Water Flow in Ionic Diodes: Methylation of an Intrinsically Microporous Polyamine (PIM-EA-TB). ACS Appl. Mater. Interfaces 2023, 15, 42369–42377. [Google Scholar] [CrossRef]
  62. Abdelghany, A.; Yamasaki, K.; Ichikawa, Y.; Motosuke, M. Efficient nanoparticle focusing utilizing cascade AC electroosmotic flow. Electrophoresis 2022, 43, 1755–1764. [Google Scholar] [CrossRef]
  63. Qi, C.; Sugita, N.; Shinshi, T. A Disposable Electromagnetic Bi-Directional Micropump Utilizing a Rotating Multi-Pole Ring Magnetic Coupling. Micromachines 2022, 13, 1565. [Google Scholar] [CrossRef]
  64. Meng, Z.; Tayyab, M.; Lin, Z.; Raji, H.; Javanmard, M. A Smartphone-Based Disposable Hemoglobin Sensor Based on Colorimetric Analysis. Sensors 2022, 23, 394. [Google Scholar] [CrossRef] [PubMed]
  65. Ohashi, T.; Gerrett, N.; Shinkawa, S.; Sato, T.; Miyake, R.; Kondo, N. Fluidic Patch Device to Sample Sweat for Accurate Measurement of Sweat Rate and Chemical Composition: A Proof-of-Concept Study. Anal. Chem. 2020, 92, 15534–15541. [Google Scholar] [CrossRef] [PubMed]
  66. Abate, A.R.; Weitz, D.A. Syringe-vacuum microfluidics: A portable technique to create monodisperse emulsions. Biomicrofluidics 2011, 5, 014107. [Google Scholar] [CrossRef] [PubMed]
  67. Laser, D.J.; Santiago, J.G. A review of micropumps. J. Micromech. Microeng. 2004, 14, R35. [Google Scholar] [CrossRef]
  68. Mohith, S.; Karanth, N.; Kulkarni, S.M. Analysis of annularly excited bossed diaphragm for performance enhancement of mechanical micropump. Sens. Actuators A Phys. 2022, 335, 113381. [Google Scholar]
  69. Bohm, S.; Phi, H.B.; Dittrich, L.; Runge, E. Chip-integrated non-mechanical microfluidic pump driven by electrowetting on dielectrics. Lab Chip 2024, 24, 2893–2905. [Google Scholar] [CrossRef]
  70. Zhang, C.; Xing, D.; Li, Y. Micropumps, microvalves, and micromixers within PCR microfluidic chips: Advances and trends. Biotechnol. Adv. 2007, 25, 483–514. [Google Scholar] [CrossRef]
  71. Asadi Dereshgi, H.; Dal, H.; Yildiz, M.Z. Piezoelectric micropumps: State of the art review. Microsyst. Technol. 2021, 27, 4127–4155. [Google Scholar] [CrossRef]
  72. Forouzandeh, F.; Arevalo, A.; Alfadhel, A.; Borkholder, D.A. A review of peristaltic micropumps. Sens. Actuators A Phys. 2021, 326, 112602. [Google Scholar] [CrossRef] [PubMed]
  73. Gebreyesus, E.A.; Park, A.; Guldberg, R.E.; Ong, K.G. In vitromagnetohydrodynamics system for modulating cell migration. Biomed. Phys. Eng. Express 2023, 9, 025007. [Google Scholar] [CrossRef] [PubMed]
  74. Xu, C.; Wang, K.; Huang, P.; Liu, D.; Guan, Y. Single-Cell Isolation Microfluidic Chip Based on Thermal Bubble Micropump Technology. Sensors 2023, 23, 3623. [Google Scholar] [CrossRef]
  75. Wang, Y.; Li, Z.G.; Wu, D.; Li, W.X.; Li, D.C.; Zhou, X. Design and research of ferromagnetic fluid micropump with double rotating magnetic sources. Phys. Lett. A 2024, 519, 129718. [Google Scholar] [CrossRef]
  76. Ahn, J.J.; Oh, J.G.; Choi, B. A novel type of a microfluidic system using ferrofluidics for an application of μ-TAS. Microsyst. Technol. 2004, 10, 622–627. [Google Scholar] [CrossRef]
  77. Iverson, B.D.; Garimella, S.V. Recent advances in microscale pumping technologies: A review and evaluation. Microfluid. Nanofluidics 2008, 5, 145–174. [Google Scholar] [CrossRef]
  78. Amirouche, F.; Zhou, Y.; Johnson, T. Current micropump technologies and their biomedical applications. Microsyst. Technol. 2009, 15, 647–666. [Google Scholar] [CrossRef]
  79. Gusenbauer, M.; Mazza, G.; Posnicek, T.; Brandl, M.; Schrefl, T. Magnetically actuated circular displacement micropump. Int. J. Adv. Manuf. Technol. 2018, 95, 3575–3588. [Google Scholar] [CrossRef]
  80. Nguyen, T.T.; Pham, M.; Goo, N.S. Development of a peristaltic micropump for bio-medical applications based on mini LIPCA. J. Bionic Eng. 2008, 5, 135–141. [Google Scholar] [CrossRef]
  81. Bussmann, A.; Thalhofer, T.; Hoffmann, S.; Daum, L.; Surendran, N.; Hayden, O. Microfluidic Cell Transport with Piezoelectric Micro Diaphragm Pumps. Micromachines 2021, 12, 1459. [Google Scholar] [CrossRef]
  82. Zhang, Z.; Wang, W.; Chen, X. Study on the performance of piezoelectric micro pump for insulin injection. Zhongguo Yi Liao Qi Xie Za Zhi 2015, 39, 64–67. [Google Scholar]
  83. Bao, Q.; Zhang, J.; Tang, M.; Huang, Z.; Lai, L.; Huang, J. A Novel PZT Pump with Built-in Compliant Structures. Sensors 2019, 19, 1301. [Google Scholar] [CrossRef]
  84. Uhlig, S.; Gaudet, M.; Langa, S.; Schimmanz, K.; Conrad, H.; Kaiser, B. Electrostatically Driven In-Plane Silicon Micropump for Modular Configuration. Micromachines 2018, 9, 190. [Google Scholar] [CrossRef] [PubMed]
  85. Chen, H.; Miao, X.; Lu, H.; Liu, S.; Yang, Z. High-Efficiency 3D-Printed Three-Chamber Electromagnetic Peristaltic Micropump. Micromachines 2023, 14, 257. [Google Scholar] [CrossRef]
  86. So, H.; Pisano, A.P.; Seo, Y.H. Caterpillar locomotion-inspired valveless pneumatic micropump using a single teardrop-shaped elastomeric membrane. Lab Chip 2014, 14, 2240–2248. [Google Scholar] [CrossRef]
  87. Zeng, Y.; Shin, M.; Wang, T. Programmable active droplet generation enabled by integrated pneumatic micropumps. Lab Chip 2013, 13, 267–273. [Google Scholar] [CrossRef]
  88. Ameri, A.R.; Imanparast, A.; Passandideh-Fard, M.; Mousavi Shaegh, S.A. A whole-thermoplastic microfluidic chip with integrated on-chip micropump, bioreactor and oxygenator for cell culture applications. Anal. Chim. Acta 2022, 1221, 340093. [Google Scholar] [CrossRef] [PubMed]
  89. Munas, F.R.; Melroy, G.; Abeynayake, C.B.; Chathuranga, H.L.; Amarasinghe, R.; Kumarage, P.; Dau, V.T.; Dao, D.V. Development of PZT Actuated Valveless Micropump. Sensors 2018, 18, 1302. [Google Scholar] [CrossRef]
  90. Kan, J.; Xuan, M.; Yang, Z.; Wu, Y.; Wu, B.; Cheng, G. Analysis and test of piezoelectric micropump for drug delivery. Sheng Wu Yi Xue Gong Cheng Xue Za Zhi 2005, 22, 809–813. [Google Scholar] [PubMed]
  91. Johnston, I.D.; Davis, J.B.; Richter, R.; Herbert, G.I.; Tracey, M.C. Elastomer-glass micropump employing active throttles. Analyst 2004, 129, 829–834. [Google Scholar] [CrossRef] [PubMed]
  92. Amrani, I.; Cheriet, A.; Feliachi, M. Design and experimental investigation of a bi-directional valveless electromagnetic micro-pump. Sens. Actuators A Phys. 2018, 272, 310–317. [Google Scholar] [CrossRef]
  93. Amer, S.; Badawy, W. An integrated platform for bio-analysis and drug delivery. Curr. Pharm. Biotechnol. 2005, 6, 57–64. [Google Scholar] [CrossRef]
  94. Mahnama, A.; Nourbakhsh, A.; Ghorbaniasl, G. A survey on the applications of implantable micropump systems in drug delivery. Curr. Drug Deliv. 2014, 11, 123–131. [Google Scholar] [CrossRef] [PubMed]
  95. Okura, N.; Nakashoji, Y.; Koshirogane, T.; Kondo, M.; Tanaka, Y.; Inoue, K. A compact and facile microfluidic droplet creation device using a piezoelectric diaphragm micropump for droplet digital PCR platforms. Electrophoresis 2017, 38, 2666–2672. [Google Scholar] [CrossRef] [PubMed]
  96. Kotb, Y.; Elgamal, I.; Serry, M. Shape Memory Alloy Capsule Micropump for Drug Delivery Applications. Micromachines (Basel) 2021, 12, 520. [Google Scholar] [CrossRef]
  97. Lin, J.L. Design and fabrication of pneumatically actuated valveless pumps. Micromachines 2021, 13, 16. [Google Scholar] [CrossRef]
  98. Machauf, A.; Yael, N.; Uri, D. A membrane micropump electrostatically actuated across the working fluid. J. Micromech. Microeng. 2005, 15, 2309. [Google Scholar]
