Lingxuan Hao†
,
Mohammed Imran Khan†
,
Yilin Lei†
,
Shuneng Zhou and
Bei Fan
*
Mechanical Engineering Department, Michigan State University, 48823-East Lansing, Michigan, USA. E-mail: fanbei1@msu.edu
First published on 23rd July 2025
Water, with its natural abundance and various forms, has been a promising sustainable energy source. To harness water energy more efficiently, interfacial electro-hydrodynamics (EHD) micro/nano-fluidic energy conversion and harvesting technologies have rapidly advanced over the past few decades. Compared to conventional water energy harvesting methods like hydroelectric dams and tidal power plants, EHD-based approaches exhibit unique advantages in capturing random, low-frequency, and intermittent fluid motions, enabling the effective harvesting of untapped energy sources such as raindrops, tides, and even ambient humidity. This review systematically summarizes the major interfacial EHD inspired micro/nano-fluidic energy harvesting and conversion strategies, providing an in-depth analysis of their working principles, design principles for enhancing electric output, and their potential applications that mainly include power sources and self-powered devices. Furthermore, we highlight the key challenges facing this field and discuss future research directions and breakthroughs required to facilitate the feasibility and scalability of EHD-based energy harvesting and conversion systems. With continued advancements, these technologies offer significant promise for transitioning from laboratory research to practical applications, providing sustainable and distributed energy solutions.
Water has long been a source of both mechanical and electrical energy. Traditional hydropower technologies, such as hydroelectric dams and tidal power plants, have demonstrated the feasibility of large-scale water energy harvesting by converting the kinetic energy of flowing water into electricity. While effective, these systems often require extensive infrastructure, are costly, and are limited to specific geographical locations.6 More importantly, a large amount of water energy remains untapped in the forms of raindrops, tides, and even moisture.2,6–9 These underutilized sources, found in widely distributed and dynamic environments, hold immense potential for energy harvesting. However, their low frequency, sparse distribution, and diverse forms make achieving efficient energy harvesting through conventional methods challenging.8,10–12
In addition to large-scale renewable energy solutions, the increasing demand for portable and self-powered electronic devices has driven the development of miniaturized, efficient, and low-cost energy harvesting systems.13–17 With the rapid proliferation of wearable sensors,16,18 biomedical devices,19,20 and Internet-of-Things (IOT) applications,21–23 traditional power generation methods are no longer sufficient to meet energy needs. Liquid-based energy harvesting, with its adaptability and vast potential, presents an opportunity to create compact, noise-free, and environmentally friendly power solutions.6
To harness these untapped renewable liquid energy sources, the past few decades have witnessed the rapid emergence and advancement of interfacial electro-hydrodynamics (EHD) inspired micro/nano-fluidic energy harvesting and conversion technologies.24,25 EHD, which studies the interactions between fluid motion and electric fields, presents a versatile and innovative approach to extracting energy from various liquid-based sources. Unlike conventional methods that primarily rely on large-scale hydropower infrastructure, EHD-based technologies leverage fluids as the medium for energy transfer and conversion. This enables them to efficiently harvest energy from a wide range of liquid sources, including not only low-frequency, random, and intermittent mechanical motions but also other liquid-based energy forms. These include but are not limited to thermally induced flows, evaporation-driven motion, ionic gradients, and electrochemical potential differences—all of which contribute to the vast and largely untapped reservoir of liquid energy.
The most promising micro/nano-fluidic energy harvesting technologies based on interfacial EHD include reverse electrowetting-on-dielectric (REWOD) energy harvesters,26,27 triboelectric nanogenerators (TENGs),28,29 and electrokinetic energy conversion (EKEC) devices,30,31 as illustrated in Fig. 1. These technologies leverage interfacial charge interactions and fluid dynamics to enable efficient energy conversion, offering high adaptability and broad applicability for various liquid energy harvesting scenarios.
![]() | ||
Fig. 1 Summary of the working principles and applications of REWOD, TENG, and EKEC. Reproduced with permission.26 Copyright 2020, IEEE. Reproduced with permission.27 Copyright 2016, Wiley. Reproduced with permission.28 Copyright 2017, Elsevier. Reproduced with permission.29 Copyright 2019, Springer Nature. Reproduced with permission.31 Copyright 2015, Wiley. Reproduced with permission.30 Copyright 2023, Elsevier. |
REWOD has emerged as a promising technique for fluidic energy harvesting. It leverages the fundamental principles of electrowetting on dielectric (EWOD)32,33 but operates in reverse: the mechanical deformation of droplets redistributes electrical charge, thereby generating electrical energy.34 Recent research has focused on increasing the contact area, increasing the contact area difference (specifically for water-bridge REWOD), optimizing the dielectric film, and increasing the droplet oscillation frequency to improve output performance.34–36
TENGs generate electricity by combining contact electrification and electrostatic induction, efficiently capturing mechanical energy from human motion, ocean waves, rolling droplets, liquid sloshing, and vibrations. When two materials with different electron affinities come into contact and then separate, electrons transfer between them, creating an electrostatic potential difference. Connecting an external circuit allows electron flow, generating electricity. Recent studies have focused on increasing surface charge, facilitating interface charge transfer, and enhancing liquid motion.37–39
EKEC represents a promising mechanism for harvesting energy from fluid motion. This technique takes advantage of the electrokinetic effect, which occurs when an electrolyte interacts with a charged surface to form an electric double layer (EDL) enriched with counterions near the interface. The movement of these counterions along with fluid motion creates an electrical current/potential that can be harnessed for energy. EKEC devices utilize the EDL at solid–liquid or liquid–liquid interfaces to separate charges and generate current under liquid flow. Operating at the nanoscale, EKEC systems can achieve highly efficient energy conversion even in confined spaces. Research in this field has focused on improving electrokinetic efficiency through surface engineering and exploring the interaction of ions at liquid interfaces.40–43
In addition, these fluidic energy harvesting technologies have promising applications across various fields. They can efficiently convert fluid motion or phase transitions into electrical energy, making them suitable for power sources, wearable electronics, and autonomous sensing systems. Beyond energy harvesting, they also function as self-powered sensors, enabling real-time detection of physical, chemical, and biological signals. Specifically, REWOD is particularly effective in harvesting low-frequency mechanical energy, especially from intermittent and aperiodic motion, making it highly suitable for energy harvesting from human motion. This capability allows it to power wearable44–46 and implantable devices,47,48 as well as mobile electronic systems34,49 such as GPS devices, smartphones, and medical prosthetics. TENGs are effective in both localized droplet interactions and large-scale liquid flows, capturing energy from raindrops,50 ocean waves,51,52 and pipeline flows,53,54 making them useful in diverse energy harvesting scenarios. EKEC operates efficiently across various environments, from microfluidic systems55–58 to large-scale natural water bodies such as oceans and rivers.59 It can utilize droplets, moisture gradients, evaporation, and continuous flows to generate power, serving as a stable and continuous energy source.60–62
In recent years, significant research efforts have been dedicated to developing EHD-based micro/nano-fluidic energy harvesting technologies. As these technologies continue to evolve, it is essential to provide a comprehensive review of the literature, focusing on strategies to enhance electrical energy output, assess practical applications, and address key challenges and prospects. This review summarizes recent advancements in EHD-based micro/nano-fluidic energy harvesting. It begins by outlining the fundamental transduction mechanisms and then presents key developments in structural designs and material innovations that enhance fluid–device interactions, with the aim of improving overall energy conversion efficiency. These emerging technologies show great potential for powering microelectronic devices,29,35,63 lab-on-a-chip systems,64–66 next-generation self-powered sensors,67,68 and smart wearable devices.14,16,69 Finally, future research directions are discussed, highlighting potential breakthroughs that could enhance the feasibility, scalability, and real-world applicability of EHD-based micro/nano-fluidic energy harvesting systems.
![]() | (1) |
![]() | (2) |
![]() | (3) |
![]() | (4) |
However, when the applied voltage exceeds a few hundred millivolts, it can lead to electrolyte decomposition.84 To address this issue, Berge86 introduced a thin insulating layer to prevent direct contact between the liquid and metallic electrode. With this approach, the applied voltage can be increased to several hundred volts, making this method more reliable. This technique is known as EWOD. As shown in Fig. 3a, the system can be modeled as two capacitors in series:87,88 the EDL capacitor at the insulator–droplet interface (cH) and the electrical capacitor across the dielectric film (cd). These capacitances per unit area are given by eqn (5) and (6), respectively.
![]() | (5) |
![]() | (6) |
Since the thickness of the dielectric film d generally ranges from hundreds of nanometers to several microns,48,89–91 the Debye length λD typically ranges from sub-nanometer to tens of nanometers, depending on the metal surface and the electrolyte concentration.92,93 The dielectric constants of common coating layers typically range from a few to over one hundred.36,48,94 The dielectric constant of liquid electrolytes is typically on the order of tens.94 As a result, d/εd is usually much larger than λD/εH. The total capacitance per unit area of the system, c, can be calculated as c−1 = cd−1 + cH−1. As a result, c can be approximated as cd. Consequently, Berge32,86 derived the EWOD equation (eqn (8)) by combining the Lippmann equations (eqn (2) and (3)) with Young's equation (eqn (7)),33 which later became known as the Young–Lippmann equation.85,95 Mugele et al.33 defined the last term of eqn (8) as the dimensionless electrowetting number, which represents the ratio of the electrostatic energy stored in the electrical capacitor across the dielectric film to the liquid–vapor interfacial energy.24 The electrostatic energy is given by eqn (9). As shown in eqn (8) and (9), the contact angle of the liquid droplet on the solid surface decreases with increasing voltage V, indicating that the droplet tends to spread on the dielectric surface (Fig. 3a(ii)). This behavior demonstrates that an external electric field enhances the wettability of the dielectric surface by introducing extra electrostatic energy at the interface.
![]() | (7) |
![]() | (8) |
![]() | (9) |
In 2013, Krupenkin and Taylor34 first proposed that reversing the EWOD process could convert the mechanical energy of liquid motion into electrical energy, a phenomenon known as REWOD. The schematic of REWOD is shown in Fig. 3b.24 Studies have shown that using ions as charge carriers is generally less efficient for energy generation, primarily due to strong charge trapping at the solid–liquid interface.96 In contrast, the electrons are less susceptible to interfacial trapping, resulting in a higher density of available free carriers and significantly enhanced charge transport efficiency. As a result, liquid metals are considered ideal fluid in the REWOD systems. In addition, liquid metals exhibit high electrical conductivity, high surface tension, and low vapor pressure. High conductivity enables efficient electron transport and reduces resistive losses; high surface tension contributes to greater energy storage per unit interfacial area; and low vapor pressure enhances the stability of long-term operation. As a result, in this system, a liquid metal droplet is sandwiched between two conductive electrodes, one of which is coated with a dielectric film. Electrons, supplied by an external direct current (DC) bias voltage source, act as charge carriers. When an external force is periodically applied to the electrode, the droplet moves or deforms, causing a periodic change in the solid–liquid interfacial area. This variation alters the capacitance of the capacitor across the dielectric film,49 thereby converting the periodic mechanical energy into electrical energy.34 It is noted that the bias voltage itself does not contribute to net energy production during the energy generation process; a higher bias voltage can enhance output power by increasing the number of electrons. However, increasing the bias voltage also raises the risk of dielectric film breakdown. More importantly, this contradicts the fundamental goal of energy harvesting, as it requires additional external energy input.35,48 In this review paper, such REWOD systems that require a bias voltage are referred to as the classical REWOD method.
To address these drawbacks, Moon et al.35 proposed a new method for alternating current (AC) electrical power generation using plain water, without the need for an external bias voltage source. Fig. 3c and d illustrate the working principle and the corresponding equivalent electrical circuits, respectively. In this system, a plain water droplet is confined between two electrodes to form a water bridge. The top electrode is coated with a hydrophobic dielectric film (PTFE), while the bottom electrode is hydrophilic (see section 2.3.2 for details). The contact area between the water and the bottom electrode remains nearly constant during each oscillation period because of its small contact angle and pinning effects. In contrast, the contact area between the top electrode and water droplet varies significantly during each oscillation cycle. Accordingly, the capacitances at the top and bottom interfaces are described by eqn (10) and (11), respectively.
The ions in the plain water act as charge carriers, and the system can be modeled as two capacitors and two resistors connected in series, as shown in Fig. 3d. In Fig. 3d(i), this corresponds to a steady-state condition in which the water bridge is fully formed and the distance between the plates remains constant over time. Here, no current flows and the voltage–capacitance relationships are given by: VT = QT/CT and VB = QB/CB. However, as the two plates move toward each other, the solid–fluid contact area at the top plate increases, whereas that at the bottom plate remains nearly unchanged. As a result, CT increases while CB remains constant. In this non-equilibrium state (as shown in Fig. 3d(ii)), electrical charging occurs at the top plate, while discharging occurs at the bottom plate. Therefore, the voltage expressions are updated as: VB(t) = (QB − q(t))/CB, VT(t) = (QT + q(t))/CT(t). If the distance between the two electrodes oscillates at frequency f, the two electrodes continuously charge and discharge in different phases, generating an AC equal to dq(t)/dt. In this case, the voltage drop ΔV between the two capacitors is given by eqn (12).
![]() | (10) |
![]() | (11) |
![]() | (12) |
Based on eqn (12), achieving a higher ΔV requires a greater difference between CT and CB. To accomplish this, a dielectric film is attached to the top electrode to increase CT (as analyzed above). To further increase the difference between CT and CB, the disparity between AT and AB should also be maximized (to be discussed in section 2.3.2).97 This method (hereafter referred to as water-bridge REWOD) eliminates the need for a bias voltage while ensuring that the output power is not too low.
According to eqn (9), the output power of REWOD during a single oscillation is directly proportional to the capacitance per unit area, the solid–liquid contact area, and the square of the bias voltage. The capacitance per unit area can be increased by enhancing the dielectric constant of the dielectric film and reducing its thickness. Specifically, in the water-bridge REWOD system, the output power also increases with the difference between the top and bottom contact areas. Moreover, the total output energy is also strongly influenced by the oscillation frequency of the droplet.
It is important to note that, as previously discussed, although increasing the bias voltage can enhance the output power and the bias voltage itself does not contribute to the net output power, this approach contradicts the fundamental goal of energy harvesting. Moreover, a higher voltage between the electrodes increases the risk of dielectric film breakdown. Therefore, relying on a higher applied voltage is not a practical strategy for improving output power. On the contrary, achieving a lower bias voltage or even a bias-free operation has become a growing trend for enhancing the practical applicability of REWOD systems.
Based on the above, the energy generated by the entire REWOD system can be increased through four main methods: (i) increasing the contact area, (ii) increasing the contact area difference, (iii) optimizing the dielectric film, and (iv) increasing the droplet oscillation frequency.
Although the maximum droplet volume is limited by the capillary length, which restricts the increase in output power, this limitation can be addressed by increasing the number of droplets operating in parallel (as shown in Fig. 4a), which directly increases the total contact area. In water-bridge REWOD systems, as explained in section 2.2, the top and bottom capacitors are continuously charged and discharged out of phase, generating an AC power. Therefore, when increasing the number of droplets, it is essential to maintain multiple uniformly sized water bridges in parallel. Experimental data indicates that the output power increases quadratically with the number of droplets. However, this quadratic relationship begins to break down when the number exceeds 17, likely due to unsynchronized top contact areas of the water bridges caused by inevitable variations in droplet size and shape.35 Approximately 1 mL of water—arranged as a droplet array (the exact number of droplets was not specified in the original paper)—can generate about 1.5 μW of power, with a maximum power density of 0.3 μW cm−2.35 This method demonstrates great potential for achieving high output in REWOD systems and has been widely adopted in recent research.35 As a result, the development of REWOD devices based on droplet arrays has become an important direction for further optimization.