  99. Lee, K.S.; Kim, B.; Shannon, M.A. Development of a peristaltic gas micropump with a single chamber and multiple electrodes. J. Micromech. Microeng. 2013, 23, 095006. [Google Scholar] [CrossRef]
  100. Tahmasebipour, M.; Paknahad, A.A. Unidirectional and bidirectional valveless electromagnetic micropump with PDMS-Fe3O4 nanocomposite magnetic membrane. J. Micromech. Microeng. 2019, 29, 075014. [Google Scholar] [CrossRef]
  101. Lin, Y.Q.; Yang, X.L.; Jin, T.; Wang, J.Y.; Yi, S.C.; Wang, Y. A Compact, High-Performance, and Deformation-Resilient Trielectrode Electrostatic Soft Pump for Soft Robotics. Adv. Intell. Syst. 2025, 7, 2400423. [Google Scholar] [CrossRef]
  102. Nicola, B.A.; Popescu, M.N.; Gáspár, S. Substrate-Controlled Bidirectional Pumping by a Bienzymatic Micropump. ACS Appl. Mater. Interfaces 2024, 16, 59556–59566. [Google Scholar] [CrossRef]
  103. Werner, E.M.; Lam, B.X.; Hui, E.E. Phase-Optimized Peristaltic Pumping by Integrated Microfluidic Logic. Micromachines 2022, 13, 1784. [Google Scholar] [CrossRef] [PubMed]
  104. Bui, G.T.; Wang, J.H.; Lin, J.L. Optimization of Micropump Performance Utilizing a Single Membrane with an Active Check Valve. Micromachines 2017, 9, 1. [Google Scholar] [CrossRef] [PubMed]
  105. Ni, J.; Li, B.; Yang, J. A pneumatic PDMS micropump with in-plane check valves for disposable microfluidic systems. Microelectron. Eng. 2012, 99, 28–32. [Google Scholar] [CrossRef]
  106. Al-Halhouli, A.T.; Demming, S.; Dietzel, A.; Büttgenbach, S. Design, Fabrication, and Characterization of a Continuous Flow Micropump System. ASME J. Therm. Sci. Eng. Appl. 2016, 8, 021006. [Google Scholar] [CrossRef]
  107. Xi, H.D.; Zheng, H.; Guo, W.; Gañán-Calvo, A.M.; Ai, Y.; Tsao, C.W. Active droplet sorting in microfluidics: A review. Lab Chip 2017, 17, 751–771. [Google Scholar] [CrossRef]
  108. Agnihotri, S.N.; Raveshi, M.R.; Nosrati, R.; Bhardwaj, R.; Neild, A. Droplet splitting in microfluidics: A review. Phys. Fluids 2025, 37, 051304. [Google Scholar] [CrossRef]
  109. Abate, A.R.; Weitz, D.A. Single-layer membrane valves for elastomeric microfluidic devices. Appl. Phys. Lett. 2008, 92, 243509. [Google Scholar] [CrossRef]
  110. Abate, A.R.; Agresti, J.J.; Weitz, D.A. Microfluidic sorting with high-speed single-layer membrane valves. Appl. Phys. Lett. 2010, 96, 203509. [Google Scholar] [CrossRef]
  111. Agnihotri, S.N.; Raveshi, M.R.; Bhardwaj, R.; Neild, A. Microvalves for integrated selective droplet generation, splitting and merging on a chip. Microfluid. Nanofluidics 2021, 25, 88. [Google Scholar] [CrossRef]
  112. Nguyen, Q.M.; Abouezzi, J.; Ristroph, L. Early turbulence and pulsatile flows enhance diodicity of Tesla’s macrofluidic valve. Nat. Commun. 2021, 12, 2884. [Google Scholar] [CrossRef] [PubMed]
  113. Liu, B.D.; Zhang, Z.; Yang, J.; Yang, J.; Li, D. An electromagnetically actuated tube type micropump with a one-way valve. Int. J. Appl. Electromagn. Mech. 2018, 56, 1–9. [Google Scholar] [CrossRef]
  114. Cesmeci, S.; Hassan, R.; Baniasadi, M. A Comparative Evaluation of Magnetorheological Micropump Designs. Micromachines 2022, 13, 764. [Google Scholar] [CrossRef]
  115. Jie, M.; Qi, Z.X.; Yu, W.X.; Ma, T.F.; Zhao, Y.J.; Cai, L.T. Design and Performance Test of Four-Chamber Series–Parallel Piezoelectric Pump. Coatings 2024, 14, 1348. [Google Scholar] [CrossRef]
  116. Grover, W.H.; Skelley, A.M.; Liu, C.N.; Lagally, E.T.; Mathies, R.A. Monolithic membrane valves and diaphragm pumps for practical large-scale integration into glass microfluidic devices. Sens. Actuators B Chem. 2003, 89, 315–323. [Google Scholar] [CrossRef]
  117. Raveshi, M.R.; Agnihotri, S.N.; Sesen, M.; Bhardwaj, R.; Neild, A. Selective droplet splitting using single layer microfluidic valves. Sens. Actuators B Chem. 2019, 292, 233–240. [Google Scholar] [CrossRef]
  118. Roy, E.; Galas, J.C.; Veres, T. Thermoplastic elastomers for microfluidics: Towards a high-throughput fabrication method of multilayered microfluidic devices. Lab Chip 2011, 11, 3193–3196. [Google Scholar] [CrossRef]
  119. Mohan, R.; Schudel, B.R.; Desai, A.V.; Yearsley, J.D.; Apblett, C.A.; Kenisl, P.J.A. Design considerations for elastomeric normally closed microfluidic valves. Sens. Actuators B Chem. 2011, 160, 1216–1223. [Google Scholar] [CrossRef]
  120. Vashi, A.; Sreejith, K.R.; Nguyen, N.T. Lab-on-a-Chip Technologies for Microgravity Simulation and Space Applications. Micromachines 2022, 14, 116. [Google Scholar] [CrossRef] [PubMed]
  121. Kim, S.; Ayan, B.; Shayan, M.; Rando, T.A.; Huang, N.F. Skeletal muscle-on-a-chip in microgravity as a platform for regeneration modeling and drug screening. Stem Cell Rep. 2024, 19, 1061–1073. [Google Scholar] [CrossRef]
  122. Krakos, A. Lab-on-chip technologies for space research—Current trends and prospects. Mikrochim. Acta 2023, 191, 31. [Google Scholar] [CrossRef]
  123. Shi, J.J.; Wu, D.; Liu, T.Z.; Hao, S.J.; Meng, B.C.; Li, S.L. Comparative of Forensic DNA Identification Using Cell Lysis Method and Magnetic Beads Method. Fa Yi Xue Za Zhi 2023, 39, 45–49, (In English and Chinese). [Google Scholar]
  124. Peng, X.Y.L.; Peng, L.; Guo, Y. Manipulating nanoliter fluid circuits on an all-glass chip by the magnetic field. iScience 2023, 26, 107659. [Google Scholar] [CrossRef]
  125. Park, C.Y.; Park, Y.H.; Kim, Y.S.; Song, H.J.; Kim, J.D. Permanent magnet actuation for magnetic bead-based DNA extraction. Biomed. Eng. Online 2018, 17 (Suppl. S2), 143. [Google Scholar] [CrossRef] [PubMed]
  126. Hamilton, J.K.; Bryan, M.T.; Gilbert, A.D.; Ogrin, F.Y.; Myers, T.O. A new class of magnetically actuated pumps and valves for microfluidic applications. Sci. Rep. 2018, 8, 933. [Google Scholar] [CrossRef]
  127. Yuan, S.Q.; Yang, S.; He, X.H.; Deng, Z.D.; Cai, S.C. Design and experimental study of a novel three-way diffuser/nozzle elements employed in valveless piezoelectric micropumps. J. Braz. Soc. Mech. Sci. Eng. 2015, 37, 221–222. [Google Scholar] [CrossRef]
  128. Zhu, C.L.; Shu, X.L.; Liu, D.C.; Du, X.H.; Li, L.X.; Pan, Q.S. Resonant-Type Piezoelectric Pump Driven by Piezoelectric Stacks and a Rhombic Micro Displacement Amplifier. Micromachines 2023, 14, 1764. [Google Scholar] [CrossRef] [PubMed]
  129. Bouzid, A.; Negassa, M.E.A.B.; Pirastesh, P.; Charrier, J.; Soulimane, S. Figure of merit study for valveless piezoelectric micropump designed to handle liquid and gas fluid in medical and environmental applications. J. Serbian Soc. For. Comput. Mech. 2024, 18, 29–44. [Google Scholar] [CrossRef]
  130. Guan, Y.F.; Meng, X.X.; Liu, Y.S.; Bai, M.Y.; Xu, F.Q. Modeling, Vibration Analysis and Fabrication of Micropumps Based on Piezoelectric Transducers. Int. J. Acoust. Vib. 2020, 25, 383–391. [Google Scholar] [CrossRef]
  131. Gidde, R.R.; Pawar, P.M.; Ronge, B.P.; Dhamgaye, V.P. Design optimization of an electromagnetic actuation based valveless micropump for drug delivery application. Microsyst. Technol. 2019, 25, 509–519. [Google Scholar] [CrossRef]
  132. Huang, Y.F.; Tsou, C.H.; Hsu, C.J.; Lin, Y.C.; Ono, T.; Tsai, Y.C. Metallic glass thin film integrated with flexible membrane for electromagnetic micropump application. Jpn. J. Appl. Phys. 2020, 59, SIIK03. [Google Scholar] [CrossRef]
  133. Gidde, R.R.; Pawar, P.M.; Dhamgaye, V.P. Fully coupled modeling and design of a piezoelectric actuation based valveless micropump for drug delivery application. Microsyst. Technol. 2020, 26, 633–645. [Google Scholar] [CrossRef]
  134. Aggarwal, R.; Agarwal, M.P.; Avasthi, R. Indian Journal of Endocrinology and Metabolism; ESICON 2016 Abstracts; National Institutes of Health: Bethesda, MD, USA, 2017; Volume 21, pp. S1–S90. [Google Scholar]
  135. Anheuer, D.; Karacan, B.; Herzog, L.; Weigel, N.; Meyer-Nieberg, S.; Gebhardt, T. Framework for Microdosing Odors in Virtual Reality for Psychophysiological Stress Training. Sensors 2024, 24, 7046. [Google Scholar] [CrossRef] [PubMed]
  136. Tjulkins, F.; Sebastian, R.; Guillerm, T.; Clover, A.J.P.; Hu, Y.; Lyness, A. Towards Micropump- and Microneedle-based Drug Delivery using Micro Transdermal Interface Platforms (MicroTIPs). Annu. Int. Conf. IEEE Eng. Med. Biol. Soc. 2022, 2022, 3020–3023. [Google Scholar]
  137. Plano, D.; Kibler, S.; Rudolph, N.; Zett, O.; Dressman, J. Silicon-Based Piezo Micropumps Enable Fully Flexible Drug Delivery Patterns. J. Pharm. Sci. 2024, 113, 1555–1565. [Google Scholar] [CrossRef] [PubMed]
  138. Silverio, V.; Canane, P.A.G.; Martins, T.A.; Afonso, R.; Cardoso, S.; Batista, E. Development of a microfluidic electroosmosis pump on a chip for steady and continuous fluid delivery. Biomed. Tech. 2022, 68, 79–90. [Google Scholar] [CrossRef]
  139. Holman, J.B.; Zhu, X.; Cheng, H. Piezoelectric micropump with integrated elastomeric check valves: Design, performance characterization and primary application for 3D cell culture. Biomed. Microdevices 2023, 25, 5. [Google Scholar] [CrossRef]
  140. Zoupanou, S.; Chiriacò, M.S.; Tarantini, I.; Ferrara, F. Innovative 3D Microfluidic Tools for On-Chip Fluids and Particles Manipulation: From Design to Experimental Validation. Micromachines 2021, 12, 104. [Google Scholar] [CrossRef]
  141. Lin, L.; Chung, C.K. PDMS Microfabrication and Design for Microfluidics and Sustainable Energy Application: Review. Micromachines 2021, 12, 1350. [Google Scholar] [CrossRef]
  142. Du, Z.; Sun, W.; Mi, S. Multi-step PDMS curing and a controlled separation method for mass manufacturing of high-performance and user-friendly micro-devices: Valved micropumps. Lab Chip 2024, 24, 843–853. [Google Scholar] [CrossRef]
  143. Akram, M.M.; Aghajani, A.; Shi, W.; Gosselin, B. A closed-loop micropump and flow meter for high-precision drug delivery in an implantable neural probe. In Proceedings of the 2024 46th Annual International Conference of the IEEE Engineering in Medicine and Biology Society (EMBC), Orlando, FL, USA, 15–19 July 2024; pp. 1–4. [Google Scholar]
  144. Ahmadianyazdi, A.; Miller, I.J.; Folch, A. Tunable resins with PDMS-like elastic modulus for stereolithographic 3D-printing of multimaterial microfluidic actuators. Lab Chip 2023, 23, 4019–4032. [Google Scholar] [CrossRef]
  145. Spencer, W.J.; Corbett, W.T.; Dominguez, L.R.; Shafer, B. D, An electronically controlled piezoelectric insulin pump and valves. IEEE Trans. Sonics Ultrason. 1978, 25, 153–156. [Google Scholar] [CrossRef]
  146. Calvo-López, A.; Ymbern, O.; Puyol, M.; Alonso-Chamarro, J. Soluble reactive phosphorous determination in wastewater treatment plants by automatic microanalyzers. Talanta 2021, 221, 121508. [Google Scholar] [CrossRef] [PubMed]
  147. Dhwaj, A.; Roy, N.; Jaiswar, A.; Prabhakar, A.; Verma, D. 3D-Printed Impedance Micropump for Continuous Perfusion of the Sample and Nutrient Medium Integrated with a Liver-On-Chip Prototype. ACS Omega 2022, 7, 40900–40910. [Google Scholar] [CrossRef]
  148. Kim, J.; Kim, S.; Choi, J.; Koo, C. A 3D Miniaturized Glass Magnetic-Active Centrifugal Micropump Fabricated by SLE Process and Laser Welding. Micromachines 2022, 13, 1331. [Google Scholar] [CrossRef]
  149. Alexandre-Franco, M.F.; Kouider, R.; Kassir Al-Karany, R.; Cuerda-Correa, E.M.; Al-Kassir, A. Recent Advances in Polymer Science and Fabrication Processes for Enhanced Microfluidic Applications: An Overview. Micromachines 2024, 15, 1137. [Google Scholar] [CrossRef] [PubMed]
  150. Li, Y.; Zhang, H.; Li, Q.; Deng, Y.; Ye, Z.; Gui, L. Texture-structure-based liquid metal filling for blind-end microchannels and its application on multi-layer chips. RSC Adv. 2023, 13, 24228–24236. [Google Scholar] [CrossRef] [PubMed]
  151. Ni, J.; Huang, F.; Wang, B.; Li, B.; Lin, Q. A planar PDMS micropump using in-contact minimized-leakage check valves. J. Micromech. Microeng. 2010, 20, 095033. [Google Scholar] [CrossRef]
  152. Regmi, S.; Poudel, C.; Adhikari, R.; Luo, K.Q. Applications of Microfluidics and Organ-on-a-Chip in Cancer Research. Biosensors 2022, 12, 459. [Google Scholar] [CrossRef]
  153. Lan, Z.; Chen, R.; Zou, D.; Zhao, C.X. Microfluidic Nanoparticle Separation for Precision Medicine. Adv. Sci. 2025, 12, e2411278. [Google Scholar] [CrossRef]
  154. Abu-Dawas, S.; Alawami, H.; Zourob, M.; Ramadan, Q. Design and Fabrication of Low-Cost Microfluidic Chips and Microfluidic Routing System for Reconfigurable Multi-(Organ-on-a-Chip) Assembly. Micromachines 2021, 12, 1542. [Google Scholar] [CrossRef]
  155. Vante, A.B.; Kanish, T.C. Fluid-structure interaction and experimental studies of passive check valve based piezoelectric micropump for biomedical applications. Adv. Mater. Process. Technol. 2024, 10, 2095–2121. [Google Scholar] [CrossRef]
  156. Liu, X.P.; Li, X.Q.; Wang, M.; Cao, S.Q.; Wang, X.F.; Liu, G.J. A High-Performance Piezoelectric Micropump with Multi-Chamber in Series. Appl. Sci. 2022, 12, 4483. [Google Scholar] [CrossRef]
  157. Lv, W.C.; Ni, J.F.; Xuan, W.P.; Li, Y.X.; Huang, X.Y.; Sun, L.L. A simulation and experimental study of a valveless piezoelectric micropump based on the synthetic jet principle. Nanotechnol. Precis. Eng. 2025, 8, 023009. [Google Scholar] [CrossRef]
  158. Chuang, W.C.; Lee, H.L.; Chang, P.Z.; Hu, Y.C. Review on the modeling of electrostatic MEMS. Sensors 2010, 10, 6149–6171. [Google Scholar] [CrossRef]
  159. Surendran, N.; Durasiewicz, C.P.; Hoffmann, T.; Wille, A.; Bussmann, A.B.; Richter, M. Microfluidic Delivery of High Viscosity Liquids Using Piezoelectric Micropumps for Subcutaneous Drug Infusion Applications. IEEE Open J. Eng. Med. Biol. 2024, 5, 21–31. [Google Scholar] [CrossRef]
  160. Bussmann, A.B.; Grünerbel, L.M.; Durasiewicz, C.P.; Thalhofer, T.A.; Wille, A.; Richter, M. Microdosing for drug delivery application—A review. Sens. Actuators A Phys. 2021, 330, 112820. [Google Scholar] [CrossRef]
  161. Ochoa, M.; Ziaie, B. A fermentation-powered thermopneumatic pump for biomedical applications. Lab Chip 2012, 12, 4044–4048. [Google Scholar] [CrossRef] [PubMed]
  162. Li, Y.; Zhu, K.; Liu, X.; Zhang, Y.S. Blood-Vessel-on-a-Chip Platforms for Evaluating Nanoparticle Drug Delivery Systems. Curr. Drug Metab. 2018, 19, 100–109. [Google Scholar] [CrossRef] [PubMed]
  163. Liu, G.; Lu, Y.; Zhang, F.; Liu, Q. Electronically powered drug delivery devices: Considerations and challenges. Expert Opin. Drug Deliv. 2022, 19, 1636–1649. [Google Scholar] [CrossRef]
  164. You, R.; Fu, X.; Duan, X. Acoustofluidic Based Wireless Micropump for Portable Drug Delivery Applications. Annu. Int. Conf. IEEE Eng. Med. Biol. Soc. 2021, 2021, 1276–1279. [Google Scholar]
  165. Rasekh, M.; Harrison, S.; Schobesberger, S.; Ertl, P.; Balachandran, W. Reagent storage and delivery on integrated microfluidic chips for point-of-care diagnostics. Biomed. Microdevices 2024, 26, 28. [Google Scholar] [CrossRef] [PubMed]
  166. Wang, Q.; Liu, J.; Yin, W.; Wang, A.; Zheng, J.; Wang, Y. Microscale tissue engineering of liver lobule models: Advancements and applications. Front. Bioeng. Biotechnol. 2023, 11, 1303053. [Google Scholar] [CrossRef]