![]() | ||
Fig. 4 Strategies for improving electric output power. a) (i) Power versus the number of water bridges; the solid curve represents a parabolic fit, and data are presented as mean ± standard deviation. (ii) Experimental setup for REWOD with a droplet array. Reproduced with permission.35 Copyright 2013, Springer Nature. b) Working mechanism of high surface area REWOD energy harvesting under pulsating pressure, leading to AC voltage generation. Reproduced with permission.98 Copyright 2022, Elsevier. c) Video freeze shots showing three stages of the electrolyte between electrodes during modulation, with an electrode distance of 4 mm, 2.5 mm, and 1.5 mm. Reproduced with permission.99 Copyright 2021, Springer Nature. d) Representative cross-sectional view of a rough electrode model. Reproduced with permission.48 Copyright 2022, Wiley. e) Field emission scanning electron microscopy image of a fabricated atomic layer deposition film sample. Reproduced with permission.100 Copyright 2017, Elsevier. f) Schematic diagram of laminated structures. Reproduced with permission.36 Copyright 2021, Springer. g) (i) Conceptual design of the bubbler: (1) REWOD chip, (2) membrane, (3) top plate, (4) an array of bubbles, (5) an array of electrodes; (ii) schematics of a simplified single-electrode device used in the experiment: (1) REWOD chip, (2) membrane, (3) top plate, (4) bubble, (5) metal electrode, (6) dielectric coating, and (7) conductive liquid; (iii) equivalent electrical circuit of the single-electrode device. Reproduced with permission.101 Copyright 2015, Springer Nature. |
Previous research on REWOD solely relied on planar electrodes which have a fixed interfacial area, thereby limiting capacitance and consequently limiting power density output.36,49,99,100 Adhikari et al.98 perforated silicon (Si) wafers with numerous micro-sized pores to increase the electrode surface area. Fig. 4b shows that the porous electrode is coated with dielectric materials (SiO2 and CYTOP). A conductive electrolyte is inserted and retracted through the pores under a time varying pulsating pressure, which periodically modulates the electrical capacitance, generating an AC voltage.103 Without applying any external bias voltage, and for a 38 μm pore-size electrode at a 5 Hz modulation frequency, the maximum current and voltage densities per unit planar area were 3.77 μA cm−2 and 1.05 V cm−2, respectively. Additionally, the root mean square power density was measured to be 4.8 μW cm−2.
Moon et al.35 deliberately modified the wetting properties of the electrodes: the top electrode was made hydrophobic by coating polytetrafluoroethylene (PTFE) onto the indium–tin-oxide (ITO) surface, while the bottom electrode (ITO) remained hydrophilic. The reported contact angles for water droplets on these surfaces are 62.5° (top electrode) and 107° (bottom electrode), respectively. Due to the strong pinning effects on the hydrophilic surface, the contact area between the water and the bottom electrode remained nearly constant throughout the oscillation cycle.97 In contrast, the contact area on the hydrophobic top electrode varied significantly during each oscillation cycle, effectively increasing the difference between the top and bottom contact areas.
Adhikari et al.99 also used a similar method. In their design, the top electrode was coated with a metal layer that served as a current collector. The bottom electrode was also coated with a metal layer for conduction, followed by a dielectric layer (e.g., Al2O3 or SiO2) and a fluoropolymer coating (e.g., Teflon or CYTOP) to enhance surface hydrophobicity. As a result, the bottom electrode forms a much larger contact angle with the liquid compared to the top hydrophilic surface. Fig. 4c shows the video freeze shots of the water bridge at different electrode distances. At the maximum displacement (i.e., when the electrode–electrolyte interfacial area is minimized) of 4 mm, the interfacial area between the top electrode and the electrolyte is slightly larger than that of the bottom electrode. Without using any external bias voltage, AC current generation was demonstrated, achieving a power density of 53.3 nW cm−2 at an external excitation frequency of 3 Hz with an optimal external load.
Adhikari et al.48 also enhanced the contact area difference between top and bottom interfaces by increasing the roughness of the bottom electrode through reactive ion etching. As shown in Fig. 4d, this modification enabled a maximum power density of 3.18 μW cm−2 to be harvested from a 50 μL deionized water droplet oscillating at a 5 Hz, without any applied bias voltage, which is 4 times higher than that achieved with planar electrodes.
Krupenkin and Taylor34 investigated the breakdown characteristics of various dielectric materials. Among the tested materials, oxide films performed poorly, while inorganic fluorides exhibited better performance. The best results were observed with fluoropolymers, such as Cytop107 and Teflon-like films. However, despite its minimal charge trapping, Krupenkin and Taylor34 observed in their experiments that Cytop exhibited unreliable breakdown resistance, leading to a higher risk of breakdown, making it less suitable for high-performance applications. In contrast, Ta2O5, with its relatively high dielectric constant of 25, demonstrated superior breakdown resistance. Ultimately, to optimize both charge stability and breakdown performance, nanometer-thick multilayer dielectric films were used, which consist of Cytop deposited on Ta2O5.34
Under the same thickness, dielectric films with higher density have higher breakdown strength and lower leakage current caused by charge trapping, thus enhancing the effective output energy. It is also challenging to use non-vacuum-based film fabrication methods, such as spray deposition and dip-coating, to obtain dense films with thicknesses in the order of nanometers. Vacuum-based methods, particularly atomic layer deposition (ALD), are ideal due to their self-limiting growth process.108 ALD thin films have higher dielectric constants, breakdown strengths, and lower leakage currents, making them suitable for REWOD energy harvesting. Yang et al.100 developed a REWOD energy harvester with an Al2O3 ALD film (see in Fig. 4e). A 45 nm-thick ALD film was demonstrated to achieve lower leakage current density (<10−8 A cm−2), higher capacitance (245.19 nF cm−2), and higher power density (5.59 mW cm−2) at a bias voltage of 13.5 V.
In subsequent research, Yang et al.36 developed a laminated structure (see in Fig. 4f) that is able to further reduce the leakage current in high-dielectric constant materials. This laminated structure consists of two materials: TiO2, a high-dielectric-constant material, and Al2O3, known for its high electrical resistance and band gap energy.109 A REWOD energy harvester incorporating this lamination layer achieved a high power density of 15.36 mW cm−2 at a bias voltage of 30 V, exhibiting lower leakage current and relatively higher capacitance compared to a single layer of TiO2 or Al2O3.
To systematically compare the performance differences among various dielectric films, Table 1 summarizes key parameters including the material type, fabrication method, thickness, dielectric properties, breakdown voltage, leakage current, and corresponding power density. It is evident that both material selection and processing techniques significantly influence the performance of dielectric films. Future optimization efforts should continue to focus on the synergy between material innovation and structural design.
Material | Fabrication method | Thickness (nm) | Capacitance (nF cm−2) | Dielectric constant | Breakdown voltage (V) | Leakage current test field strength (mV cm−1) | Leakage current density (A cm−2) | Power density test bias voltage (V) | Power density (mW cm−2) | Ref. |
---|---|---|---|---|---|---|---|---|---|---|
Ta2O5/Cytop | Sputtering + anodic oxidation | <100 | 16 | 2/23 | ∼10 | N/A | N/A | N/A | N/A | 34 |
Ta2O5/parylene C/Cytop | Chemical vapor deposition | 200/200/70 = 470 | Not measured | 2/3/23 | >50 | N/A | N/A | N/A | N/A | 107 |
Al2O3 | Thermal oxidation | 101 | 53 | 6 | 40–50 | 4.94 | <10−7 | N/A | N/A | 110 |
Al2O3 | Sputter | 85 | 89.27 | 8.57 | 18 | 0–1.0 | <10−2 | 12 | 1.5 | 100 |
Al2O3 | ALD | 140 | 69.61 | 11.01 | >20 | 0–1.0 (film: 100 nm) | <10−8 | 13.5 | 3.38 | 100 |
Al2O3 | ALD | 45 | 245.19 | 12.46 | 13.5 | 0–1.0 (film: 100 nm) | <10−8 | 13.5 | 5.59 | 100 |
TiO2/Al2O3 | ALD | 85/65 = 150 | 146 | 24.73 | 27 | 0–0.8 | <10−8 | 27 | 14.87 | 36 |
Al2O3/TiO2/Al2O3 | ALD | 65/100/65 = 230 | 102 | 26.5 | 30 | 0–0.8 | <10−7 | 30 | 15.36 | 36 |
Hsu et al.49 further investigated the bubble oscillation mechanisms in REWOD systems, proposing an analytical and computational framework to model the collapse and rebound dynamics of the gas–liquid interface. In their formulation, the collapse time tcollapse is derived from a balance between the applied pressure difference and the inertia of the surrounding conductive liquid, given by eqn (13). For the rebound phase, a spherical cap geometry is assumed, and the growth time tgrowth is expressed as eqn (14). The total oscillation period T is approximated as the sum of the collapse and growth times, i.e., T = tcollapse + tgrowth, and the frequency is estimated as f = 1/T. Their results indicate that the oscillation frequency increases with pressure and decreases with bubble radius, while the effects of fluid viscosity and gap thickness are comparatively minor. The model predictions were validated by both CFD simulations and experimental measurements, demonstrating that self-oscillation frequencies exceeding 2 kHz can be achieved under moderate pressures (e.g., ΔP ≈ 0.08 bar using mercury). Furthermore, the internal high-frequency response enables energy harvesting even from low-frequency mechanical excitations (<1 Hz). Reported instantaneous power densities above 104 W m−2 under high-frequency and biased conditions further underscore the potential of bubble-based REWOD systems for efficient energy conversion.
![]() | (13) |
![]() | (14) |
![]() | ||
Fig. 5 Application of REWOD as power sources. a) Schematic of a footwear embedded REWOD microfluidic energy harvester. Reproduced with permission.34 Copyright 2011, Springer Nature. b) Cross-section of a footwear-integrated energy harvester based on the bubbler approach. (1) Top flexible chamber filled with compressed inert gas, (2) REWOD chip, (3) bottom gas chamber. Reproduced with permission.101 Copyright 2015, Springer Nature. c) A REWOD-based vibration harvester. Reproduced with permission.34 Copyright 2011, Springer Nature. d) Experimental results: voltage drop and instantaneous power with time for the case of a 14 droplet REWOD system. Reproduced with permission.35 Copyright 2013, Springer Nature. e) (i) Schematics of an oil pump and the possible location where to place the bubbler; (ii) schematic of a bubbler. Reproduced with permission.101 Copyright 2015, Springer Nature. f) Schematic of flexible electrodes. Reproduced with permission.122 Copyright 2024, IEEE. g) A REWOD unit for a wearable sensor system. Reproduced with permission.26 Copyright 2020, IEEE. |
Hsu et al.101 also illustrated a simple bubbler device (a type of REWOD component mentioned in section 2.2.3) integrated into footwear to enable energy harvesting from human locomotion (about 1 Hz frequency), as shown in Fig. 5b. Bubbles self-oscillate naturally whenever external pressure is applied to the device, allowing effective coupling with mechanical energy sources across a wide range of forces, displacements, and frequencies. The bubbler chip is located between two chambers filled with the pressurized gas. During the heel strike, the top elastic chamber is compressed, forcing the gas through the REWOD chip and inducing thousands of bubble oscillations, each converting a portion of the heel strike's mechanical energy into electrical energy. During the toe-off process, the compressed gas flows from the bottom chamber back to the top chamber through an auxiliary bypass check valve, completing the cycle. This device was estimated to generate about 1 W of usable electrical power. These examples demonstrate the feasibility and potential of footwear-based REWOD energy harvesters for powering wearable or portable electronic devices.
The REWOD process can also harvest mechanical energy from vibrations, a common source found in floors, stairs, vehicles, and equipment housings. This process enables novel harvester architectures that greatly increase power output. For example, Krupenkin and Taylor34 demonstrated a REWOD-based vibration harvester that uses an array of conductive droplets squeezed between two dielectric-coated electrodes, separated by a millimeter-thick elastic spacer, which also acts as a mounting ‘pad’ for the load device (as shown in Fig. 5c). In this system, mechanical vibrations cause periodic changes in the solid–liquid contact area, generating electrical current. With a film stack capacitance of 102 nF cm−2, the power density can be scaled up to 10−1 W cm−2 at a vibration frequency of 50 Hz, allowing for the development of producing power outputs of several watts.
Additionally, the water-bridge REWOD is well-suited for utilizing vibrations and holds significant potential for future microfluidic power generation systems. This system uses water instead of liquid metal, with the ions in the water serving as charge carriers. It can provide AC current power without requiring an external bias voltage source (shown in Fig. 5d). Moreover, the use of water eliminates the toxicity risks associated with liquid metals.
In the supplementary information of ref. 101, Hsu et al. introduced another bubbler device (Fig. 5e) capable of harvesting energy from machine motions. This device is located at the point where the rod string is attached to an oil pump head and can power the pump load cell and associated equipment. As the pump head moves up and down (about 0.2 Hz frequency), the force exerted by the rod string on the harvester's top plate alternates, causing periodic compression of the bellows chamber. This motion displaces the dielectric fluid contained in the chamber through the REWOD chip, generating electrical power. This device was estimated to generate more than 5 W of electrical power.
Traditional REWOD components rely on rigid planar electrodes, which limits their suitability for powering wearable sensors. For the next generation of these devices, continuous and sustainable operation under various bending conditions, without bias voltage, is essential. To address this limitation, Adhikari et al.48 developed a bias-free voltage REWOD component for powering wearable sensors. They used a 3D-printed polydimethylsiloxane (PDMS) and carbon black flexible substrate as the bottom REWOD electrode. The flexible substrate was first coated via electron-beam physical vapor deposition (EBPVD) with a ∼200 nm Ti conductive layer, followed by a ∼50 nm Cr adhesion bilayer, and then a ∼200 nm Al2O3 dielectric layer. The top conductive electrode was a doped Si wafer coated with a 100 nm Ti layer. Using a 50 μL electrolyte droplet, a maximum AC current of ∼340 nA was generated at 2 Hz. This study demonstrated the feasibility of implementing REWOD for powering wearable sensors using flexible electrodes.46
Coating on the polyimide substrate is more consistent compared to PDMS and exhibits greater thermal stability at high temperatures (>200 °C) encountered during the EBPVD deposition process.127 Kakaraparty et al. designed a REWOD device using a polyimide sheet (Kapton) as a flexible electrode.44 The top electrode was fabricated on a 0.11 mm polyimide sheet via EBPVD (see in Fig. 5f), with a 50 nm Cr layer followed by a 150 nm Ti layer. The bottom electrode was similarly coated with 50 nm Cr and 150 nm Ti, with an additional 100 nm Al2O3 dielectric layer. A 50 μL DI water droplet was encapsulated between the electrodes. When the electrodes were maintained in a planar configuration without any bias voltage, the REWOD generated a maximum power density of 0.002 μW cm−2 at 5 Hz. When the flexible electrodes were bent to a 60° angle, the power density reached 0.05 μW cm−2, about 25 times higher than that in the planar state. This is due to the reduced gap between the electrodes in the bent state, which increases the contact area between the electrodes and the electrolyte, thereby leading to a higher capacitance. Moreover, a narrower gap reduces electrical resistance across the electrodes, minimizing energy dissipation.127 This output is relatively high with no bias voltage and is sufficient to power ultra-low-power bio-wearable chips, which typically require only nanowatt-level power to operate.128
Kakaraparty et al.122 transitioned to a polyimide substrate for the flexible electrodes and employed a sputtering-based physical vapor deposition technique to deposit a high-dielectric metal oxide (MnO2) onto the polyimide sheet. This modification led to a significant increase in power density, reaching 476.21 μW cm−2 with a 50 μL droplet. These studies represent an important initial step toward developing a single electrolyte encapsulated flexible REWOD component, with the long-term goal of characterizing arrays of REWOD structures for future applications.