  167. Sarkar, T.; Nguyen, T.; Moinuddin, S.M.; Stenmark, K.R.; Nozik, E.S.; Saha, D. A Protocol for Fabrication and on-Chip Cell Culture to Recreate PAH-Afflicted Pulmonary Artery on a Microfluidic Device. Micromachines 2022, 13, 1483. [Google Scholar] [CrossRef]
  168. Al-Hilal, T.A.; Keshavarz, A.; Kadry, H.; Lahooti, B.; Al-Obaida, A.; Ding, Z. Pulmonary-arterial-hypertension (PAH)-on-a-chip: Fabrication, validation and application. Lab Chip 2020, 21, 3334–3345. [Google Scholar] [CrossRef] [PubMed]
  169. Shinha, K.; Nihei, W.; Nakamura, H.; Goto, T.; Kawanishi, T.; Ishida, N. A Kinetic Pump Integrated Microfluidic Plate (KIM-Plate) with High Usability for Cell Culture-Based Multiorgan Microphysiological Systems. Micromachines 2021, 12, 1007. [Google Scholar] [CrossRef] [PubMed]
  170. Noh, K.H.; Kang, H.M.; Oh, S.J.; Lee, J.Y.; Kim, D.H.; Kim, M. A new experimental model to study human drug responses. Biofabrication 2020, 12, 045029. [Google Scholar] [CrossRef]
  171. Kim, S.; Kim, J.; Joung, Y.H.; Ahn, S.; Park, C.; Choi, J. Monolithic 3D micromixer with an impeller for glass microfluidic systems. Lab Chip 2020, 20, 4474–4485. [Google Scholar] [CrossRef]
  172. Kalinowski, S.; Kościelniak, P.; Wierzbicka, E.; Koronkiewicz, S. A Multi-Pumping Gradient Calibration Module for Potentiometric Determination of Nitrate in Surface Water. Molecules 2023, 28, 493. [Google Scholar] [CrossRef]
  173. Park, J.H.; Seo, M.Y.; Ham, Y.B.; Yun, S.N.; Kim, D.I. A study on high-output piezoelectric micropumps for application in DMFC. J. Electroceram. 2013, 30, 102–107. [Google Scholar] [CrossRef]
  174. Michael, I.J.; Kumar, S.; Oh, J.M.; Kim, D.; Kim, J.; Cho, Y.K. Surface-Engineered Paper Hanging Drop Chip for 3D Spheroid Culture and Analysis. ACS Appl. Mater. Interfaces 2018, 10, 33839–33846. [Google Scholar] [CrossRef]
  175. Wu, Q.; Liu, J.; Wang, X.; Feng, L.; Wu, J.; Zhu, X. Organ-on-a-chip: Recent breakthroughs and future prospects. Biomed. Eng. Online 2020, 19, 9. [Google Scholar] [CrossRef]
  176. Srivastava, S.K.; Foo, G.W.; Aggarwal, N.; Chang, M.W. Organ-on-chip technology: Opportunities and challenges. Biotechnol. Notes 2024, 5, 8–12. [Google Scholar] [CrossRef] [PubMed]
  177. Chliara, M.A.; Elezoglou, S.; Zergioti, I. Bioprinting on Organ-on-Chip: Development and Applications. Biosensors 2022, 12, 1135. [Google Scholar] [CrossRef] [PubMed]
  178. Zhao, Y.; Lv, X.; Li, X.; Rcheulishvili, N.; Chen, Y.; Li, Z. Microfluidic actuated and controlled systems and application for lab-on-chip in space life science. Space Sci. Technol. 2023, 3, 0008. [Google Scholar] [CrossRef]
  179. Thorsen, T.; Maerkl, S.J.; Quake, S.R. Microfluidic large-scale integration. Science 2002, 298, 580–584. [Google Scholar] [CrossRef]
  180. Khandelwal, A.; Athreya, N.; Tu, M.Q.; Janavicius, L.L.; Yang, Z.; Milenkovic, O. Self-assembled microtubular electrodes for on-chip low-voltage electrophoretic manipulation of charged particles and macromolecules. Microsyst. Nanoeng. 2022, 8, 27. [Google Scholar] [CrossRef]
  181. Prakash, J.; Dharmendra, T.; Bég, O.A. Comparative study of hybrid nanofluids in microchannel slip flow induced by electroosmosis and peristalsis. Appl. Nanosci. 2020, 10, 1693–1706. [Google Scholar] [CrossRef]
  182. Xu, H.; Hang, Y.; Wu, Z.; Lei, X.; Deng, J.; Yang, J. Capillary-driven microchip integrated with nickel phosphide hybrid-modified electrode for the electrochemical detection of glucose. Anal. Chim. Acta 2024, 1316, 342882. [Google Scholar] [CrossRef]
  183. Wu, D.; Shi, B.; Li, B.; Wu, W. A Novel Self-Activated Mechanism for Stable Liquid Transportation Capable of Continuous-Flow and Real-time Microfluidic PCRs. Micromachines 2019, 10, 350. [Google Scholar] [CrossRef]
  184. Ma, H.K.; Chen, R.H.; Hsu, Y.H. Development of a piezoelectric-driven miniature pump for biomedical applications. Sens. Actuators A Phys. 2015, 234, 23–33. [Google Scholar] [CrossRef]
  185. Liu, B.D.; Ma, C.X.; Yang, J.H.; Li, D.S.; Liu, H.B. Study on the Heat Source Insulation of a Thermal Bubble-Driven Micropump with Induction Heating. Micromachines 2021, 12, 1040. [Google Scholar] [CrossRef] [PubMed]
  186. Yang, Z.G.; Dong, L.T.; Wang, M.; Li, X.Q.; Liu, X.P.; Liu, G.J. A miniature piezoelectric pump with high performance. AIP Adv. 2022, 12, 065316. [Google Scholar] [CrossRef]
  187. Motohashi, T.; Ogawa, N.; Akai, H.; Shintake, J. Peristaltic micropump using polyvinyl chloride gels with micropatterned surface. Sci. Rep. 2022, 12, 22608. [Google Scholar] [CrossRef] [PubMed]
  188. Huang, J.; Affane, H.; Zhang, B.; Kuang, M.; Xiong, J.; Zhang, S.Y. Investigation of Flow Characteristics in Valveless Piezoelectric Pumps with Airfoil Baffles at Varying Angles of Attack. Appl. Sci. 2025, 15, 445. [Google Scholar] [CrossRef]
  189. Dong, J.S.; Cao, Y.; Chen, Q.Q.; Wu, Y.; Liu, R.G.; Liu, W.S. Performance of single piezoelectric vibrator micropump with check valve. J. Intell. Mater. Syst. Struct. 2020, 31, 117–126. [Google Scholar] [CrossRef]
  190. Rusli, M.Q.A.; Chee, P.S.; Arsat, R.; Lau, K.X.; Leow, P.L. Electromagnetic actuation dual-chamber bidirectional flow micropump. Sens. Actuators A Phys. 2018, 282, 17–27. [Google Scholar] [CrossRef]
  191. Chen, S.; Xie, Y.L.; Fang, X.; Zou, Z.M.; Lu, J.; Zhang, Z.H. A piezoelectric pump with composite vibrator for bubble resistance. Sens. Actuators A Phys. 2024, 379, 115972. [Google Scholar] [CrossRef]
  192. Cao, D.; Ji, S.; Yang, Z. Design and analysis of a wheel valve piezoelectric micropump with high performance. Sens. Actuators A Phys. 2024, 377, 115624. [Google Scholar] [CrossRef]
  193. Pan, Q.; Jiang, H.; Huang, Z.; Huang, B.; Li, R.; Wana, B. Development of a piezoelectric pump with ball valve structure. J. Intell. Mater. Syst. Struct. 2021, 32, 2289–2299. [Google Scholar] [CrossRef]
  194. Ye, Y.; Chen, J.; Pan, Q.S.; Feng, Z.H. Suppressing the generation of cavitation by increasing the number of inlet check valves in piezoelectric pumps. Sens. Actuators A Phys. 2019, 293, 56–61. [Google Scholar] [CrossRef]
  195. Wang, J.L.; Zhang, F.; Gui, Z.Z.; Wen, Y.X.; Zeng, Y.H.; Xie, T. Design and Analysis of a Cardioid Flow Tube Valveless Piezoelectric Pump for Medical Applications. Sensors 2024, 24, 122. [Google Scholar] [CrossRef]
  196. Dereshgi, H.A.; Yıldız, M.; Parlak, N. Performance comparison of novel single and bi-diaphragm PZT based valveless micropumps. J. Appl. Fluid Mech. 2020, 13, 401–412. [Google Scholar] [CrossRef]
  197. Stephen, P.S. Low Viscosity Magnetic Fluid Obtained by the Colloidal Suspension of Magnetic Particles. U.S. Patent US-PATENT-3,215,572, 1965. [Google Scholar]
  198. Yang, W.R.; Zhang, Y.M.; Yang, X.R.; Sun, C.X.; Chen, Y. Systematic analysis of ferrofluidic: A visualization review, advances engineering applications, and challenges. J. Nanopart. Res. 2022, 24, 102. [Google Scholar] [CrossRef]
  199. Li, W.; Li, Z.; Han, W.; Li, D.; Yan, S.; Zhou, J. Study of the flow characteristics of pumped media in the confined morphology of a ferrofluidic pump with annular microscale constraints. J. Fluids Eng. 2025, 147, 021201. [Google Scholar] [CrossRef]
  200. Pamme, N. Magnetism and microfluidics. Lab Chip 2006, 6, 24–38. [Google Scholar] [CrossRef]
  201. Kim, E.G.; Oh, J.; Choi, B. A study on the development of a continuous peristaltic micropump using magnetic fluids. Sens. Actuators A Phys. 2006, 128, 43–51. [Google Scholar] [CrossRef]
  202. Yunas, J.; Mulyanti, B.; Hamidah, I.; Mohd Said, M.; Pawinanto, R.E.; Wan Ali, W.A.F. Polymer-based MEMS electromagnetic actuator for biomedical application: A review. Polymers 2020, 12, 1184. [Google Scholar] [CrossRef] [PubMed]