When considering the entire system, some researchers have begun integrating the REWOD component into circuitry to power motion sensors. Adhikari et al.66 developed a novel bias-free REWOD unit to power motion sensors, capable of generating an unconditioned AC output of 95–240 mV using a 50 μL droplet of 0.5 M NaCl electrolyte. When integrated with commercial components, this AC signal was rectified, boosted, and amplified. A seven-stage rectifier utilizing Schottky diodes with a forward voltage drop of 135–240 mV and a forward current of 1 mA converted the AC signal into DC voltage. The boost converter produced approximately 3 V DC, demonstrating the feasibility of this system in powering motion sensors.
Tasneem et al.26 reported a self-powered motion sensor based on REWOD energy harvesting. The energy harvester (as shown in Fig. 5g) includes a rectifier and a voltage regulator to provide the DC supply voltage to the analog front-end and the transmitter for wireless data transfer from the motion sensor. The on-chip circuitry includes a seven-stage voltage-doubler based rectifier, an amplifier, an analog-to-digital converter, and a transmitter. The recycled folded cascode based charge amplifier has a closed-loop gain of 53 dB within the bandwidth of 1150 Hz, which is suitable for detecting low-frequency motion signals. The wireless motion sensing device using REWOD is suitable for quantitatively monitoring the motion-related data as a wearable sensor.
Sah et al.129 also proposed a REWOD-based system where the REWOD-generated charge is amplified by a charge amplifier with a gain of 2 V/V to improve the signal-to-noise ratio and transmitted to a digital receiver. To prevent power flickering, the rectifier's filter circuit was integrated with a supercapacitor, ensuring a constant power supply for 5 minutes with a power conversion efficiency of over 80% at 1 Hz.
In recent years, temperature- and force-sensing networks have become increasingly important to meet the demands of electronic skins (e-skins) and artificial intelligence systems.27,140–143 Simultaneous sensing of temperature and force has gained significant attention to mimic the functionality of human skin. To achieve this, most existing designs rely on separate temperature and force sensors, thereby detecting multiple stimuli independently.144–146 However, developing a true multimodal sensor capable of measuring both temperature and force within a single sensing unit remains a challenge. While some researchers have proposed potential solutions,147,148 many of these approaches involve nanoscale fabrication or micro-level current/voltage signals, leading to complex manufacturing processes and post-signal processing. Consequently, there is a pressing need for a novel sensing strategy that offers high accuracy while maintaining a simplified architecture. Liu et al.27 developed a self-powered multimodal temperature and force sensor based on REWOD and the thermogalvanic effect in a K3[Fe(CN)6]/K4[Fe(CN)6] droplet, as shown in Fig. 6a. The droplet's deformation enables force detection, while the temperature difference across the droplet generates both an alternating pulse voltage and a direct voltage, enabling simultaneous sensing of external force and temperature. To demonstrate this capability, the researchers constructed an integral system in which the droplet sensor responds to both external temperature and force stimuli. The sensor exhibits a displacement detection sensitivity of at least 0.05 mm, while the minimum force detection sensitivity depends on the mechanical properties of the external support structure. Upon receiving an external stimulus (e.g., a single press), the droplet sensor converts it into a voltage signal, which is then amplified and processed by a microcontroller unit. Finally, four two-tone light-emitting diodes respond spontaneously to the stimuli, indicating the object's temperature and the magnitude of the applied force through changes in color and the number of LEDs activated. These results demonstrate the potential of the droplet sensor as a multimodal temperature and force sensor for artificial intelligence applications.
![]() | ||
Fig. 6 Application of REWOD as self-powered sensors. a) (i) Schematic structure and principle of the sensor; (ii) the voltage–time curve of a droplet-based sensor when coupled temperature and force stimuli is applied; (iii) schematic representation of the integral sensing system. Reproduced with permission.27 Copyright 2016, Wiley. b) (i) The structures of the electrode with different EHE area ratios; (ii) side view of coplanar REWOD configuration, where counter electrode and energy harvesting electrode both situate on the bottom plate, leaving design flexibility on top structures. Reproduced with permission.130 Copyright 2022, IEEE. |
Moyo et al.130 developed a facile coplanar REWOD structure for sensing applications. Traditionally, REWOD units adopt a two-plate configuration, consisting of a top counter electrode and a bottom energy-harvesting electrode coated with a dielectric layer.27,34,46 In contrast, the coplanar design integrates both the counter electrode and the dielectric-coated energy-harvesting electrode into a single planar structure. To maintain a constant ratio between the solid–liquid contact area on the bottom dielectric-layer-coated electrode and the total liquid–electrode contact area during the spreading and deformation of a centered droplet, the coplanar electrodes were patterned symmetrically in a “flower petal” configuration (as seen in Fig. 6b(i)). Fig. 6b(ii) and (iii) show side views of the coplanar REWOD configuration. A vibrating top plate (without a metal electrode) can be added to assist in controlling droplet deformation and spreading (the vibrating plate is not shown in Fig. 6b). The exposed top surface in this coplanar REWOD structure offers significant potential for expanding sensing applications, such as raindrop energy harvesting. Furthermore, when integrated with flexible/stretchable electrodes, the coplanar REWOD system shows promise for tactile sensing and motion detection in soft robotics.27,34,46,130,149
Although classical REWOD units require a bias voltage to charge electrons to improve output power, recent advancements have shifted toward bias-free operation to enhance the practical applicability of REWOD. Despite these advancements, further improvements in bias-free REWOD output energy are necessary. Scaling up power generation in such systems will require precise synchronization and control of thousands of droplets with respect to their positions and velocities.150 Currently, most studies are limited to proof-of-concept devices using only a single droplet. Future work is expected to focus on the development of electrolyte-enclosed structures incorporating multiple droplets,127 which would significantly increase the electrolyte–electrode interfacial area, thereby enhancing interfacial capacitance and overall power output. Furthermore, future research should further investigate the effects of dielectric film thickness, variations in electrolyte volume, and electrolyte composition on overall device performance, as well as the development of high-dielectric constant metal oxide-coated array structures.44,122 These efforts will collectively contribute to improved output power and energy conversion efficiency, advancing the development of practical and scalable REWOD systems.
In addition, advances in flexible electrode technology are essential. The development of flexible conductive materials with good electrical conductivity and mechanical stability, as well as the ability to accommodate deformations such as bending or stretching, can ensure the stable performance of REWOD systems under practical operating conditions. Furthermore, integrating flexible electrodes with micro-structured surfaces is expected to further enhance droplet control and improve energy conversion efficiency.
Moreover, eliminating reliance on costly materials and complicated fabrication processes is another critical factor for large-scale adoption. Enhancing output by evaluating commercial off-the-shelf metalized polymer films for REWOD or experimenting with spin coating to deposit low-cost dielectric layers for REWOD electrodes are promising directions for further research.45,46,151,152
While the basic working principle of TENGs can be described through contact electrification and electrostatic induction, a deeper and quantitative understanding of their current generation mechanism requires a shift to electromagnetic theory. Classical Maxwell's equations were originally formulated for static or uniformly moving media, and do not fully account for the dynamic interfacial charge redistribution and moving dielectric boundaries intrinsic to TENG operation. Recent theoretical advancements particularly those by Wang's group highlight the critical role of displacement current in TENGs, and propose extensions to the conventional Maxwell framework to incorporate moving dielectric interfaces and time-dependent surface polarization.
Maxwell's equations, formulated by James Clerk Maxwell, describe the fundamental interactions between electric field, magnetic fields, charge density and current density.164 In Maxwell's equations, the displacement field (D) is generally expressed as:165
D = ε0E + P | (15) |
D = ε0E + P + Ps | (16) |
D′ = ε0E + P | (17) |
∇·D′ = ρf − ∇·PS | (18) |
∇·B = 0 | (19) |
![]() | (20) |
![]() | (21) |
![]() | (22) |
Further, if the surface charge density function σs(r,t) on the material surfaces is defined by a shape function f(r,t), where time t accounts for the instantaneous shape of the material under external influence (Fig. 7), then the equation for Ps can be expressed as:165
Ps = −∇φs(r,t) | (23) |
![]() | (24) |
![]() | (25) |
![]() | ||
Fig. 7 a) A schematic illustration of a TENG connected to an external load, along with the corresponding coordinate system used for mathematical modeling and analysis. b) Schematic showing displacement current and conduction current in a TENG. Reproduced with permission.165 Copyright 2020, Elsevier. |
Additionally, the voltage across the TENG is determined by the path integral of the electromotive force.166
![]() | (26a) |
![]() | (26b) |
![]() | (27) |
To overcome these limitations, researchers have turned their attention to liquid-based TENGs, which utilize the fluid nature of liquids to achieve more reliable and efficient contact electrification. Their fluidity175 ensures full contact, reducing wear176 during the process. Humidity has little or even no impact, especially when using water.177 In this section, we will focus on liquid based TENGs, specifically liquid–solid TENGs, where a liquid interacts with a solid surface and liquid–liquid TENGs which involve contact between two immiscible liquids. Although contact electrification at liquid–gas interfaces have also been studied,178,179 research in this area remains limited. The low density of gas and the slow motion of liquid make it difficult to achieve sufficient contact electrification for direct observation.180
Since Lin et al. first introduced liquid–solid TENGs in 2013,182 this technology has undergone significant development. Initially they designed a system where a PDMS surface interacted with water inside a tank, generating electricity through periodic contact and separation. Over time, liquid–solid TENGs has evolved to capture energy from droplets, waves, and fluid motion.196–198
A thorough understanding of contact electrification at the liquid interface is critical for improving liquid based TENG performances. Xu et al.199 developed the electron-cloud-potential-well model, often termed the “Wang Transition” model, to describe electron transfer mechanisms at material interfaces. When two materials come into contact, electrons transfer due to an asymmetric double-well potential created by overlapping electron clouds (Fig. 9). As the materials separate, the transferred electrons remain trapped in the receiver material due to a potential barrier. However, as temperature increases, this barrier weakens, making it easier for electrons to escape.
![]() | ||
Fig. 9 Electron-cloud-potential-well model. Reproduced with permission.199 Copyright 2018, Wiley-VCH GmbH. |
Various theoretical and experimental studies have been carried out to better understand the mechanisms of contact electrification showing electron transfer dominance at the interface. At the atomic level, theoretical models confirm that electron transfer occurs during contact electrification.200 Nan et al. used density functional theory (DFT) to study how electrons transfer between water and seven commonly used polymers in liquid–solid TENGs. They analyzed factors like bonding style, functional groups, work function, and contact distance affecting electron movement. Their findings showed that the LUMO (lowest unoccupied molecular orbital)–HOMO (highest occupied molecular orbital) energy gap plays a key role in determining the strength of electron transfer.201,202 Similarly, molecular dynamics simulations have verified electron transfer at the Pt(111)–water interface in water–metal interactions.203 Experimental studies have further substantiated electron transfer in contact electrification and revisited the formation of the EDL, previously believed to be primarily driven by ion transfer.204,205 For water–polymer interactions, Nie et al.206 examined charge transfer in droplets in contact with a PTFE membrane. Their findings demonstrated that the observed charge transfer significantly exceeded predictions based solely on the ion-transfer model, indicating that electron transfer is the predominant mechanism at the liquid–solid interface, accounting for over 90% of total transferred charges.
Various studies including above ones have provided strong evidence for the “two-step” mechanism proposed by Wang et al.207 in explaining how the EDL forms at the liquid–solid interface. In the first step, electron transfer occurs when the liquid comes into contact with the solid surface. This process leads to the formation of a sparse distribution of surface charges, meaning that only a small fraction of atoms on the solid surface (i.e., approximately one in every 30000) carries a charge. In the second step, ions present in the liquid reorganize near the interface due to electrostatic forces. These ions arrange themselves in response to the surface charge, creating the EDL.
Building on the understanding of contact electrification and EDL formation at the liquid–solid interface, these factors need special consideration for optimizing the performance and stability of TENGs: material selection, surface modification, and environmental control. When choosing solid materials for liquid–solid TENGs, their position in the triboelectric series208 matters. This relative position in the triboelectric series helps determine how well they gain or lose electrons when in contact with a liquid. A stable electron and ion transfer direction at the interface improves charge transfer, making the TENG more efficient. PDMS, fluorinated ethylene propylene (FEP), polyethylene terephthalate (PET), and PTFE are commonly used solid materials in liquid–solid TENGs due to their strong hydrophobic nature and high tribo-negative characteristics. Likewise, using deionized (DI) water is preferable because ionic screening in more conductive solutions can hinder electron transfer.206,209,210 In general, as ion concentration increases beyond a certain threshold, excess ions accumulate near the solid surface, resulting in screening effects that weaken the interfacial electric field and suppress charge transfer efficiency.
The efficiency of charge transfer in liquid based TENGs is closely linked to surface properties. Hydrophilic surfaces favor ion-based transfer through ionization reactions, while hydrophobic surfaces primarily support electron transfer.211 Environmental factors, such as high temperatures, can further reduce performance by causing thermionic charge leakage.211 Despite ongoing research, the power output of liquid–solid TENGs remains limited typically below 1 W m−2 due to fundamental constraints at the interface.212 For instance, increasing the surface roughness of the hydrophobic surface can enhance charge transfer. However, if the roughness is too high, it may trap air at the solid–liquid interface, which reduces effective contact and limits electrification. As a result, optimizing wettability requires balancing the contact area and hydrophobicity. These interfacial factors ultimately restrict surface charge generation. Strategies to address such challenges are discussed in detail in section 3.2.
![]() | (28) |
In another related study, Wu et al.214 introduced a charge trapping-based electricity generator that employed an advanced homogeneous electrowetting-assisted charge injection method (Fig. 10a). By pre-charging hydrophobic fluoropolymer surfaces, they obtained an exceptionally high and stable negative surface charge density of 1.8 mC m−2. This innovative approach yielded unprecedented performance metrics, including instantaneous currents exceeding 2 mA, power densities surpassing 160 W m−2, and energy conversion efficiencies exceeding 11%. These results further validate pre-charging as an effective method to substantially enhance TENG performance.
![]() | ||
Fig. 10 Increase surface charges by a) charge injection using the homogeneous-EWCI method. Reproduced with permission.214 Copyright 2020, Wiley-VCH GmbH. b) Corona charging on the CYTOP surface. Reproduced with permission.217 Copyright 2020, Elsevier. c) Functional group impact on SiO2 surfaces using silane coupling agents. Reproduced with permission.218 Copyright 2020, American Chemical Society. Enhance liquid motion by d) laser direct writing method for PTFE surface wettability. Reproduced with permission.219 Copyright 2023, American Chemical Society. e) Reactive ion etching modifying the FEP surface. Reproduced with permission.220 Copyright 2021, Wiley-VCH GmbH. f) Novel SLIPS-TENG. Reproduced with permission.221 Copyright 2019, Oxford University Press. |
Jang et al.217 proposed a monocharged electret technique for improving surface charge density in liquid–solid TENGs (Fig. 10b). By employing a simple corona charging method to pre-inject stable negative charges into an amorphous fluoropolymer layer, they achieved an exceptionally high surface charge density (as demonstrated by the significant increase in electrical output). For instance, compared to pristine FEP which generated only ∼4 V, corona-charged FEP produced 17 V (∼4 times increase), while corona-charged cyclized perfluoro-polymer (CYTOPcorona) reached ∼130 V (∼33 times increase). This pre-charging approach led to a remarkable 1000-fold improvement in peak power output, increasing from ∼1 μW in pristine FEP to 1 mW in CYTOPcorona. Furthermore, the injected charges exhibited excellent stability, maintaining ∼80 V output even in highly conductive electrolyte solutions (e.g., 1 M NaCl) which typically degrade TENG performance. Even after two days, the charge-stabilized surface retained 15 times higher output voltage than uncharged FEP, confirming the long-term effectiveness of charge injection. Their results further illustrate the substantial potential of pre-charging techniques in enhancing TENG robustness and performance.
Where electrowetting-assisted charge injection and corona charging embed extra charge within the surface layer, an alternative route is to store charge in a subsurface capacitor and discharge it only at the moment of droplet contact. This internal-reservoir strategy avoids both surface ion screening and the lengthy waiting intervals required between successive impacts of droplets. Wu et al.222 exemplified this approach with the first externally charge-pumped liquid–solid TENG. A high-insulation ceramic layer beneath a super-hydrophobic Kapton/FEP film acts as a charge-storage capacitor that is continuously biased by a rotary triboelectric pump fitted with a coaxial polarity commutator (Fig. 11a). When a 50 μL water droplet spreads and briefly bridges the pre-charged ceramic to an offset aluminum strip, the stored electrons are released as a displacement-current burst. This design delivers a short-circuit current density of 50.31 μA cm−2 and a peak power density of 231.8 W m−2, a 1.43-fold improvement over the previous liquid–solid record; a single droplet can light 600 commercial LEDs, and a continuous droplet train sustains stable output for 30 min in a 3 × 3 array test. These results establish external charge pumping as a powerful means of elevating surface charge density and, consequently, overall TENG performance.
![]() | ||
Fig. 11 a) Externally charge-pumped liquid–solid TENG. Reproduced with permission.222 Copyright 2024, Wiley-VCH GmbH. b) Enhance TENG performance through charge trapping facilitated by photon-induced carrier generation.223 Copyright 2021, Wiley-VCH GmbH. c) Field-effect-enhanced droplet electricity generator.224 Copyright 2023, Wiley-VCH GmbH. |
After showing that a mechanical pump can “force-feed” the surface with charge from below, we next look at a strategy where visible light becomes the charger itself, growing fresh trap sites inside the tribo-layer every time it shines. Chen et al.223 boosted a PDMS-based TENG by embedding a g-C3N4/MXene–Au carbon-cluster composite into the tribo-layer. Under ≈60 mW cm−2 visible-light irradiation, g-C3N4 creates photo-generated carriers that act as charge traps, while Au nanoclusters (via surface-plasmon “hot” electrons) and highly conductive MXene sheets stabilize additional carriers, together building a dense trap network that allows much greater charge accumulation per cycle (Fig. 11b). A 4 cm × 4 cm device employing this composite delivered an output voltage of 510 V, a current of 80 μA, and a peak power of 20 mW at an impedance of 2 MΩ and output power density of 12.5 μW mm−2 representing roughly a six-fold improvement over an undoped PDMS control, while its internal impedance dropped by about 50%.
Surface charge density in triboelectric nanogenerators is typically governed by the intrinsic chemical and physical properties of materials, which sets a natural upper bound for output performance. To overcome this limitation, Shao et al.224 developed a field-effect-enhanced droplet electricity generator (FE-DEG), where a bias voltage is applied to a bottom electrode to regulate the surface potential at the dielectric interface. The device consists of a p-doped silicon bottom electrode, a thermally grown SiO2 dielectric layer (∼2 μm), and a spin-coated Hyflon top layer that provides hydrophobicity and serves as the contact surface (Fig. 11c). Upon applying a bias voltage, an electric field is established across the dielectric, inducing polarization and pre-conditioning the surface prior to droplet impact. This pre-polarized interfacial state enhances droplet–solid contact electrification, resulting in increased charge transfer. The authors demonstrated real-time modulation of charge output by adjusting the gate voltage, confirming dynamic tunability of device performance. Under optimized conditions, the FE-DEG achieved a peak instantaneous power density of 576.1 W m−2 at a load resistance of 6.2 kΩ. Additionally, the device exhibited stable output under humid environments and across a pH range of 4 to 11. The charge output of the FE-DEG exceeded that of several previously reported droplet energy harvesting devices, validating the effectiveness of the field-effect modulation strategy. By enabling electrostatic control of surface potential without altering material composition, this method introduces a robust and tunable mechanism for enhancing energy harvesting efficiency.
Lin et al.218 examined how different chemical functional groups affect electron and ion transfer in contact electrification between liquids and silicon dioxide (SiO2) surfaces. The researchers modified SiO2 using silane coupling agents with perfluorododecyl, oxypropyl, dodecyl, and aminopropyl groups to see how these changes impact charge transfer. They found that both electrons and ions contribute to contact electrification when DI water is used, but electron transfer is more affected by surface functional groups. Perfluorododecyl, which has a high electron affinity, tends to gain electrons, while aminopropyl, with low electron affinity, donates electrons. However, when organic liquids like cyclohexane and dichloromethane were used, electron transfer became the main mechanism since these liquids do not contain free ions. Importantly, the study found that electron transfer in organic liquids does not depend on functional groups because these liquids lack the ability to influence electron exchange significantly. To explain these effects, the researchers proposed an energy band model (Fig. 10c) showing how functional groups change electron energy barriers and charge transfer efficiency. When a solid like SiO2 comes into contact with a liquid, electron transfer occurs based on their energy band structures. In DI water, hydrogen bonding and dissolved ions create impurity states, which introduce additional energy levels. Since these states are at a higher energy level than the surface states of SiO2, electrons transfer from water to SiO2, leading to charge redistribution. This transfer can be influenced by surface functionalization with high electron affinity groups (e.g., perfluorododecyl) which introduce low-energy unoccupied states, increasing electron transfer from water to SiO2, whereas low electron affinity groups (e.g., aminopropyl) introduce high-energy occupied states, making SiO2 more likely to donate electrons to water. In contrast, organic solvents lack hydrogen bonding and impurity states, so charge transfer occurs through their molecular orbitals. When SiO2 comes into contact with an organic solvent, electrons move from its surface states to the LUMO of the organic molecules, causing SiO2 to become positively charged. However, because molecular orbitals are localized and discrete, electron transfer is highly limited (typically at most two electrons per molecule), and modifying the SiO2 surface has minimal effect on charge transfer with organic solvents. This distinction highlights how electron transfer in water is strongly influenced by surface functionalization, while in organic solvents, it follows a more restricted molecular orbital model. These changes affect how much charge stays on the surface after contact.
In another study, Vu et al.227 developed a high-performance liquid–solid TENG using a functionalized polyvinylidene fluoride (PVDF) membrane, modified with silica nanoparticles (SiNPs) and 1H,1H,2H,2H-perfluorooctyltriethoxysilane (FOTS), forming the FOTS/SiNPs/PVDF (FSiP) membrane. The proposed membrane integrates fluorine-bearing silane chains, which enhance interfacial polarization, significantly improving its hydrophobicity and dielectric constant. These enhancements lead to a 10.8-fold increase in power density to 420 mW m−2, along with improved current (5.79 μA) and voltage (28.3 V) output.
To achieve optimal performance, surface roughness must be carefully engineered to balance the trade-offs between effective charge generation and droplet mobility.187 Some nanostructured superhydrophobic surfaces (with increased roughness) have demonstrated higher electrical outputs, possibly because high-velocity droplet impacts collapsed the trapped air layer, increasing the contact area and enhancing the charge transfer.189,231 However, at lower droplet impact speeds, the persistent air layer can reduce the effective contact area and limit the extent of electrification.229 Thus, for practical applications, surface roughness should be carefully engineered to optimize triboelectric charge transfer while maintaining efficient solid–liquid interactions, ensuring stable and effective energy harvesting. The most common ways to increase the contact angle involve modifying the surface's morphology or adding texture layers, followed by coating with a low surface energy material such as fluoropolymers. This can be done through morphological changes like laser direct writing,232,233 etching220 and many other ways.
The femtosecond laser direct writing method is a highly effective technique for modifying surface roughness, enhancing both charge transfer efficiency and triboelectric performance in liquid-based TENGs. This process creates a porous micro/nanostructured layer on hydrophobic solid surfaces, significantly improving their hydrophobicity and surface charge density without altering the material's chemical composition. Zhang et al.219 applied a line-by-line femtosecond laser scanning approach using a 1035 nm wavelength, 10 kHz pulse frequency, and 350 fs pulse width, with an optimal laser power of 1.07 W (Fig. 10d). This treatment increased the water contact angle from 105° to 160°, preventing liquid residues and improving self-cleaning properties. More importantly, the modification played a crucial role in charge transfer dynamics by optimizing the contact area. While the superhydrophobic nature of the treated PTFE initially reduced droplet contact at low speeds, its porous microstructure expanded the effective contact area at high impact speeds, leading to greater charge induction. Additionally, the rapid spread and retraction of droplets on the modified surface further enhanced triboelectric output by accelerating charge transfer. Due to the enhanced contact electrification, the open-circuit voltage and short-circuit current output increased by factors of 3 and 1.5, respectively.
Wang et al.220 demonstrated the effectiveness of reactive ion etching in modifying FEP surfaces to achieve superhydrophobic properties (Fig. 10e). Using reactive ion etching treatment in a plasma-therm chamber for 15 minutes, they created a highly water-repellent surface with a contact angle of 168.5°. The modification played a crucial role in droplet-based energy harvesting, significantly improving the anti-wetting behavior of the surface. One of the major advantages of this superhydrophobic FEP surface was its ability to prevent droplet coalescence, ensuring that each droplet detached instantly instead of forming a continuous liquid layer. This property was particularly beneficial under high-frequency droplet impacts exceeding 140 Hz, where maintaining individual droplet separation was essential for maximizing energy conversion. A direct comparison between the superhydrophobic surface-based droplet electricity generator (SHS-DEG) and a conventional droplet electricity generator (DEG) demonstrated a significant enhancement in output performance. Specifically, the average voltage output of SHS-DEG reached 14.63 V, which was three times higher than that of DEG, while the average current output increased to 3.98 μA, which was twice that of DEG.
Unlike superhydrophobic surfaces, SLIPS-TENGs maintain stable electricity generation even in harsh environments due to their unique liquid-infused surface, which forms a smooth and continuous liquid–liquid interface. Xu et al.221 in 2019 introduced a SLIPS-based TENG by incorporating a perfluorinated lubricant (PFPE) into a porous PTFE structure (Fig. 10f). They present SLIPS-TENG as a superior alternative to conventional superhydrophobic surface-TENGs (SHS-TENGs), addressing issues like surface wear, ice formation, and contamination. By integrating a lubricant-infused surface, the SLIPS-TENG enhances mechanical durability, charge transfer efficiency, and environmental stability. A key discovery is the charge transparency phenomenon, where a critical lubricant layer thickness maintains charge transfer like solid–liquid interfaces, akin to wetting transparency in graphene. Unlike SHS-TENGs, which suffer from droplet bouncing and reduced charge efficiency, the SLIPS-TENG ensures sustained droplet contact, leading to higher energy conversion rates. Its self-healing ability enables stable power generation even after surface damage. Additionally, it excels under extreme conditions, including low temperatures (−3 °C), high humidity, and submerged environments, where SHS-TENGs typically fail. Experimental results show an order-of-magnitude improvement in energy harvesting, with the SLIPS-TENG charging a 1 μF capacitor to 5 V within 55 seconds at −3 °C, demonstrating its efficiency under harsh environmental conditions.221
![]() | ||
Fig. 12 a) First bulk effect droplet-based electricity generator. Reproduced with permission.90 Copyright 2020, Springer Nature. b) Transistor-like design for droplet energy harvesting. Reproduced with permission.234 Copyright 2023, Elsevier. c) Universal single-electrode configuration. Reproduced with permission.235 Copyright 2021, Elsevier. |
The transistor-like architecture, originally introduced by Xu et al.,90 revolutionizes the charge transfer process in liquid–solid TENGs by shifting from a surface-dominated mechanism to a bulk-driven charge transfer process. This design consists of a PTFE tribo-surface sandwiched between an ITO bottom electrode and an aluminum strip top electrode. Unlike conventional TENGs, where charge transfer is confined to the liquid–solid interface, this bulk-driven approach enables charge movement through the entire water droplet, significantly enhancing energy conversion efficiency. This setup allows the water droplet to act as a conductive bridge, simultaneously interacting with the PTFE surface, ITO electrode, and aluminum electrode, thereby increasing the effective charge separation and transfer (Fig. 12a). As a result, this transistor-like droplet energy generator achieves an energy conversion efficiency of approximately 2.2%, which is several orders of magnitude higher than that of traditional single-electrode liquid–solid TENGs. Furthermore, this architecture facilitates an impressive charge transfer of 49.8 nC per 100 μL droplet, making it one of the most efficient designs for high-output energy harvesting in droplet-based TENGs.
Several optimization strategies have been proposed to enhance transistor-like droplet electricity generators. Wang et al.236 found that incorporating a CYTOP intermediate layer between the tribo-surface and bottom electrode improves charge retention, with factors such as increased thickness, roughness, and surface charge injection all contributing to improved performance. Zhang et al.234 investigated the effects of electrode positioning, demonstrating that placing the top electrode directly at the droplet's maximum spreading radius (d = 0) or at zero vertical distance (h = 0) from the tribo-surface shown in Fig. 12b results in the strongest polarization and highest charge density. Extending the bulk effect beyond individual droplets, Zhang et al.237 proposed a water-column-based energy harvester, in which a 10 cm water column slides through a tilted PTFE tube, achieving a record-breaking output of 904 V, 509 μA, and 118.1 mW at ∼2 MΩ load. This approach highlights the scalability of bulk-effect-based energy harvesting beyond single droplet interactions.
Building upon the advancements in transistor-like architectures, the lubricant-armored transistor-like electricity generator238 introduces a novel approach to bulk-effect-based energy harvesting by integrating a SLIPS with a transistor-like electrode configuration. This design incorporates a lubricated PTFE membrane, which forms a low-friction, transparent interface that enhances droplet movement during droplet impact. As a droplet comes into contact with both the top and bottom electrodes, it momentarily establishes a closed-loop circuit, facilitating effective charge redistribution analogous to a transistor gating mechanism. This architecture significantly improves energy conversion efficiency while ensuring robustness in challenging environmental conditions, such as high salinity, extreme pH variations, and fluctuating humidity. Under optimized conditions, featuring a 46 μm SLIPS layer thickness and a 0.1 M NaCl solution, the lubricant-armored transistor-like electricity generator has demonstrated an instantaneous power density of 118 W m−2, a peak voltage of 65 V, and a current of 200 μA, outperforming conventional SLIPS-based TENGs.238 Additionally, its scalable structure, developed on a printed circuit board (PCB) substrate, minimizes wiring complexity and enhances integration, making it well-suited for large-scale water energy harvesting applications.
To increase design flexibility, Zhang et al.235 introduced the universal single-electrode liquid–solid TENG, which eliminates the bottom electrode, allowing any solid tribo-surface with a top electrode to function as an energy harvester (Fig. 12c). Here, charge transfer occurs through the solid surface, droplet, top electrode, and ground, with water droplet movement serving as the switching mechanism for charge exchange. The output voltage V in this system follows the equation:
![]() | (29) |
A comparative study by Li et al.239 analyzed the performance differences between universal single-electrode and transistor-like liquid–solid TENGs. Their findings indicate that universal single-electrode devices exhibit lower internal impedance, enabling them to operate efficiently with smaller optimal load resistances. Furthermore, an increase in water-to-ground capacitance extends the electrical relaxation time, leading to enhanced energy harvesting efficiency. This distinction highlights that while the transistor-like design excels in absolute power output, the universal single-electrode configuration provides greater adaptability and reduced impedance constraints.