  203. Rathore, A.S. Joule heating and determination of temperature in capillary electrophoresis and capillary electrochromatography columns. J. Chromatogr. A 2004, 1037, 431–443. [Google Scholar] [CrossRef]
  204. Wang, Y.; Li, Z.G.; Li, D.C.; Chen, F.; Zhao, Q.; Qing, J. A Double-Rotating ferrofluidic Vane Micropump with an Embedded Fixed Magnet. Actuators 2024, 13, 308. [Google Scholar] [CrossRef]
  205. Hatch, A.; Kamholz, A.E.; Holman, G.; Yager, P.; Bohringer, K.F. A ferrofluidicic magnetic micropump. J. Microelectromech. systems. 2001, 10, 215–221. [Google Scholar] [CrossRef]
  206. Wang, H.; Gao, Z.; Meng, Y.; Li, Y.; Zhang, Q. Performance of the ferrofluidic-spiral labyrinth combined seal for sealing liquid. Int. J. Appl. Electromagn. Mech. 2023, 72, 87–101. [Google Scholar] [CrossRef]
  207. Liu, Y.; Yang, X.; Dou, X.; Han, L. Experimental study of symmetrical ferrofluidic seals with single magnetic source and small clearance. Tribol. Int. 2024, 193, 109467. [Google Scholar] [CrossRef]
  208. Liu, B.D.; Zhang, Z.; Yang, J.; Yang, J.; Li, D. A rotary ferrofluidicic vane micropump with C shape baffle. Sens. Actuators B Chem. 2018, 263, 452–458. [Google Scholar] [CrossRef]
  209. Zhu, G.; Li, Z.Q.; Li, W.; Liu, S.X.; Wang, Z.C.; Cheng, C. Design and Dynamic Characteristics Research of Fishmouth Shaped Baffle ferrofluidic Micro-Pump. Iran. J. Sci. Technol. Trans. Mech. Eng. 2025, 49, 1205–1214. [Google Scholar] [CrossRef]
  210. Tang, Z.; Shao, X.; Huang, J.; Yao, J.; Ding, G. Manipulate microfluid with an integrated butterfly valve for micropump application. Sens. Actuators A Phys. 2020, 306, 111965. [Google Scholar] [CrossRef]
  211. Yu, A.; Wang, Y.F.; Lv, S.J.; Tang, Q.H. Numerical analysis of the cavity vorticity transport and entropy production in a micropump. Int. Commun. Heat Mass Transf. 2024, 159, 108144. [Google Scholar] [CrossRef]
  212. Shaker, S.; Hajjawi, M.; Khan, A.; Kilani, M. Effect of inlet and outlet angles on the flow performance of the ferrofluidicic magnetic micropump. Cogent Eng. 2023, 10, 2158611. [Google Scholar] [CrossRef]
  213. Nguyen, N.T.; Wu, Z. Micromixers—A review. J. Micromech. Microeng. 2004, 15, R1. [Google Scholar] [CrossRef]
  214. Unger, M.A.; Chou, H.P.; Thorsen, T.; Scherer, A.; Quake, S.R. Monolithic microfabricated valves and pumps by multilayer soft lithography. Science 2000, 288, 113–116. [Google Scholar] [CrossRef]
  215. Hartshorne, H.; Backhouse, C.J.; Lee, W.E. ferrofluidic-based microchip pump and valve. Sens. Actuators B Chem. 2004, 99, 592–600. [Google Scholar] [CrossRef]
  216. Li, W.; Li, Z.; Han, W.; Wang, Y.; Zhao, J.; Zhou, J. Morphologic transformation of ferrofluidic during micropump driving under field control. Ann. N. Y. Acad. Sci. 2025, 1543, 194–203. [Google Scholar] [CrossRef] [PubMed]
  217. Doganay, S.; Cetin, L.; Ezan, M.A.; Turgut, A.A. rotating permanent magnetic actuator for micropump devices with magnetic nanofluids. J. Micromech. Microeng. 2020, 30, 075012. [Google Scholar] [CrossRef]
  218. Chang, Y.J.; Hu, C.Y.; Lin, H.S. A microchannel immunoassay chip with ferrofluidic actuation to enhance the biochemical reaction. Sens. Actuators B Chem. 2013, 182, 584–591. [Google Scholar] [CrossRef]
  219. Fu, L.M.; Fang, W.C.; Hong, T.F.; Lee, C.Y. A magnetic micropump based on ferrofluidicic actuation. Int. J. Autom. Smart Technol. 2014, 4, 77–82. [Google Scholar] [CrossRef]
  220. Ashouri, M.; Shafii, M.B.; Moosavi, A.; Hezave, H.A. A novel revolving piston minipump. Sens. Actuators B Chem. 2015, 218, 237–244. [Google Scholar] [CrossRef]
  221. Ashouri, M.; Shafii, M.B.; Moosavi, A. Theoretical and experimental studies of a magnetically actuated valveless micropump. J. Micromech. Microeng. 2016, 27, 015016. [Google Scholar] [CrossRef]
  222. Lee, C.Y.; Leong, J.C.; Wang, Y.N.; Fu, L.M. A ferrofluidicic magnetic micropump for variable-flow-rate applications. Jpn. J. Appl. Phys. 2012, 51, 047201. [Google Scholar] [CrossRef]
  223. Ando, B.; Ascia, A.; Baglio, S.; Pitrone, N. Development of novel ferrofluidicic pumps. In Proceedings of the 2006 International Conference of the IEEE Engineering in Medicine and Biology Society, New York, NY, USA, 30 August–3 September 2006; pp. 2828–2831. [Google Scholar]
  224. Yamahata, C.; Chastellain, M.; Parashar, V.K.; Petri, A.; Hofmann, H.; Gijs, M.A. Plastic micropump with ferrofluidicic actuation. J. Microelectromech. Syst. 2005, 14, 96–102. [Google Scholar] [CrossRef]
  225. Kim, H.; Ali, A.; Kang, Y.; Lim, B.; Kim, C. Surface-Driven Particle Dynamics: Sequential Synchronization of Colloidal Flow Attempted in a Static Fluidic Environment. ACS Appl. Mater. Interfaces 2025, 17, 12772–12781. [Google Scholar] [CrossRef] [PubMed]
  226. Molinero-Fernández, Á.; Moreno-Guzmán, M.; López, M.Á.; Escarpa, A. Magnetic Bead-Based Electrochemical Immunoassays On-Drop and On-Chip for Procalcitonin Determination: Disposable Tools for Clinical Sepsis Diagnosis. Biosensors 2020, 10, 66. [Google Scholar] [CrossRef] [PubMed]
  227. Li, Z.; Zu, X.Y.; Du, Z.; Hu, Z.G. Research on magnetic bead motion characteristics based on magnetic beads preset technology. Sci. Rep. 2021, 11, 19995. [Google Scholar] [CrossRef]
  228. Zhou, M.; Su, H.; Wang, B.; Wan, C.; Du, W.; Chen, P. A magnet-actuated microfluidic array chip for high-throughput pretreatment and amplification and detection of multiple pathogens. Analyst 2022, 147, 2433–2441. [Google Scholar] [CrossRef]
  229. Giacometti, M.; Milesi, F.; Coppadoro, P.L.; Rizzo, A.; Fagiani, F.; Rinaldi, C. A Lab-On-chip Tool for Rapid, Quantitative, and Stage-selective Diagnosis of Malaria. Adv. Sci. 2021, 8, 2004101. [Google Scholar] [CrossRef] [PubMed]
  230. Zhang, G.H.; Wang, Y.H.; Zhou, W.H.; Lei, Y.M.; Lu, J.S.; Yin, W.J. A Magnetically Driven Tandem Chip Enables Rapid Isolation and Multiplexed Profiling of Extracellular Vesicles. Angew. Chem. Int. Ed. Engl. 2023, 62, e202315113. [Google Scholar] [CrossRef]
  231. Golozar, M.; Chu, W.K.; Casto, L.D.; McCauley, J.; Butterworth, A.L.; Mathies, R.A. Fabrication of high-quality glass microfluidic devices for bioanalytical and space flight applications. MethodsX 2020, 7, 101043. [Google Scholar] [CrossRef]
  232. Zeng, L.; Hu, S.; Chen, X.; Zhang, P.; Gu, G.; Wang, Y. Extraction of small extracellular vesicles by label-free and biocompatible on-chip magnetic separation. Lab Chip 2022, 22, 2476–2488. [Google Scholar] [CrossRef]
  233. İçöz, K.; Akar, Ü.; Ünal, E. Microfluidic Chip based direct triple antibody immunoassay for monitoring patient comparative response to leukemia treatment. Biomed. Microdevices 2020, 22, 48. [Google Scholar] [CrossRef]
  234. Chen, H.; Chen, C.; Zhou, Y.; Yang, C.; Song, G.; Hou, F. Evaluation and Optimization of a Cross-Rib Micro-Channel Heat Sink. Micromachines 2022, 13, 132. [Google Scholar] [CrossRef]
  235. Waqas, H.; Khan, S.A.; Farooq, U.; Muhammad, T.; Alshehri, A.; Yasmin, S. Thermal transport analysis of six circular microchannel heat sink using nanofluid. Sci. Rep. 2022, 12, 8035. [Google Scholar] [CrossRef] [PubMed]
  236. Zhou, W.; Su, H.; Liu, H.; Zhang, Y.; Qiu, N.; Gao, B. Leakage and rotordynamic performance of a semi-Y labyrinth seal structure for centrifugal pump based on multi-frequency whirl method. J. Eng. Gas Turbines Power 2025, 147, 101022. [Google Scholar] [CrossRef]