While bulk-effect-based TENGs typically employ vertically stacked, transistor-like architectures to guide charge flow across dielectric interfaces, recent developments in coplanar electrode design present an alternative route for achieving high output through electrostatic interaction and droplet-bridging-triggered charge redistribution. In the droplet energy harvesting (DEH) panel developed by Xu et al.,240 both source and drain electrodes are positioned on the same substrate surface in a planar layout. The source electrode is fully encapsulated by a FEP dielectric, while the drain remains exposed (Fig. 13a). When a water droplet lands on the FEP-covered source region, triboelectric charge separation occurs at the dielectric–droplet interface, generating a negative surface charge on the FEP and a corresponding positive charge on the droplet. As the droplet spreads to come into contact with the adjacent drain electrode, it forms a transient conductive bridge that facilitates electron flow through the external circuit (Fig. 13b). This electrostatic bridging event avoids vertical capacitive stacking and thus minimizes parasitic capacitance, resulting in faster circuit response and reduced energy loss.
![]() | ||
Fig. 13 (a) Schematic of the droplet energy harvesting (DEH) panel and its transistor-inspired unit cell. Reproduced with permission.240 Copyright 2022, Royal Society of Chemistry. (b) Working mechanism illustrating charge transfer from the source to the drain upon droplet spreading, followed by charge recovery as the droplet retracts and exits. Reproduced with permission.240 Copyright 2022, Royal Society of Chemistry. (c) (i) Conceptual framework of TCNG systems integrating power generation and management. (ii) Representative charge induction steps in TC-TENG operation. (iii) Bioinspired TC-TENG design featuring dual charge reservoirs and neural-like regulation. Reproduced with permission.241 Copyright 2023, Royal Society of Chemistry. |
Beyond its architectural elegance, the DEH panel demonstrates strong spatial adaptability and system-level efficiency. Xu et al.240 designed two panel-topology strategies: a checkerboard electrode layout for structured droplet impacts, and a strip-like U-shaped drain configuration optimized for spatially random rain events. The A4-sized panel supports asynchronous, position-independent droplet activation, with each event triggering a localized output without signal interference. Instead of assigning a separate rectifier to each unit, all outputs are routed through a single full-wave bridge rectifier. This is possible because each droplet-induced current pulse lasts only a few milliseconds. Natural droplet impacts occur at low frequencies with minimal overlap, allowing the system to combine signals without interference or loss. The device achieves an average open-circuit voltage of 266.6 V and a short-circuit current of 273.6 μA. It is further shown to successfully power a wireless forest monitoring system including temperature and smoke sensors by charging a 10 mF capacitor to 10.69 V under natural rainfall.
While vertically stacked and coplanar TENG architectures, such as Xu et al.'s240 droplet energy harvesting panel, have improved output via enhanced spatial charge collection and reduced parasitic losses, they still rely on AC output and require external rectifiers. As a further advancement in electrode configuration, Dong et al.241 proposed a bioinspired total current nanogenerator (TCNG) that achieves DC output without rectification, drawing design cues from the electrical discharge organs of rays. By embedding voltage-gated switching into the electrode system, the TCNG eliminates the need for external rectifiers and synchronization, offering a scalable solution for high-power DC output with simplified circuit design and easier integration into power conversion modules (PCMs).
The TCNG features a multi-channel structure consisting of eight charge-collecting needles (CCNs) positioned above a PTFE triboelectric surface (Fig. 13c). Water droplets serve as charge carriers, generating triboelectric charges upon contact and transferring them to the CCNs and a shared bottom electrode. A gas discharge tube (GDT) connects the electrodes and functions as a voltage-controlled switch, initiating synchronized discharge once the accumulated potential exceeds ∼600 V. This configuration enables asynchronous charge accumulation and timed energy release, producing stable, rectifier-free DC output.
The device operates through five stages: charge splitting, negative charge transfer, spatiotemporal charge separation, positive charge transfer, and charge neutralization. During this cycle, displacement current (Id) builds up and is converted into conduction current (Ic) upon GDT activation, with the total output defined as It = Id + Ic. The TCNG achieves an open-circuit voltage of 3000 V, a short-circuit current of 12 mA, and demonstrates practical capability by powering 1260 LEDs in a single discharge cycle.
More recently, Li et al.242 advanced the TCNG architecture by introducing a fully integrated droplet-based nanogenerator system designed for real-world applications. Their platform combines a water-charge-shuttle design with a field effect-enhanced triboelectric interface, a DC–DC buck converter for power management, and physical compatibility with solar panels. This system reaches a record-high open-circuit voltage of 4200 V, powers 1440 LEDs, and supports smart greenhouse applications such as Bluetooth monitoring and thermohygrometer operation. These results demonstrate how total current based architectures can evolve from single-event demonstrations into scalable, field-deployable energy platforms, reaffirming the importance of electrode structure, charge routing, and system-level integration in next-generation liquid–solid TENG design.
The high-density design approach aims to increase the power density by arranging multiple liquid based TENG units in a structured array. This configuration ensures continuous charge transfer by enabling multiple points of contact. A recent advancement in this area was demonstrated by Gu et al.244 who developed a wave energy harvester using 3D electrodes that interact with moving water (Fig. 14a). The high-density electrode array increases the number of active liquid–solid TENG units per unit time by optimizing contact with the flowing liquid, leading to improved charge transfer and enhanced energy harvesting. This system achieved an instantaneous power density of 11.7 W m−2, demonstrating the effectiveness of structural innovations in boosting mechanical energy harvesting efficiency.
![]() | ||
Fig. 14 a) Wave energy harvester utilizing a liquid–solid TENG planar array. Reproduced with permission.244 Copyright 2021, Elsevier. b) Stacked configuration of a liquid–solid TENG and solar cell. Reproduced with permission.245 Copyright 2023, Wiley-VCH GmbH. |
Another promising approach is hybrid integration, where liquid–solid TENGs are combined with complementary energy harvesting technologies, such as solar cells, piezoelectric nanogenerators, or electromagnetic generators. This synergy bridges the energy gap during periods when the liquid–solid TENG operates below the optimal conditions. For instance, Ye et al.245 recently introduced a vertically stacked liquid–solid TENG and solar cell hybrid system, where a transparent liquid–solid TENG was placed above a solar cell. This configuration not only maintained light transmission but also boosted power density from 37.03 mW m−2 to 40.80 mW m−2 under simulated rainy conditions (Fig. 14b). Such hybrid designs leverage multiple energy sources efficiently, ensuring a more stable and enhanced power output. Both high-density and hybrid structural designs continue to evolve, offering innovative pathways to enhance the practicality and efficiency of liquid–solid TENGs for real-world applications.
A notable example of droplet electricity TENGs is the cactus-inspired droplet electricity generator,158 which enhances droplet motion and charge transfer through an asymmetric amphiphilic surface. Inspired by cactus spines and Namib Desert beetle elytra, the system features an amphiphilic cellulose ester coating that provides hydrophilic sites for droplet nucleation from fog and hydrophobic regions for efficient removal (Fig. 15a). A hydrophobic FEP channel guides the droplets smoothly toward the collection area. The spines are inclined at 60 degrees, optimizing Laplace pressure to drive droplets toward the hydrophobic channel. As droplets impact the droplet electricity TENG surface as shown in Fig. 15a(ii), they generate triboelectric charges, leading to electrostatic adsorption. This process enables the TENG to generate an open-circuit voltage of 103.2 V, capable of lighting 400 commercial light emitting diodes (LEDs), while maintaining a surface charge of 42.6 nC after 770 droplets. By combining bioinspired structural engineering with triboelectric charge transfer, the system achieves a high water-harvesting efficiency of 93.18 kg m−2 h−1, making it one of the most effective triboelectric-enhanced water collection technologies.158
![]() | ||
Fig. 15 Droplet electricity generator. a) Cactus inspired droplet electricity generator. Reproduced with permission.158 Copyright 2022, Springer Nature. b) Raindrop energy-powered autonomous wireless hyetometer. Reproduced with permission.159 Copyright 2022, Springer Nature. c) A hybrid solar panel with a TENG array for simultaneous raindrops and solar energy harvesting. Reproduced with permission.245 Copyright 2023, Wiley-VCH GmbH. d) Rainwater energy captured using a combination of pre-charged and grounded liquid films. Reproduced with permission.29 Copyright 2019, Springer Nature. e) Liquid–liquid TENG system for rainwater energy harvesting, utilizing HFE for electron transfer and electrostatic induction for displacement current generation. Reproduced with permission.157 Copyright 2022, Wiley-VCH GmbH. f) A self-repairing hydrophobic textile designed for efficient water droplet energy harvesting. Reproduced with permission.246 Copyright 2021, American Chemical Society. |
Another examples is the raindrop TENG developed by Xu et al.159 for rainwater energy harvesting (Fig. 15b). Using PCB technology, they fabricated a copper electrode array coated with a hydrophobic PTFE layer, ensuring smooth droplet rolling. Operating in single-electrode mode, this device harnesses charge transfer as droplets move across the PTFE surface, generating a potential difference that drives current through an external circuit. Under a rainfall intensity of 71 mm min−1, this TENG exhibited an average short-circuit current of 15 μA, an open-circuit voltage of 1800 V, and a peak output power of 325 μW. Similarly, Ye et al.245 developed a raindrop-TENG optimized for raindrop energy harvesting (Fig. 15c). By layering a FEP film on an ITO electrode, they achieved efficient triboelectric charge transfer as raindrops rolled over the surface. To prevent signal disruption caused by excessive droplet impacts during heavy rainfall, they refined the array structure of the raindrop-TENG by optimizing electrode arrangement, spatiotemporal separation of raindrops and the panel tilt angle to minimize charge cancellation and enhance energy output. This optimization resulted in a consistent average power density of up to 40.8 mW m−2.
Expanding based upon droplet electricity triboelectric systems, the liquid–liquid TENG has emerged as another promising alternative, leveraging liquid interfaces to enhance charge transfer and enable self-replenishing operation. Nie et al.29 introduced a liquid–liquid TENG that utilizes precharged and grounded liquid membranes to harvest energy from falling water droplets. Unlike conventional liquid–solid TENGs, this system achieves energy conversion entirely at the liquid–liquid interface, offering unique advantages in permeability and self-healing capability. The permeability of the liquid membrane which let droplets pass through continuously ensures uninterrupted energy production. This sets it apart from conventional solid surfaces, which can trap leftover liquid and reduce charge collection efficiency. Moreover, the membrane's self-healing nature allows it to regain its structure after each droplet impact, enabling consistent operation. The design consists of two liquid membranes: a pre-charged liquid membrane and a grounded liquid membrane (Fig. 15d). The pre-charged membrane accumulates positive charges induced by an adjacent negatively charged FEP film, which is pre-charged through triboelectric contact with a nylon film. The grounded membrane serves as a charge collector. When a neutral water droplet comes into contact with the pre-charged membrane, the positive charges redistribute between the droplet and the membrane. As the droplet detaches, it removes a portion of the charge, disturbing the electrostatic balance. To restore equilibrium, charges are replenished from the ground to the liquid membrane, inducing a current flow and enabling continuous energy generation. This system demonstrated a peak power output of 137.4 nW for 40 μL droplets, making it highly effective for energy harvesting in rainwater collection, irrigation currents, microfluidic systems, and electrostatic charge removal from solid surfaces.
In another example, Zhang et al.157 introduced a liquid–liquid TENG utilizing hydrofluoroether (HFE7500) fluid as a triboelectric material to efficiently charge rainwater droplets via electron transfer (Fig. 15e). As droplets pass through HFE7500, electrons transfer from water to the fluid, progressively increasing the positive charge of the droplets. A ring-shaped electrode positioned around the droplet path generates a displacement current via electrostatic induction, ensuring efficient energy conversion. Due to the fluidity of the triboelectric layer, the contact surface is constantly refreshed, preventing charge saturation and maximizing energy output. The system achieves a charge density of 3.63 μC L−1 and sustains triboelectric charge exchange for up to 1200 seconds before saturation. With a crest factor (evaluates stability of the output signal) of approximately 1.1, the device exhibits stable output and is capable of generating DC, eliminating the need for rectifiers and reducing energy loss from switching components. Beyond rainwater energy harvesting, this approach is adaptable to microfluidic systems, hydropower applications, and self-powered environmental sensors, paving the way for scalable liquid-based energy solutions.
Advancements in materials science have greatly enhanced the efficiency of droplet electricity TENGs. One such innovation is the all-fabric triboelectric nanogenerator, which incorporates a superhydrophobic surface made of SiO2 nanoparticles, polyvinylidene fluoride–hexafluoropropylene (PVDF–HFP), and perfluorodecyltrichlorosilane (FDTS) coatings (Fig. 15f). This multi-layered fabric structure provides high breathability, self-repairing hydrophobicity, and long-term durability. The device efficiently converts raindrop impacts into electrical energy through triboelectric charge transfer. Compared to conventional polymer-based triboelectric nanogenerators, the fabric TENG delivers stable performance even under prolonged environmental exposure, making it a promising solution for wearable electronics and outdoor energy harvesting applications.246 Another example is the waterproof and fabric-based multifunctional TENG, which combines water-repellent ethylene-vinyl acetate friction layers with conductive and mesh fabric layers.247
![]() | ||
Fig. 16 Bulk liquid based generator. a) Illustration of the OS-TENG structure and its potential for large-scale ocean wave energy harvesting. Reproduced with permission.249 Copyright 2020, Elsevier. b) Asymmetric plate structure of the dielectric layer-based TENG. Reproduced with permission.185 Copyright 2023, Wiley-VCH GmbH. c) Schematic of liquid metal-TENG harvesting energy through continuous water flow in and out. Reproduced with permission.250 Copyright 2023, Wiley-VCH GmbH. d) Liquid-metal-based TENG optimized for low-frequency and multidirectional vibrations. Reproduced with permission.184 Copyright 2021, Frontiers Media S.A. |
Munirathinam et al.250 introduced a liquid-metal-based TENG for energy harvesting from flowing water in pipes/channels. It features a Galinstan liquid-metal electrode encapsulated within a PDMS friction layer, utilizing a liquid–solid interface for triboelectric charge generation (Fig. 16c). It generates electricity through contact electrification and electrostatic induction as flowing water interacts with the PDMS layer, creating an electrostatic potential difference that induces charge transfer in the Galinstan liquid-metal electrode. This liquid-metal based TENG exhibits superior performance with a peak output of 6.2 V and 3.6 μA at a flow rate of 2.5 L min−1, attributed to its flexibility and larger contact area. As water flows over the PDMS friction layer, the PDMS deforms and stretches, causing a reduction in its thickness due to the Poisson effect.251 This decrease in thickness shortens the distance between the triboelectric charges in water and the Galinstan electrode, thereby increasing the electrostatic force acting on the electrode. Since electrostatic force is inversely proportional to the separation distance, a thinner PDMS layer enhances charge induction, leading to improved energy conversion efficiency. Additionally, the low Young's modulus of Galinstan allows it to stretch in response to PDMS deformation, further increasing the effective contact area between the electrode and the friction layer. This dynamic adaptation enhances triboelectric charge generation and charge transfer efficiency, leading to higher electrical output. Conversely, if the PDMS layer is too thick, the increased separation distance weakens the electrostatic attraction between the triboelectric charges and the electrode, thereby reducing the number of induced charges and lowering overall device performance. This design is highly effective for remote water flow monitoring in pipelines. In addition, it is also capable of powering LEDs and liquid crystal displays. Its ability to resist corrosion and prevention of electrode cracking ensures extended durability, even under low flow rates and continuous operation.
In another example, Deng et al.184 developed a liquid-metal-based TENG optimized for low-frequency and multidirectional vibrations (Fig. 16d). The fluidic nature of the liquid metal enhances its responsiveness to vibrations, maximizing energy conversion. The device operates in freestanding mode, meaning the liquid metal is not fixed to any electrode but moves freely within a sealed chamber, functioning as an independent triboelectric layer. It generates electricity through contact electrification and electrostatic induction as the liquid metal moves freely within a sealed chamber, repeatedly making and breaking contact with the Kapton triboelectric layers due to external vibrations. This contact-separation motion induces charge transfer, creating an electrostatic potential difference, which drives electron flow through the external circuit, resulting in AC output. By combining freestanding movement with contact-separation charge generation, the device achieves a peak voltage of 252 V. It successfully charged a 10 μF capacitor to 6.46 V in 60 seconds at 7.5 Hz and demonstrated its capability by powering 100 LEDs during walking tests. With a power density of 33000 mW m−3, the device is well-suited for applications in marine buoys and vehicle suspensions, offering an effective approach for vibration energy harvest.
![]() | ||
Fig. 17 Biomedical monitoring and sensing. a) Liquid–solid TENG used for blood flow monitoring, designed to break down and be absorbed by the body over time. Reproduced with permission.252 Copyright 2021, Wiley-VCH GmbH. b) A self-powered flow sensor for monitoring biofluids in the healthcare system. Reproduced with permission.186 Copyright 2020, American Chemical Society. c) A dual-signal self-powered device utilizing a solid–liquid TENG for precise dopamine concentration detection. Reproduced with permission.156 Copyright 2019, Wiley-VCH GmbH. d) Triboelectric nanosensors for label-free screening of anti-tumor drugs. Reproduced with permission.253 Copyright 2024, Elsevier. |
A further medical-oriented design involves a tube-shaped liquid–solid TENG with a circular cross-section to monitor clinical drainage procedures.186 This device employs a superhydrophobic SiNPs layer as the triboelectric surface and trimmed copper electrodes on the inner wall of a silicone rubber tube. By connecting the tube to a smaller-diameter drip line, pulsed current outputs are different for various liquids such as HCl, urine, blood, NaOH, and PBS demonstrating its potential for clinical IoT applications. As shown in Fig. 17b(i), the triboelectric layer, along with trimmed copper electrodes, is positioned on the inner wall of a round silicone rubber tube. This tube is connected to a smaller-diameter silicone tube, which facilitates the flow of test liquids in droplet form through a tubular drop counter. When these droplets pass through the drop counter, they generate a pulsed current output, as depicted in Fig. 17b(ii). The recorded current variations correspond to different liquid droplets, demonstrating the system's ability to differentiate between fluids. These findings highlight the potential of the fabricated droplet counter for clinical monitoring applications, as illustrated in Fig. 17b(iii).186
Jiang et al.156 introduced a dual-signal dopamine detection sensor utilizing a single-electrode solid–liquid TENG, composed of a PTFE film, copper electrode, and glass substrate. Dopamine detection is facilitated by its self-polymerization into polydopamine (PDA) on the PTFE surface which modifies surface charge properties and hydrophilicity, leading to distinct electrical responses. Two primary signals were generated: ITENG, resulting from contact electrification and electrostatic induction at the oil–water–PTFE interface, and Iinterface arising from electrostatic effects within the oil–water interface (Fig. 17c). Experimental findings demonstrated that contact angle measurements decreased with increasing PDA concentration (0–500 μM L−1), confirming effective surface modification. Furthermore, the dual-signal detection approach yielded detection limits of 5.15 μM for ITENG and 3.96 μM for Iinterface, establishing a sensitive and selective dopamine sensing platform.
A TENG-based nanosensor253 (Fig. 17d) was developed for high-throughput drug screening by detecting electrical signal variations caused by molecular interactions. This platform measures voltage changes when molecules bind to its surface, making it effective for identifying drug–protein interactions. Using the FK506-binding protein (FKBP) rapamycin system as a model, the sensor successfully detected drug interactions with the autophagy-related cysteine protease ATG4B. It identified S130 and tioconazole as inhibitors, while dexamethasone exhibited no binding effect. To validate the findings, Kelvin probe force microscopy measured surface potential shifts, confirming how drugs alter the charge distribution on the sensor. Additionally, enzyme activity assays verified the inhibitory effects of certain compounds, ensuring precise drug screening. This self-powered platform provides a sensitive, reliable, and cost-effective solution for drug discovery.253
![]() | ||
Fig. 18 Applications of liquid-based TENGs in sensors. a) Detection and classification of microplastics using deep learning integrated with liquid-based TENG technology. Reproduced with permission.155 Copyright 2023, American Chemical Society. b) Working mechanism of a TiO2-based liquid–solid triboelectric nano-sensor for catechin detection. Reproduced with permission.255 Copyright 2020, Elsevier. c) Working mechanism of a capillary-tube triboelectric nanogenerator. Reproduced with permission.256 Copyright 2018, Wiley-VCH GmbH. d) Rainfall sensing: establishing an approximately linear relationship between rainfall intensity and the device's electrical output. Reproduced with permission.29 Copyright 2019, Springer Nature. |
Chatterjee et al.255 developed a triboelectric nanosensor based on liquid–solid contact electrification for catechin detection, while also optimizing liquid–solid TENG performance (Fig. 18b). The sensor uses TiO2 nanosheet arrays as the solid triboelectric material, with water, ethanol, and acetone serving as the liquid contact media. Catechin molecules interact with the TiO2 surface, forming a ligand-to-metal charge transfer complex, which alters surface charge density and improves electron transfer efficiency. This interaction leads to a measurable increase in voltage, reaching 1.2 V at 80 μM catechin concentration. Ultraviolet photoelectron spectroscopy analysis confirmed that catechin adsorption reduced the TiO2 work function from 6.54 eV to 5.52 eV, indicating improved charge transfer.
Similarly, Gao et al.254 developed a TENG-based self-powered urea sensor with high sensitivity and specificity by incorporating enzyme-catalyzed reactions into a dual-electrode configuration. The sensing mechanism relies on the hydrolysis of urea by urease, which alters the pH of the solution. This pH change influences the triboelectric charge transfer between the PDMS surface and the liquid, modifying the electrical output of the TENG. The device effectively detects urea concentrations as low as 4 μM and remains unaffected by common fertilizer ions. High-performance liquid chromatography validation confirmed a strong correlation between output voltage and urea concentration, demonstrating its applicability in agricultural.
Chen et al.256 further expanded TENG sensing to microfluidics, creating a capillary-tube triboelectric nanogenerator (ct-TENG) for non-destructive microliter sampling (0.5 μL) (Fig. 18c). By harnessing Maxwell's displacement current from liquid flow, electrical outputs indicated variations in total aerobic count (TAC) in barreled water. The ct-TENG senses TAC by detecting changes in its electrical signals (voltage, current, charge transfer) as microbial growth increases. Bacteria release charged byproducts that weaken the triboelectric effect, causing a progressive decline in output signals. Over time, increased TAC led to decreased current, voltage, and charge transfer signals, enabling determination of optimal drinking intervals before water quality deteriorated.
In the maritime domain, Wang et al.68 designed a ring-shaped TENG for ship tilt angle detection. The device consists of a PTFE ring tube with an external copper electrode. An internal liquid layer moves as the ship tilts, causing contact electrification and generating electrical signals. The open-circuit voltage, short-circuit current, and transferred charge correlate with the tilt-induced arc length of fluid displacement. An integrated LED provides real-time feedback, while an alarm system warns the crew if the tilt angle surpasses a critical threshold, ensuring timely evacuation measures.
Durability remains a critical challenge for liquid based TENGs, as factors like electrode corrosion, material wear, contamination, and wetting258 can degrade performance over time. Mechanical deformation and abrasion can lead to electrode delamination and tribo-surface degradation, particularly in wearable and portable applications. Strategies to enhance durability include material engineering for stronger tribo-surfaces, hydrophobic modifications to prevent wetting and contamination, and corrosion-resistant electrode designs.259,260 Additionally, SLIPS have emerged as an effective solution, offering self-healing properties, anti-icing capabilities, and stable performance under extreme environmental conditions.221,261 These approaches are essential for improving the longevity and reliability of liquid based TENGs in real-world applications.
Environmental conditions play a crucial role in the performance of liquid based TENGs and should be carefully considered. Temperature is a key factor where liquid based TENGs exhibit a decline in electrical output as temperature increases.209,262 This reduction occurs because higher temperatures accelerate electron dissipation into the environment, following the principles of thermal emission theory. Light is another influential factor, typically enhancing TENG performance.263 Exposure to photons increases the concentration of electron–hole pairs, boosting charge generation.264 The extent of this enhancement depends on wavelength and light intensity. The shorter wavelengths and higher light intensity lead to greater output performance since they produce higher luminous flux, resulting in more electron–hole pair generation at the triboelectric interface.
Researchers have explored various strategies and advancements aimed at enhancing TENG performance and facilitating its industrialization. In section 3.2, we discussed in detail the methods to improve the performance of TENGs including increasing surface charges, facilitating interface charge transfer and enhancing liquid motion. Additionally, the bulk effect phenomenon, where an additional top electrode is incorporated to improve charge transfer efficiency, was examined. To further boost TENG output, high-density designs have been explored, where multiple arrays of TENGs are connected to amplify overall energy generation. Moreover, integrating TENGs into hybrid systems, such as solar or wind energy harvesters, can further optimize performance and expand practical applications. Another promising approach involves hybrid TENG designs that combine TENGs with electromagnetic generators or other energy-harvesting methods. These hybrid configurations help broaden frequency response and improve overall power output. However, the complexity and high volume of such hybrid systems remain as challenges, limiting their practicality. To address this, advancements in compact circuit designs and refined power management strategies are essential for maximizing overall system efficiency. While significant progress has been made in TENG technology, further innovations and optimizations are required to achieve large-scale commercialization. Continued research into materials, design scalability, and energy storage integration will be key to bridging the gap between lab-scale prototypes and real-world applications.
As the liquid based TENGs developed over time, production costs will decline, miniaturization will improve, and the scope of applications ranging from everyday water usage to wearable devices will continue to grow. With further exploration into sustainable materials, advanced structural designs, and robust circuit integration, liquid based TENGs are poised to play a central role in future energy solutions, promising both reliability and scalable power generation in practical settings.
Water in nature exists in diverse forms, from bulk liquid (e.g., flowing streams and ocean waves) to droplets and vapor. The electrokinetic mechanisms can harvest energy stored in different forms of water. Compared with liquid-based TENGs, which were discussed in the previous sections, EKEC systems represent a fundamentally different energy conversion mechanism. Although both rely on interactions between liquid and solid surfaces and are enhanced by fluid motion, their working principles diverge significantly. Liquid-based TENGs generate electricity through contact electrification and electrostatic induction. The surface charges are exchanged when a liquid comes into contact with and separates from a dielectric surface, leading to transient current pulses. In contrast, EKEC exploits the EDL formed at charged solid–liquid interfaces. As the liquid moves along the surface, it induces the transport of counterions in the EDL, producing an electrical output. Furthermore, the output from the liquid-based TENG is typically AC due to its periodic contact-based operation, whereas EKEC can produce either AC or DC, depending on the specific flow. A more comprehensive comparison of these two technologies is provided in Table 2. To systematically explain how EKEC works for various forms of liquids, this section categorizes them into three main approaches: drawing potential, waving potential, and streaming potential.
Comparison dimension | REWOD | TENGs | EKEC |
---|---|---|---|
Working principle | Capacitance variation induced by droplet deformation through electrowetting | Electrostatic induction and charge transfer triggered by contact-separation | Flow potential generated by charge dragging induced by fluid shear |
Typical output power density | 10−3 ≃ 100 W m−2 (theoretical max: ∼104 W m−2) | 10−5 – 160 W m−2 | 10−3 ≃ 1 W m−2 |
Energy conversion efficiency | <10% | 0.01–11% | ∼1–4% (theoretical max: 40%) |
Fabrication cost | Highly dependent on high-dielectric-constant metal oxide coatings, complex micro/nanofabrication, sophisticated electrode manufacturing, and challenging fabrication of enclosed droplet structures, resulting in overall system complexity and high fabrication cost | Mostly utilize low-cost, mature fabrication techniques such as electrospinning, spraying, printing, and textile integration | Relies on precise interfacial control and microchannel/nanopore fabrication. Involves diverse materials, lacks standardized fabrication routes, and presents low overall process maturity, leading to high and difficult-to-control fabrication costs and poor scalability |
Materials: ∼$0.5 per cm2 | Materials are easily accessible, processes are simple, suitable for large-area manufacturing with low production cost | Materials: ∼$1 per cm2 | |
Micro/nanofabrication: ∼$10–50 per cm2 | Materials: ∼$0.1 per cm2 | Micro/nanofabrication: ∼$10–50 per cm2 | |
Assembly: ∼$1–5 per cm2 | Micro/nanofabrication: ∼$0.1–1 per cm2 | Assembly: ∼$5–10 per cm2 | |
Total estimated cost: ∼$11.5–55.5 per cm2 | Assembly: ∼$0.2 per cm2 | Total estimated cost: ∼$16–66 per cm2 | |
Total estimated cost: ∼$0.3–1.3 per cm2 | |||
Applicable scenarios | • Pressure-type wearable energy harvesting (e.g., shoe soles, flexible components) | • Contact- and friction-based wearable textile energy harvesting | • Energy harvesting from droplets, waves, fluid pipelines, and surface moisture |
• Pressure-type mechanical motion energy harvesting (e.g., oil pumps, vibration devices) | • Droplet and wave impact energy harvesting | • Self-powered sensors for respiration, humidity variation, and seepage monitoring | |
• Self-powered sensors for pressure and motion monitoring | • Self-powered sensors for biological and chemical composition detection | ||
• Self-powered environmental and rainfall alarm sensors | |||
Advantages | • High energy density harvesting under high-frequency and strong compression scenarios | • Ultra-low cost | • Capable of adapting to complex fluid environments such as waves, humidity, and evaporation. |
• High energy conversion potential | • Simple fabrication process | • Simple structure | |
• Suitable for large-area textile and friction-based scenarios | • Generate both AC and DC | ||
• Expandable to hybrid systems for diversified energy harvesting | |||
Disadvantages | • Complex fabrication process with high cost, difficult to achieve mass production | • Poor long-term durability, prone to wear, contamination, moisture, and corrosion | • Low energy conversion efficiency with insufficient understanding of interfacial charge mechanisms |
• Low power output, highly dependent on complex high-dielectric films and micro/nanofabrication systems | • Poor environmental adaptability, significant reduction in energy output under high temperature and humidity due to electron leakage | • Lack of standardized synthesis and device design processes | |
• Synchronization and control of multiple droplets is challenging, requiring the development of enclosed multi-droplet microstructures | • Biofouling and surface blockage can accumulate on the device, leading to performance degradation |
![]() | (30) |
![]() | (31) |
V = −lRsqψC0v | (32) |
I = q0Wv | (33) |
![]() | (34) |
![]() | (35) |
![]() | (36) |
The EKEC efficiency is an important metric to quantify the performance of streaming potential systems. It is defined as the ratio of the harvested electrical power to the mechanical power input required to drive the flow. The EKEC efficiency ηeff can be mathematically expressed as:60
![]() | (37) |
One direct strategy for increasing zeta potential is to enhance the intrinsic surface charge density. Metal oxides and their derivatives are well known for carrying substantial surface charges in aqueous suspensions.281 Due to their high zeta potential, metal-based nanomaterials (e.g., nanoparticles and nanowires) are highly suitable for EKEC applications. For example, Al2O3 nanoparticles with a zeta potential of 40.3 mV have been incorporated into hydrophilic PET membranes, yielding a sustained voltage of 2.5 V and a current of 0.8 mA.282 The output voltage from other metal oxides materials, such as Fe2O3, Mn3O4, and TiO2, showed a positive correlation with the strength of their zeta potential.282
Furthermore, specific functional groups on solid surfaces, such as hydroxyl (–OH), carbonyl (–CO), and carboxyl (–COOH) groups, ionize in solution, imparting a negative charge to the solid surface and further enhancing the zeta potential,283 thereby improving energy conversion efficiency (Fig. 20a). For example, Li et al.284 functionalized carbon nanoparticles with poly(sodium-p-styrene sulfonate) (PSS), 1,2,3,4-butanetetracarboxylic acid (BTCA), polyethyleneimine (PEI), and poly(diallyl dimethylammonium chloride) (PDADMAC). The corresponding zeta potentials of the modified carbon nanoparticles shifted from the initial −30 mV to −58 mV, −35 mV, 20 mV, and 56 mV, respectively. Consequently, as shown in Fig. 20b, the resulting streaming potential increased from 1.2 V to 3.0 V, 2.8 V, −2.5 V, and −1.5 V, respectively.