  237. Li, W.X.; Li, Z.; Han, W.; Li, R.; Zhang, Y. Mechanism of bubble generation in ferrofluidic micro-pumps and key parameters influencing performance. Powder Technol. 2025, 467, 121562. [Google Scholar] [CrossRef]
  238. Lin, Y.H.; Piñan Basualdo, F.N.; Kalpathy Venkiteswaran, V.; Misra, S. Untethered soft magnetic pump for microfluidics-based Marangoni surfer. Sci. Rep. 2024, 14, 20280. [Google Scholar] [CrossRef]
  239. Han, J.; Yeom, J.; Mensing, G.; Flachsbart, B.; Shannon, M.A. Characteristics of electrostatic gas micro-pump with integrated polyimide passive valves. J. Micromech. Microeng. 2012, 22, 095007. [Google Scholar] [CrossRef]
Figure 1. Some applications of microfluidics: (a) fully integrated nucleic acid pretreatment, amplification, and detection on a paper-based platform for identifying EGFR mutations in lung cancer cells [37]; (b) assembled detection chip [37]; (c) microfluidic chip fabricated using multi-layer soft lithography [38]; (d) preparation and calibration process of ion-selective electrodes on a microchip [39]; (e) electromagnetic valve manifold employed in photolithography-based microfluidic chips [38]; (f) a representative application case of digital microfluidics. Open Drop [40]. (a,b) Copyright 2018, Elsevier. All rights reserved; (c,e) Copyright 2021, Yang and Ho; (d) Copyright 2018, Elsevier. All rights reserved; (f) Copyright 2017, MDPI (Basel, Switzerland).
Figure 1. Some applications of microfluidics: (a) fully integrated nucleic acid pretreatment, amplification, and detection on a paper-based platform for identifying EGFR mutations in lung cancer cells [37]; (b) assembled detection chip [37]; (c) microfluidic chip fabricated using multi-layer soft lithography [38]; (d) preparation and calibration process of ion-selective electrodes on a microchip [39]; (e) electromagnetic valve manifold employed in photolithography-based microfluidic chips [38]; (f) a representative application case of digital microfluidics. Open Drop [40]. (a,b) Copyright 2018, Elsevier. All rights reserved; (c,e) Copyright 2021, Yang and Ho; (d) Copyright 2018, Elsevier. All rights reserved; (f) Copyright 2017, MDPI (Basel, Switzerland).
Actuators 14 00460 g001
Figure 2. Implantable microfluidic systems: (a) A pump-free blood filtration device driven by the pressure difference between arteries and veins, fabricated via laser cutting and bonding. This approach enables the microfluidic geometry to adapt to each neural probe’s shape, facilitating integration with various existing neural probes while avoiding the complexity of monolithic microfluidic integration [48]. (b) A 3D-printed neural probe incorporating microfluidic channels [49]. (c) Schematic of a wirelessly controlled implantable insulin delivery system (WIIDS) [50]. (d) A novel integrated transdermal drug delivery system made of polymer, featuring micropumps and microneedles [51]. (e,f) Conceptual diagram and physical prototype of a wearable, rapidly fabricated microneedle patch with enhanced stability for closed-loop diabetes management [52]. (a) Copyright 2021, MDPI (Basel, Switzerland) unless otherwise stated. (b) Copyright 2023, Authors. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CCBY). (c) Copyright 2025, Springer Nature Limited. (d) Copyright 2022, MDPI (Basel, Switzerland) unless otherwise stated. (e,f) Copyright 2025, Springer Nature Limited.
Figure 2. Implantable microfluidic systems: (a) A pump-free blood filtration device driven by the pressure difference between arteries and veins, fabricated via laser cutting and bonding. This approach enables the microfluidic geometry to adapt to each neural probe’s shape, facilitating integration with various existing neural probes while avoiding the complexity of monolithic microfluidic integration [48]. (b) A 3D-printed neural probe incorporating microfluidic channels [49]. (c) Schematic of a wirelessly controlled implantable insulin delivery system (WIIDS) [50]. (d) A novel integrated transdermal drug delivery system made of polymer, featuring micropumps and microneedles [51]. (e,f) Conceptual diagram and physical prototype of a wearable, rapidly fabricated microneedle patch with enhanced stability for closed-loop diabetes management [52]. (a) Copyright 2021, MDPI (Basel, Switzerland) unless otherwise stated. (b) Copyright 2023, Authors. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CCBY). (c) Copyright 2025, Springer Nature Limited. (d) Copyright 2022, MDPI (Basel, Switzerland) unless otherwise stated. (e,f) Copyright 2025, Springer Nature Limited.
Actuators 14 00460 g002
Figure 3. A disposable electromagnetic bidirectional micropump utilizing rotating multipole annular magnetic coupling, developed by Meng et al. [64]. (a) The microfluidic device; (b) dip the microfluidic device in a finger; (c) apply pressure down the micropump to create a negative pressure environment; (d) device in 3D-printed phone clamp after adsorbing; (e) 3D-printed mold for the extension part and PDMS extension with dipping glass capillary; (f) whole device after assembling; (g) filter paper with the blood completely absorbed. Copyright 2021, MDPI (Basel, Switzerland) unless otherwise stated.
Figure 3. A disposable electromagnetic bidirectional micropump utilizing rotating multipole annular magnetic coupling, developed by Meng et al. [64]. (a) The microfluidic device; (b) dip the microfluidic device in a finger; (c) apply pressure down the micropump to create a negative pressure environment; (d) device in 3D-printed phone clamp after adsorbing; (e) 3D-printed mold for the extension part and PDMS extension with dipping glass capillary; (f) whole device after assembling; (g) filter paper with the blood completely absorbed. Copyright 2021, MDPI (Basel, Switzerland) unless otherwise stated.
Actuators 14 00460 g003
Figure 4. Schematic diagrams of conventional mechanical micropumps [33]. Copyright 2019, Elsevier Ltd.
Figure 4. Schematic diagrams of conventional mechanical micropumps [33]. Copyright 2019, Elsevier Ltd.
Actuators 14 00460 g004
Figure 5. Various types of actuators: (a) piezoelectric actuator; (b) electromagnetic actuator; (c) thermopneumatic actuator; (d) phase change actuation; (e) shape memory alloy (SMA) actuator; (f) electrostatic actuator; (g) electroactive polymer (EAP) actuator [33]. (ag) Copyright 2019, Elsevier Ltd.
Figure 5. Various types of actuators: (a) piezoelectric actuator; (b) electromagnetic actuator; (c) thermopneumatic actuator; (d) phase change actuation; (e) shape memory alloy (SMA) actuator; (f) electrostatic actuator; (g) electroactive polymer (EAP) actuator [33]. (ag) Copyright 2019, Elsevier Ltd.
Actuators 14 00460 g005
Figure 6. Diagrams of some pneumatic micropump. (a) A single-layer valve is installed at a bifurcation of the sorter, with a flexible elastic diaphragm separating the two microchannels. When the valve is activated, droplets are directed to the lower branch. (b) Four parallel walls divide the main channel into five sampling channels, with two microactuators positioned on both sides. When the microactuators are activated, the walls bend, leaving only one sampling channel fully open. Droplets then flow into the sampling channel. The sampling channel can be selected by adjusting the pressure of the two microactuators. (c) The single-layer membrane valve consists of an air channel on one side of the bifurcation structure. When the valve is activated, droplets flow into the lower channel. (d) The PDMS wall deforms under pneumatic pressure. When the bottom valve is activated, droplets flow into the upper branch.