![]() | ||
Fig. 20 Strategies to improve electrical output. a) Schematic of modification with different molecules on carbon particles. Reproduced with permission.284 Copyright 2019, Elsevier. b) Streaming potential generated on different PSS, BCTA, PEI, PDADMAC modified surfaces with different zeta potential levels. Reproduced with permission.284 Copyright 2019, Elsevier. c) The waving potential generated across electrolytes with different ion radii. Reproduced with permission.274 Copyright 2014, Springer Nature. d) Schematic of velocity field near the interface with no-slip boundary condition (left) and Naiver-slip boundary condition (right). Reproduced with permission.327 Copyright 2021, Elsevier. e) Increasing the slip length from 0 nm to 50 nm could theoretically boost the maximum conversion efficiency. Reproduced with permission.300 Copyright 2008, Wiley. f) The SLIPS can overperform the generated streaming potential compared to flat and superhydrophobic surfaces with the same pattern. Reproduced with permission.310 Copyright 2018, Springer Nature. g) The overlapped EDLs can significantly enhance the counterion concentration at the center of the channel. Reproduced with permission.328 Copyright 2012, Springer Nature. h) As the ratio between the channel height and Debye length increased, the dimensionless potential significantly reduced. Reproduced with permission.322 Copyright 2010, American Institute of Physics. i) Variation of dimensionless streaming potential influenced by the synergistic effect of dimensionless EDL thickness and ionic Peclet number. Reproduced with permission.323 Copyright 2013, Elsevier. |
Besides optimizing solid material properties, the ionic properties (e.g., ion species and ion concentration) of the electrolyte significantly influence the streaming potential/current in EKEC. Ion species also play a crucial role in electrokinetic phenomena,285 influencing ion adsorption, hydration, and mobility at the solid–liquid or liquid–liquid interfaces. Zhang et al.286 experimentally measured the streaming potential of LiCl, NaCl, and KCl solutions flowing through ion exchange membranes with the same ion concentration to determine the influence of ion species contributing to zeta potential. The zeta potential decreased in the order of Li+ > Na+ > K+, which is attributed to variations in ionic mobility.287 Also, Yin et al.274 experimentally measured the drawing and waving potential generated by LiCl, NaCl, KCl, NaF, NaCl, NaBr, and NaI solutions. As shown in Fig. 20c, with a consistent 0.6 M concentration, the higher output voltage was generated with a smaller ionic radius for both cations (Li+ > Na+ > K+) and anions (F− > Cl− >Br−). In addition, the higher valency of ions also enhances the zeta potential. Löbbus et al.288 experimentally investigated the influence of mono-, di-, and trivalent counterions on zeta potential. The results indicated that electrolytes with trivalent counterions exhibit much larger zeta potential than electrolytes with di- and monovalent counterions.
In addition to ion size, higher electrolyte concentration can significantly reduce electrokinetic energy conversion efficiency. With higher ion concentrations in the electrolyte, a higher fraction of the excess charge ions is in the Stern layer,289 resulting in the reduction of the thickness of the EDL and zeta potential on the solid surface.290 By fitting various experimental results,291–294 the relationship between ζ and ion concentration, c, can be approximated as: ζ ≃ logc for the low ion concentration range (<1 M).280,295 Kirby et al.280 concluded that the zeta potential yields a linear relationship with pC (−log
c = pC) while different aqueous solutions are in contact with the solid substrate.
To quantify the effect of slip on electrokinetic energy conversion, the Stokes equation can be solved under the Navier boundary condition. This yields a velocity profile: , where H is the channel height. By substituting the velocity and local charge density profile in the streaming current equation, Is = ∫H0ρe(y)u(y)dy, the streaming current under Navier slip boundary condition yields:
.298 For no-slip boundary condition, the streaming current results in:
.280 Then, the streaming current enhancement factor is:
. This results in an effective zeta potential ζe at the slipping interface, which is higher than the intrinsic zeta potential at the no-slip surface ζ, with
.299 Ren et al.300 theoretically analyzed the effect of varying slip lengths on energy conversion efficiency in nanofluidic channels. As shown in Fig. 20e, for a 100 nm-high channel, increasing the slip length to 50 nm could theoretically boost the maximum conversion efficiency to approximately 40%.
Recent studies also suggest that surface charge distribution influences slip length. Xie et al.301 found that strong coulombic interactions between counter-ions and the non-uniformly charged surface increase friction, thereby reducing the slip length. In contrast, materials such as graphene and carbon nanotubes exhibit more uniform surface charge distributions, which contribute to enhanced slip.
A straightforward strategy to enhance slip length is utilizing slippery surfaces to reduce flow friction and improve energy conversion efficiency.302 Inspired by lotus leaves, superhydrophobic surfaces can extend slip lengths to the submicron or even micron scale.303 The superhydrophobic surfaces are characterized by a very high contact angle (>150°).304 They are normally created by combining micro/nano scale textures with a low surface energy material.305 The trapped air within these textures minimizes the moving friction between the liquid and solid surface, allowing liquid to flow more freely over the surface with higher slip length. Malekidelarestaqi et al.306 simulated the EKEC efficiency on superhydrophobic surfaces with varying slip lengths. Their results showed that increasing the slip length from 0 to 144 nm could improve the efficiency by a factor of 3.4. However, while superhydrophobic surfaces can increase slip length through trapped air pockets, the limited surface charge at the liquid–air interface negatively impacts EKEC efficiency. Squires307 utilized the Lorentz reciprocal theorem for the Stokes flow and proposed that if the slip region is uncharged, the effective electrokinetic flow is not enhanced compared to the charged non-slip region. In addition, Zhao et al.308 demonstrated that at moderate to high zeta potentials, electrokinetic flow over a charge-free liquid–gas interface generates lower streaming currents than that over a uniformly charged no-slip surface.
To overcome this limitation, SLIPS have been introduced. SLIPS combine micro/nano scaled textured materials with a lubricating liquid layer infused into those textures.309 The liquid layer reduces friction and enhances fluid slip by preventing direct contact between the liquid and the solid surface. In the meanwhile, the liquid–liquid interaction provided higher surface charge density than the liquid–air interface in superhydrophobic surfaces.310 These advantages make SLIPS a promising candidate for electrokinetic applications. Fan et al.310–315 demonstrated that the SLIPS achieved a higher figure of merit than superhydrophobic surfaces, using NaCl solution on SLIPS infused with a low-dielectric-constant lubricant oil. As shown in Fig. 20f, they achieved a figure of merit enhancement of 0.043 mV Pa−1, outperforming superhydrophobic surfaces with similar surface patterns.
Notably, theoretical and experimental studies have demonstrated that surface charge and boundary slip are intrinsically coupled rather than independent phenomena. Specifically, an increase in surface charge density corresponds to a decrease in slip length.297,316,317 This dependence of slip length on surface charge inevitably influences fluid behavior at the micro- and nanoscale. Therefore, it is crucial to account for this interdependence while designing the substrate.
These findings demonstrate that increasing interfacial slip is a useful strategy for enhancing EKEC efficiency. By reducing frictional losses and promoting ion mobility, approaches such as superhydrophobic surfaces and SLIPS significantly improve charge transport and power generation. The integration of these surface engineering strategies offers a promising route toward higher electrokinetic performance and more efficient energy harvesting.
Osterle et al.320 were among the first to employ thermodynamic analysis to couple fluid flow with electrical current. They predicted a maximum EKEC of approximately 0.9% in a 100 nm glass capillary. Their group further simulated that the efficiency can be improved to ∼4% by reducing the characteristic length of the channel until the EDLs overlap (less than twice of λD).321 Ban et al.322 developed a simple model for water flow without salts or dissolved gases to analyze the effects of overlapped EDLs on electro conductivity and potential in the channel. As shown in Fig. 20h, The potential distribution reveals a higher potential near the channel walls and a lower potential at the channel center. When the channel height is equal to or smaller than the Debye length (kh = 0.5, where k = 1/λD), the dimensionless potential at the channel center increases significantly. With advancements in microfabrication techniques and materials science, experimental verification of EKEC-based energy harvesting has become feasible. Das et al.323 theoretically investigated the synergistic effect of dimensionless EDL thickness and ionic Peclet number (
, where Uc is the characteristic speed, Lc is the characteristic length, and α is the thermal diffusivity) on streaming potential. As shown in Fig. 20i, the dimensionless streaming potential rose rapidly with higher
until the EDLs overlapped. After the EDLs overlapped, or with large R (R > 1), the dimensionless streaming potential varies weakly with
. For R ≪ 1, the dimensionless streaming potential continuously rises with higher
.
Although overlapping the EDLs improves electrokinetic electric output by enabling unipolar ion transport, it may also significantly increase the viscous resistance and pressure drop required for driving liquid motion in the channel.324 According to the Hagen–Poiseuille equation,325 , the pressure drop increases sharply as the channel radius decreases. The hydraulic resistance, defined as
, scales inversely with the fourth power of the channel radius. For a cylindrical nanochannel with a radius of 100 nm and a length of 1 mm, maintaining a flow rate of 1 nL min−1 requires a pressure drop of ∼108 Pa. Furthermore, the wettability of the channel wall plays a critical role by influencing both the direction and magnitude of the capillary pressure. The capillary pressure can be expressed with the Young–Laplace equation:326
, where γ is the surface tension, r is the radius of the nanochannel, and θ is the contact angle. For hydrophilic channels (θ < 90°), the capillary pressure can serve as a passive driving force to assist fluid infiltration. The maximum assisting pressure can reach up to ∼106 Pa. For hydrophobic surfaces (θ > 90°), the cos
θ becomes negative, effectively preventing spontaneous liquid infiltration, and increasing the difficulty of liquid flow through the nanochannel. Therefore, the maximum critical threshold pressure for initiating flow in hydrophobic channels can reach up to the order of ∼106 Pa. As a result, the nanochannel with overlapping EDLs demands enormous driving forces, making continuous flow energetically expensive and practically infeasible without external pumping. This comparison emphasizes the severe hydrodynamic constraints in overlapped-EDL systems, where electrical output gains must be carefully weighed against mechanical feasibility.
![]() | ||
Fig. 21 Application of a droplet energy generator. a) (i) The droplet slides on the graphene sheet drawn by Si wafer, generating a drag potential. a) (ii) Drawing a droplet from the left end of the graphene strip to the right end and back again generates a voltage of 0.15 mV. The reversed droplet velocity will generate a negative output voltage. a) (iii) The voltage induced when moving one, two, and three droplets. The output voltage is proportional to the number of droplets. Reproduced with permission.274 Copyright 2014, Springer Nature. b) (i) The voltage generated by a 0.6 M NaCl solution when it moves in a graphene microchannel. The voltage direction varies with the flow direction of the droplet. b) (ii and iii) The output voltage of the graphene microchannel device changes at different channel widths and heights. Reproduced with permission.334 Copyright 2022, American Chemical Society. c) (i) The droplet energy generator consisted of graphene/polymer film on a SiO2 plate. c) (ii) The graphene/PET film can generate 100 mV pulse voltage, while it is undetectable on graphene/PMMA. Reproduced with permission.336 Copyright 2018, American Chemical Society. d) (i) Mechanisms of electricity generation while the droplet slid on MoS2 film. d) (ii) Voltage generated while MoS2 film is deposited on different substrates. Reproduced with permission.339 Copyright 2020, Elsevier. |
Following this breakthrough, advancements in materials science have led to the development of more efficient droplet energy harvesters. Zhao et al.333 designed a flexible single-electrode raindrop energy harvester using a graphene/reduced graphene oxide (G/rGO) composite film. They recorded periodic current and voltage signals under continuous raindrop stimulation to evaluate the efficiency of converting rainwater into electricity. This device leveraged the pseudocapacitive charging/discharging cycle at the G/rGO–droplet interface, generating stable electrical signals for over 1000 s. By optimizing the graphene ratio, the harvester achieved a maximum output current in the sub-mA order per droplet and a voltage in the hundreds of mV order per droplet. However, prolonged exposure to rain led to ion-induced electron trapping, which degraded the performance.
Kong et al.334 explored the potential of graphene-based droplet energy harvesting devices by optimizing the device architecture. They reported a graphene microchannel structure that significantly enhanced energy conversion efficiency. As shown in Fig. 21b(i), when a 50 μL, 0.6 M NaCl droplet flows through the graphene microchannel at a velocity of 20.55 cm s−1 from left to right, an output voltage of 150 mV is generated. When the droplet moves in the opposite direction at the same velocity, an equal but opposite voltage is induced. The output performance far surpasses that of graphene in an open planar configuration. To further investigate the effect of channel dimensions on energy harvesting performance, they systematically varied the channel width. As illustrated in Fig. 21b(ii), when the channel width increases from 150 μm to 350 μm, the expanded liquid–solid interface enhances electron extraction from the graphene surface, thereby increasing the induced voltage. However, in channels wider than 400 μm, the droplet height cannot reach the top of the channel during the flow, leading to certain loss in flow velocity. Therefore, the response of output voltage is reduced in a wider channel. Since the droplet motion is driven by a pressure gradient, the channel length also plays a critical role in energy output. As shown in Fig. 21b(iii), when the channel length increases from 2 cm to 5 cm, the output voltage rises from 2.1 mV to 147 mV. However, further elongation of the channel length from 6 cm to 8 cm results in a gradual decline in voltage due to a reduction in flow velocity. Based on these findings, the optimal channel parameters were determined to be 350 μm in width and 5 cm in length, providing a well-balanced design for maximum droplet-induced energy harvesting.
Furthermore, the drag-induced potential effect has been observed at the water interface of polymer-coated insulator–semiconductor structures.335 Yang et al.336 investigated the influence of polymer polarity on droplet energy harvesting by integrating different polymer/graphene layers onto a SiO2 substrate (Fig. 21c(i)). To simulate rainfall conditions, 5 mm-diameter, 0.6 M NaCl droplets continuously fell on a 60°-inclined surface at an initial velocity of ∼1 m s−1. As shown in Fig. 21c(ii), graphene/PET films, which exhibit stronger Na+ adsorption, produced a pulsed voltage of ∼100 mV, whereas no detectable output was observed on graphene/PMMA surfaces. The results indicate that the polymer substrate plays a dominant role in attracting ions to the water/graphene interface, with the PET substrate exhibiting a significantly stronger attraction to Na+ ions compared to the PMMA substrate.