Figure 6. Diagrams of some pneumatic micropump. (a) A single-layer valve is installed at a bifurcation of the sorter, with a flexible elastic diaphragm separating the two microchannels. When the valve is activated, droplets are directed to the lower branch. (b) Four parallel walls divide the main channel into five sampling channels, with two microactuators positioned on both sides. When the microactuators are activated, the walls bend, leaving only one sampling channel fully open. Droplets then flow into the sampling channel. The sampling channel can be selected by adjusting the pressure of the two microactuators. (c) The single-layer membrane valve consists of an air channel on one side of the bifurcation structure. When the valve is activated, droplets flow into the lower channel. (d) The PDMS wall deforms under pneumatic pressure. When the bottom valve is activated, droplets flow into the upper branch.
Actuators 14 00460 g006
Figure 7. Some common processing methods for micropumps [33]. Repainted with permission, copyright 2019, Elsevier Ltd.
Figure 7. Some common processing methods for micropumps [33]. Repainted with permission, copyright 2019, Elsevier Ltd.
Actuators 14 00460 g007
Figure 8. Multi-layered micropump architectures. (a) Three-layer pneumatic micropump based on micro-blower technology [150]. (b) Multi-layer piezoelectric micropump configuration for experimental investigation of integrated heat sink thermal management performance [154]. (c) Multi-layer pneumatic micropump developed by Lin et al. [97]. (d) Multi-layer single-chamber and quad-chamber piezoelectric micropumps [150]. (a) Copyright 2023, Royal Society of Chemistry; (b) Copyright 2021, MDPI (Basel, Switzerland) unless otherwise stated; (c) Copyright 2021, MDPI (Basel, Switzerland) unless otherwise stated; (d) Copyright 2022, MDPI (Basel, Switzerland) unless otherwise stated.
Figure 8. Multi-layered micropump architectures. (a) Three-layer pneumatic micropump based on micro-blower technology [150]. (b) Multi-layer piezoelectric micropump configuration for experimental investigation of integrated heat sink thermal management performance [154]. (c) Multi-layer pneumatic micropump developed by Lin et al. [97]. (d) Multi-layer single-chamber and quad-chamber piezoelectric micropumps [150]. (a) Copyright 2023, Royal Society of Chemistry; (b) Copyright 2021, MDPI (Basel, Switzerland) unless otherwise stated; (c) Copyright 2021, MDPI (Basel, Switzerland) unless otherwise stated; (d) Copyright 2022, MDPI (Basel, Switzerland) unless otherwise stated.
Actuators 14 00460 g008
Figure 9. Applications of micropump chips in cell culture systems. (a) Schematic of a cell culture chip for reconstructing pulmonary arterial cells affected by PAH (Pulmonary Arterial Hypertension) [167]. (b) PAH-on-a-chip platform designed to culture three primary pulmonary arterial cell types (adventitial fibroblasts, smooth muscle cells, and endothelial cells) in three-layered structures mimicking pulmonary arteries/arterioles [168]. (c) High-availability kinetic pump-integrated microfluidic plate (KIM-Plate) for cell culture-based multi-organ microphysiological systems. The KIM-Plate consists of a lid, chamber layer, and channel layer, mounted on a magnetic stirrer motor base to control kinetic pumps and perfuse culture media [169]. (a) Copyright 2022, MDPI (Basel, Switzerland) unless otherwise stated; (b) Copyright 2020, Royal Society of Chemistry 2025; (c) Copyright 2021, MDPI (Basel, Switzerland) unless otherwise stated.
Figure 9. Applications of micropump chips in cell culture systems. (a) Schematic of a cell culture chip for reconstructing pulmonary arterial cells affected by PAH (Pulmonary Arterial Hypertension) [167]. (b) PAH-on-a-chip platform designed to culture three primary pulmonary arterial cell types (adventitial fibroblasts, smooth muscle cells, and endothelial cells) in three-layered structures mimicking pulmonary arteries/arterioles [168]. (c) High-availability kinetic pump-integrated microfluidic plate (KIM-Plate) for cell culture-based multi-organ microphysiological systems. The KIM-Plate consists of a lid, chamber layer, and channel layer, mounted on a magnetic stirrer motor base to control kinetic pumps and perfuse culture media [169]. (a) Copyright 2022, MDPI (Basel, Switzerland) unless otherwise stated; (b) Copyright 2020, Royal Society of Chemistry 2025; (c) Copyright 2021, MDPI (Basel, Switzerland) unless otherwise stated.
Actuators 14 00460 g009
Figure 10. Mechanical micropumps. (a) Prototype of a valveless PZT-driven micropump developed by Ma et al. [184]. (b) Thermal bubble-actuated micropump created by Liu et al. [185], with dimensions comparable to a coin. (c) A single-chamber piezoelectric pump [186] utilizing a circular monocrystalline piezoelectric actuator and cantilever check valves, demonstrating superior output performance and compact size. (d) Disposable hemoglobin sensor chip integrated with a smartphone-based colorimetric analyzer, developed by Meng et al. [64], capable of detecting abnormally shaped red blood cells. (e) Piezoelectric peristaltic micropump employing PVC gel and micropatterned surfaces, developed by Motohashi et al. [187]. (a) Copyright 2015, Elsevier B.V. All rights reserved; (b) Copyright 2021, MDPI (Basel, Switzerland) unless otherwise stated; (c) Copyright 2022, AIP Publishing LLC; (d) Copyright 2022, MDPI (Basel, Switzerland) unless otherwise stated; (e) Copyright 2022, Springer Nature Limited.
Figure 10. Mechanical micropumps. (a) Prototype of a valveless PZT-driven micropump developed by Ma et al. [184]. (b) Thermal bubble-actuated micropump created by Liu et al. [185], with dimensions comparable to a coin. (c) A single-chamber piezoelectric pump [186] utilizing a circular monocrystalline piezoelectric actuator and cantilever check valves, demonstrating superior output performance and compact size. (d) Disposable hemoglobin sensor chip integrated with a smartphone-based colorimetric analyzer, developed by Meng et al. [64], capable of detecting abnormally shaped red blood cells. (e) Piezoelectric peristaltic micropump employing PVC gel and micropatterned surfaces, developed by Motohashi et al. [187]. (a) Copyright 2015, Elsevier B.V. All rights reserved; (b) Copyright 2021, MDPI (Basel, Switzerland) unless otherwise stated; (c) Copyright 2022, AIP Publishing LLC; (d) Copyright 2022, MDPI (Basel, Switzerland) unless otherwise stated; (e) Copyright 2022, Springer Nature Limited.
Actuators 14 00460 g010
Figure 11. Ferrofluid.
Figure 11. Ferrofluid.
Actuators 14 00460 g011
Figure 12. Schematic diagram of a conventional toroidal ferrofluidic micropump, from Hatch et al. [205]. Repainted with permission. Copyright 2015, IEEE—All rights reserved, including rights for text and data mining and training of artificial intelligence and similar technologies.
Figure 12. Schematic diagram of a conventional toroidal ferrofluidic micropump, from Hatch et al. [205]. Repainted with permission. Copyright 2015, IEEE—All rights reserved, including rights for text and data mining and training of artificial intelligence and similar technologies.
Actuators 14 00460 g012
Table 1. Advantages, disadvantages, and applications of various types of mechanical micropumps.
Table 1. Advantages, disadvantages, and applications of various types of mechanical micropumps.
TypeAdvantagesDisadvantagesApplications
Piezoelectric
  • High output pressure
  • Fast response
  • Low power consumption
  • No thermal effect
  • High driving voltage
  • Fragile
  • Sensitive to air bubbles
  • Complex manufacturing
  • Biomedical Engineering
  • Electronics Cooling
Electromagnetic
  • High output pressure
  • Wide flow range
  • No thermal effect
  • Diverse fluid compatibility
  • Interference of magnetic field
  • High power consumption
  • Large size
  • Complex structure
  • Point-of-Care Testing
  • Drug Infusion Pump
Thermopneumatic
  • Simple structure
  • Fast response
  • Capability to manipulate extremely small droplets
  • Significant heating effect
  • Compatible with only certain fluids
  • Low pressure
  • Inkjet Print Head
  • Microthruster for Satellites
  • Spacecraft
Phase change
  • Enormous driving force
  • High output pressure
  • Simple structure
  • Disposable use and packaging
  • Low energy efficiency
  • Thermal effects
  • Unsteady flow rate
  • High dependency on fluid properties
  • MEMS Cooling
  • Micromixer
Shape memory alloy
  • High output pressure
  • Compact design
  • Long stroke
  • Low noise
  • Low work efficiency
  • Thermal effect issues
  • Limited fatigue life
  • Industrial Automation
  • Automotive Industry
Electroactive polymer
  • Low driving voltage
  • Potential biocompatibility
  • Relatively simple structure
  • Low output force
  • Low pressure
  • Slow response speed
  • Biomedical Engineering
  • Microdosing Applications
Table 2. Geometric and performance characteristics of nozzle/diffuser (valveless): (length: diffuser tube length; Width 1: narrow-end width; Width 2: wide-end width; No—not mentioned).
Table 2. Geometric and performance characteristics of nozzle/diffuser (valveless): (length: diffuser tube length; Width 1: narrow-end width; Width 2: wide-end width; No—not mentioned).
Ref.YearTypeLength (μm)Width 1 (μm)Width 2 (μm)Height (μm)Angle (°)Material
[120]2015PE3000No75No8Si
[121]2023PE700300100500NoPDMS
[122]2020PE10936740No7Si
[116]2019PE(PZT-5A)13001307010010No
[123]2018EM(Cu)11001255050010PDMS
[124]2020ES(Cu)2300500100No16PDMS
[125]2018EM1000NoNoNo15.64PDMS
[126]2020PE(PZT)No2000200100NoPMMA
[127]2014FMP748427961054100013.2PMMA
[128]2015FMP496310002004509.5PMMA
[129]2012FMP748427961054100013.2PMMA
[88]2022PMNoNoNoNo9.5PMMA
[130]2024EM15531080600No17.6PDMS
[131]2018PM1000182.83020014No
[132]2023No145227741000150062.8PMMA
Table 3. Geometric and performance parameters of some mechanical micropumps. (PEA: piezoelectric actuator; ESA: electrostatic actuator; PMA: pneumatic actuator; EMA: electromagnetic actuator. Note: The following symbols in the table represent: \—not applicable; No—not mentioned. Additionally, to ensure the clarity and esthetics of the table, the units of physical quantities are specified as follows and will not be repeated within the table: driving voltage: U–V, back pressure: P–kPa, flow rate: Q–mL/min, frequency: f–Hz, and length dimensions: thickness–μm; size–mm).
Table 3. Geometric and performance parameters of some mechanical micropumps. (PEA: piezoelectric actuator; ESA: electrostatic actuator; PMA: pneumatic actuator; EMA: electromagnetic actuator. Note: The following symbols in the table represent: \—not applicable; No—not mentioned. Additionally, to ensure the clarity and esthetics of the table, the units of physical quantities are specified as follows and will not be repeated within the table: driving voltage: U–V, back pressure: P–kPa, flow rate: Q–mL/min, frequency: f–Hz, and length dimensions: thickness–μm; size–mm).
Ref.YearTypeDiaphragmThicknessMaterialSizeValveUPQf
[120]2019PE(PZT)PDMS250NoΦ12Diffuser500.220100
[121]2016PE(PZT)Si<20Si10 × 10\803.10.362.52 k
[101]2015PMPDMS200\4 × 4YesNo304963
[188]2025PE(PZT)NoNoNoΦ38\1000.78200.743
[115]2024PE(PZT)Copper200NoNoUmbrella11042.3767400
[189]2019PEPMMANoNoΦ40, t0.8Umbrella600.82No300
[131]2018EM(Cu)PDMS100PDMSΦ10DiffuserNo0.350.5245
[132]2020ES(Cu)Iron-based MGNoPDMSΦ8Diffuser20No0.171
[113]2018EM(Cu)NdFeB5000TeflonΦ2, H35BallNo0.540.215
[81]2021PENo200stainless steelΦ18Spring300NoNo60
[63]2022EM(NdFeB)PDMS275PMMAΦ6No4.80.50.8650
[128]2023PEPDMSNoPMMA78 × 12Yes1501.7700716
[190]2018EMPDMSNoPDMSNoDiffuser1.41.276 μL9
[157]2025PE(PZT)NoNoresin22 × 22 × 5\42No61822.5 k
[191]2024PE(PZT)PET50–200PMMAΦ25, H3Umbrella30018.741.4No
[192]2024PE(PZT)NoNoPET10 × 10 × 1Cantilever20010099.6790
[193]2021PE(PZT)PDMS200PMMA60 × 60 × 12Ball44815.3131.6750
[194]2019PE(PZT)PI40PMMA100 × 20 × 15Cantilever300150.160
[129]2024PE(PZT)PDMS600PDMSΦ10, H0.5Diffuser9No5.6925
[195]2024PE(PZT)No\resin35 × 35 × 0.8\10019.7420.12120
[196]2020PE(PZT)Si100PMMAΦ25, 4Diffuser45No9.1No
[130]2020PE(PZT)PMMA200PDMSΦ10Diffuser1000.35150600
Table 4. Geometric parameters and performance characteristics of representative ferrofluidic micropumps (No—data not reported in the reference; \—parameter not applicable; +—variant based on prototype design; Φ—diameter; H—pump chamber/magnet height; d—hydraulic diameter of flow channel. The units for some parameters are listed as follows: Pressure—kPa, velocity—μL/min, size—mm, field—mT).
Table 4. Geometric parameters and performance characteristics of representative ferrofluidic micropumps (No—data not reported in the reference; \—parameter not applicable; +—variant based on prototype design; Φ—diameter; H—pump chamber/magnet height; d—hydraulic diameter of flow channel. The units for some parameters are listed as follows: Pressure—kPa, velocity—μL/min, size—mm, field—mT).
Ref.YearConstructionPressureVelocityMaterialLayerSizeFabricationMagnetField
[75]2024Circle1.351756.3PMMA2Φ10, H0.5NoPM, 1, Φ5495
[217]2020HemicycleNo9.02PMMA2Φ50, 0.4 × 0.4CNCPM, 2, Φ15, t5160
[79]2018AnnularNo3.5 × 105Stratasys J7502Φ80, H30, d73D PrintingEM, 10, 200 N, 1–6 ANo
[201]2006AnnularNo3.8Si4NoEtching coatingPM, 295
[218]2013AnnularNoNoPDMS3Φ15, H0.5Injection moldingPM, 1No
[219]2014Annular+No128PMMA2Φ21, d1CO2 laserPM, 2No
[77]2004Circle2NoSi215 × 28 × 0.8MEMESPM, 1, Φ3, t2340
[220]2015Liner+0.99934PMMA6Φ6, H1CO2 laserPM, 1510
[221]2015Circle0.651310PMMA4Φ3.5, H2CO2 laserPM, 1, 4 × 6 × 21200
[208]2018Annular1.1349.32PDMS2Φ25, H3.5Soft lithographyPM450
[222]2012Annular0.6693PMMA2Φ21, d4CO2 laserPMNo
[215]2004Liner12NoGlass23 × 4Photo etchingPM, 2No
[205]2001Annular1.3245.8Si4NoPhoto etchingPM, 1, Φ6, t3350
[223]2006Liner0.21200Glass\Φ44NoEM, 7No
[224]2005Liner2.530PMMA70.1 × 22 × 6Powder blastingPM, 180
Table 5. Comparison of pumping performance between mechanical micropumps and ferrofluidic micropumps.
Table 5. Comparison of pumping performance between mechanical micropumps and ferrofluidic micropumps.
Ref.First AuthorYearTypeP (kpa)Q (mL/min)
[75]Wang, Y.2024Ferrofluidic Micropump1.351756.3
[220]Ashouri, M.2015Ferrofluidic Micropump0.994934
[221]Ashouri, M.2015Ferrofluidic Micropump0.6471310
[202]Liu, B.D.2018Ferrofluidic Micropump1.1349.32
[199]Hatch, A.2001Ferrofluidic Micropump1.3245.8
[120]Gidde, R.R.2019Piezoelectric Micropump (PZT-5A)0.220
[134]Aggarwal, S.2016Piezoelectric Micropump (PZT)3.10.36
[188]Huang, J.2025Piezoelectric Micropump (PZT-5A)0.784200.7
[63]Qi, C.2022Electrostatic Micropump (NdFeB)0.50.86
[208]Liu, B.D.2018Electrostatic Micropump (Cu)0.5390.21
[190]Rusli, M.Q.A.2018Electrostatic Micropump1.20.076
[238]Lin, Y.2015Pneumatic Micropump30496
[97]Lin, J.L.2022Pneumatic Micropump68.9512.48
[138]Ni, J.2012Pneumatic Micropump250.034
[98]Machauf, A.2005Electrostatic Micropump (Cu)\0.001
[239]Han, J.2012Electrostatic Micropump (Cu)\0.11
[99]Lee, K.S.2013Electrostatic Micropump (Cu)\40/250/150
[160]Ochoa, M.2012Thermopneumatic Micropump5.860.23
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhou, X.; Li, Z.; Han, B.; Guo, Q.; Qing, Z. Performance Comparison of Mechanical and Ferrofluidic Micropumps: Structural and Operational Perspectives. Actuators 2025, 14, 460. https://doi.org/10.3390/act14090460

AMA Style

Zhou X, Li Z, Han B, Guo Q, Qing Z. Performance Comparison of Mechanical and Ferrofluidic Micropumps: Structural and Operational Perspectives. Actuators. 2025; 14(9):460. https://doi.org/10.3390/act14090460

Chicago/Turabian Style

Zhou, Xing, Zhenggui Li, Baozhu Han, Qinkui Guo, and Zhichao Qing. 2025. "Performance Comparison of Mechanical and Ferrofluidic Micropumps: Structural and Operational Perspectives" Actuators 14, no. 9: 460. https://doi.org/10.3390/act14090460

APA Style

Zhou, X., Li, Z., Han, B., Guo, Q., & Qing, Z. (2025). Performance Comparison of Mechanical and Ferrofluidic Micropumps: Structural and Operational Perspectives. Actuators, 14(9), 460. https://doi.org/10.3390/act14090460

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