Recent studies have identified molybdenum disulfide (MoS2) as a promising alternative to graphene due to its higher electrical resistance, enabling greater voltage generation.337,338 Aji et al.339 developed a single-layer MoS2-based droplet energy harvester (Fig. 21d(i)), demonstrating a pulsed voltage of 6 V and a current of 5 nA when a 50 μL, 1 M NaCl droplet was deposited on a 45°-tilted surface, which outperforms all previously reported droplet-based energy harvesters. They further investigated the impact of different substrates (polyethylene naphthalene (PEN), sapphire, and SiO2/Si) on power generation. The PEN substrate, characterized by reduced n-type doping,340,341 increased the MoS2 film's resistance, resulting in a higher output voltage (Fig. 21d(ii)). Similarly, Kumar et al. replaced the substrate with Si3N4/Si and achieved an output of 7.3 V and 11 nA when a 60 μL, 0.6 M NaCl droplet dropped from a height of 4 cm and slid across the MoS2 surface.
These advancements highlight the potential of droplet-based energy harvesting as a viable and scalable approach for sustainable energy conversion. By leveraging novel material architectures and optimizing interfacial properties, researchers continue to push the limits of energy output, opening new possibilities for self-powered systems and environmental energy harvesting applications.
![]() | ||
Fig. 22 Application of a waving energy generator. a) (i) The graphene is placed on a PET substrate and moved vertically on the water surface in a container to generate waving potential. a) (ii) The voltage signals generated when the sample is inserted and pulled out of a 0.6 M NaCl solution at a speed of v = 3.1 cm s−1. a) (iii) The relationship between the magnitude of the positive peak voltage and the negative peak voltage and the speed at which graphene moves across the surface of a 0.6 M NaCl solution. Reproduced with permission.275 Copyright 2014, Springer Nature. b) (i) The schematic of wave energy generation stations in the ocean. b) (ii) Optimized output performance with 80 wt% G-CB films at different seawater temperatures. Reproduced with permission.344 Copyright 2018, Elsevier. c) (i) The schematic of two graphene system waving energy generators. c) (ii) The voltage generated while separately moving each graphene across NaCl solution. c) (iii) The output power density of graphene with different external resistance while inserting and pulling across NaCl solution. Reproduced with permission.345 Copyright 2019, Elsevier. |
Inspired by the advancement of waving potential, Tan et al.344 developed a thin-film wave energy generator by mixing graphene and carbon black with aqueous polyurethane, then casting the mixture onto substrates such as glass, plastic, and ceramics (Fig. 22b(i)). By optimizing parameters such as the carbon dose, wave frequency, tilt angle, and seawater temperature, the generator, with a size of 15 cm2, can produce voltages greater than 20 mV and currents greater than 10 μA (Fig. 22b(ii)). Under continuous wave impacts, these thin-film generators can maintain stable output for more than five hours. This method of harvesting wave energy offers new opportunities for low-investment, scalable power output with high stability.
Fei et al.345 improved the wave energy harvester by using a dual-graphene-sheet system. Unlike the fluctuation potential observed in single graphene sheets, the voltage in the dual-graphene system is generated within the solution. In such a system, as shown in Fig. 22c(i), the moving graphene sheet serves as the driving force for ion motion, while the stationary graphene acts as a reference electrode. Due to the low carrier density and high ion adsorption/desorption capacity of the graphene sheets, when a certain number of ions are adsorbed onto the graphene, there are insufficient electrons to shield the ion charges. As a result, a noticeable current can be induced in the external circuit outside the solution. As shown in Fig. 22c(ii), when a single graphene sheet (GrR) is inserted into the liquid surface, the voltage across the resistor gradually increases and peaks at 60 mV. After pulling GrR from the solution, the voltage drops significantly and reaches −120 mV before the sheet completely stops near the liquid surface. Shortly thereafter, the voltage returns to zero as the system reaches equilibrium. They also calculated the output power per unit area of graphene, which shows a nonlinear relationship with the external resistance. As shown in Fig. 22c(iii), during the insertion of graphene, the maximum output power occurs when the external resistance is 0.1 MΩ. During the withdrawal of graphene from the liquid surface, the maximum output power is observed at 0.6 MΩ, reaching approximately 1.6 mW m−2. They simulated that simultaneous movement of the graphene sheets in opposite directions could further enhance instantaneous power. This method of collecting water energy provides new insights for the development of wave energy harvesters.
The ongoing advancement of materials science, particularly the exploration of novel composites and coatings, will likely lead to even more efficient, cost-effective, and environmentally friendly wave energy harvesting technologies.
Zhao et al.283 reported a graphene oxide (GO) film with a carboxyl gradient capable of adsorbing and condensing moisture within its nanoscale pores. The oxygen-to-carbon (O/C) atomic ratio was approximately 0.52 at the bottom and 0.22 at the top of the film. Due to the strong hydrophilicity of the carboxyl group, the device absorbs water molecules upon exposure to moisture. As the water molecules accumulate in the oxygen-rich regions of GO film, localized solvation effects weaken the O–H bonds within the carboxyl group, leading to the release of free H+ ions. In the GO film, the carboxyl group exhibits a gradient distribution, resulting in a corresponding H+ concentration gradient. This gradient drives the diffusion of H+ ions generating a streaming potential induced by moisture flow. The device, comprising a 2.8 mm-thick GO film clamped between two asymmetrically sized gold electrodes, produced a pulse voltage of ∼20 mV and a current density of 5 μA cm−2 at ∼30% relative humidity (RH).
To expand the application of moisture energy generators in portable electronics, Cheng et al.354 developed a flexible in-plane moisture-electric converter (IPMEC) based on GO assembled films. The IPMEC device features two laser-rGO planar electrodes integrated with GO layers in between, exhibiting an oxygen-containing functional group gradient, ensuring full exposure to moisture sources. Water molecules in humid environments interact with GO and induce an ion concentration gradient along the in-plane direction of the GO layers (Fig. 23a), resulting in a substantial electrical output (∼70 mV, 12 mA cm−2) in an external circuit.
![]() | ||
Fig. 23 Application of a moisture energy generator. a) Schematic of the moisture-electric conversion cycle. The arrows represent induced voltage and current in the presence and absence of moisture. Reproduced with permission.354 Copyright 2018, Elsevier. b) (i) Schematic of the TiO2 based moisture energy generator generating electricity while harvesting moisture in the environment. b) (ii) Voltage generated in response to the 85% RH level. Reproduced with permission.355 Copyright 2018, Wiley. c) Schematic of the asymmetric sandwich structure of the moisture energy generator. The feed layer with additional GO supplements the depletion of ions in the gradient rGO, and the special electrode/h-GO interface blocks the free electrons from passing through h-GO and regulates their flow direction, which greatly promotes the electrical output. Reproduced with permission.356 Copyright 2019, Royal Society of Chemistry. d) (i) Schematic of PMEG, which consists of a PSSA membrane between two gold electrodes. The upper electrode is manufactured with holes for efficient moisture access. d) (ii) the generated output voltage on a single 1 × 1 cm2 PSSA membrane under constant moisture. d) (iii) A mist powering system built from 6 PMEG units connected in series. Reproduced with permission.357 Copyright 2019, Royal Society of Chemistry. |
Shen et al.355 further enhanced the performance of moisture energy generators by constructing a hydrophilic TiO2 nanowire network, which forms a 3D nanoconfined channel architecture to facilitate moisture absorption and diffusion (Fig. 23b(i)). Under alternating high and low humidity conditions, the device generated a pulse output of 0.6 V and 4 μW cm−2 (Fig. 23b(ii)), outperforming previously reported carbon-based moisture energy generators by two orders of magnitude. Notably, moisture from human breath provides sufficient energy to power a commercial LED. The device also exhibited intrinsic flexibility, maintaining stable output characteristics even after 10000 bending cycles.
The performance of moisture energy generators is significantly influenced by both ambient relative humidity and the hygroscopic properties of the materials. Insufficient hygroscopicity at low humidity may lead to inadequate ion dissociation, whereas excessive hygroscopicity at high humidity results in rapid saturation, limiting device longevity. To extend the sustainability of energy conversion, one promising strategy involves the design of nanostructured architectures to maintain the moisture difference in the device with heterogeneous chemical structure. For instance, Huang et al.356 proposed a sandwich-structured moisture energy generator composed of Ag/pristine GO/gradient-reduced GO/Au (Fig. 23c). The bottom GO layer offers a rich supply of mobile electrons, while the Schottky junction at the GO/Ag interface rectifies ion flow and prevents electron backflow. This device effectively harvested energy from atmospheric moisture over a wide temperature (25.8 °C–50.8 °C) and RH (5–95%) range, maintaining a stable open-circuit voltage of 0.6 V for over 120 hours. By establishing a self-sustained humidity gradient, this strategy significantly reduces limitations associated with environmental conditions.
Xu et al.357 introduced a novel approach using a flexible polyelectrolyte membrane, poly (4-styrenesulfonic acid) (PSSA), to achieve efficient moisture-induced power generation (Fig. 23d(i)). Their polymer-based moisture energy generator, consisting of a 1 cm2 PSSA membrane, delivered an open-circuit voltage of ∼0.8 V (Fig. 23d(ii)) and a short-circuit current density of ∼0.1 mA cm−2, sufficient to power various electronic devices. By integrating two polymer-based moisture energy generators in series, they designed a wearable mask capable of harvesting energy from human respiration. During normal breathing, the device could charge a capacitor to 0.9 V in 30 minutes. Additionally, both boiling water vapor and mist from an ultrasonic humidifier effectively contributed to energy harvesting. As illustrated in Fig. 23d(iii), six PMEG units were connected in series to power an LED. When placed near a commercial ultrasonic humidifier, the generated mist gradually reached the PMEGs, successfully illuminating the LED. The remarkable flexibility and superior electrical output of PMEGs highlight their tremendous potential in wearable energy-harvesting systems. These groundbreaking studies highlight the immense potential of MEG technologies, paving the way for future research and further advancements in this field.
![]() | ||
Fig. 24 Applications of an evaporation energy generator. a) Schematic of the setup for harvesting energy from water evaporation through CB film. Reproduced with permission.353 Copyright 2017, Springer Nature. b) (i) Schematic of the electricity generation induced by evaporation in a piece of wood. The microchannels of the wood piece are vertically aligned. When water flows in the microchannels, the inner surfaces of the channel are negatively charged, resulting in a positively charged flow. b) (ii and iii) Output voltage and current of different wood species and wood channel direction. Reproduced with permission.364 Copyright 2020, American Chemical Society. c) (i) Schematic of the solar-driven evaporator using a bilayer carbon nanotube/cellulose paper structure. c) (ii) Under solar irradiation, the electrical output was enhanced compared to the dark environment. Reproduced with permission.372 Copyright 2020, Elsevier. d) Schematic of the self-powered wearable wristband. While DI water evaporated, driven by body heat, the porous carbon membrane harvested the energy from evaporation to electricity. Reproduced with permission.374 Copyright 2018, Wiley. |
Inspired by the high porosity of CB, Zhang et al.360 explored the use of three-dimensional GO to construct evaporation-based energy harvesters. Their device, designed as a 3D GO sponge cup, achieved an output voltage of 0.63 V and a power density of 1.74 mW cm−2. Subsequent studies employed chemical etching to disperse GO, increasing the concentration of carboxyl groups and nanopores on GO film, thereby further enhancing device performance. These modifications improved mass transport and evaporation rates, significantly boosting overall efficiency.
Metal oxides are considered to be charged in aqueous suspension due to amphoteric dissociation of surface hydroxyl groups or adsorption of hydroxo complexes from hydrolysis products.281 They can achieve relatively high surface zeta potentials, making them highly suitable for evaporation energy harvesting, particularly in the form of nanoparticles and nanowires. Shao et al.282 utilized Al2O3 nanoparticle coated hydrophilic PET films to collect evaporation energy. Their device produced a continuous voltage of 2.5 V and a current of 0.8 mA, with sustained operation for up to 10 days. Li et al.361 fabricated a ZnO nanowire array with a 180 nm pitch, encapsulated in poly (vinyl chloride-co-vinyl-co-2-hydroxypropyl acrylate) (PVC). By inducing ethanol droplet evaporation on the ZnO surface, the device generated a voltage of 133 mV with a duration of 150 seconds. This study demonstrated that non-random distributed nanochannels can also effectively harvest energy from evaporation, highlighting the potential of nanowire structures for energy harvesting applications.
Biomass materials with intrinsic nanostructured channels, such as cellulose and lignin, offer excellent biocompatibility and degradability, making them promising candidates for environmentally friendly energy harvesting.362,363 Zhou et al.364 leveraged the abundant micro-scale channels in wood to harvest energy from evaporation (Fig. 24b(i)). By optimizing surface potential and hydrophilicity through citric acid esterification of hydroxyl groups, the wood-based device generated a continuous voltage of 300 mV and a current of 10 mA. They investigated four different wood types for evaporation energy generation, evaluating the influence of wood species and microchannel orientation on electrical output. As shown in Fig. 24b(ii and iii), measurements of a 5 cm × 5 cm × 1 cm wood device revealed that beech wood exhibited the highest performance, generating currents above 2 μA and voltages exceeding 30 mV. This performance variation was attributed to differences in hydrodynamic resistance due to their different pore dimensions among wood materials (ranging from 5–40 μm). Additionally, comparing devices with identical wood types, vertically aligned channels demonstrated significantly higher current and voltage outputs than horizontally aligned channels, because the former are aligned with the water evaporation direction. Similarly, Jiao et al.365 developed a carbonized freeze-dried carrot slice device, achieving an open-circuit voltage of 0.8 V and a short-circuit current in tens of milliamps per device.
Su et al.366 adopted an innovative approach by using 160 plasma-treated wool strips coated with Ketjen black powder to absorb and evaporate seawater, successfully charging a supercapacitor to 1.6 V within 5.5 hours. However, continuous seawater evaporation led to salt crystallization on the evaporation surface, reducing the evaporation rate with time.367–369 To address this issue, Peng et al.370 designed an asymmetric fluid evaporator enabling gravity-assisted salt collection and permeation-induced electrokinetic generation. The evaporator was fabricated by spray coating tannic acid modified MoS2 nanosheets on partial area of metal–organic framework functionalized polyacrylonitrile textiles (PAN) to create an asymmetrical structure. While drenching with water, the asymmetrically deposited textile induced varied ionic adsorption behavior at different sites, resulting in formation of a gradient EDL. The salt always tends to crystallize at one end of textile and falls off under the drive of gravity, improving the sustainability of the evaporation efficiency and electrical harvesting. Their experimental results demonstrate that after 60 hours of continuous operation in a 7.5 wt% saline solution, the device maintains a stable evaporation rate of 1.31 kg m−2 h−1 and a voltage output of 0.514 V.
For evaporation energy devices, elevated temperatures can accelerate liquid evaporation, thereby significantly enhancing power output.371 Thus, integrating photothermal materials can further improve water evaporation and energy generation efficiency. Xiao et al.372 developed a solar-driven evaporator using a bilayer carbon nanotube/cellulose paper structure, asymmetrically modified with PDMS on half area of the carbon nanotube substrate (Fig. 24c(i)). When the hydrophobic PDMS coating is applied to the surface of carbon nanotubes (CNTs)/paper, it significantly alters the light path, enhancing the visibility of the black CNTs film on the reverse side. As a result, the paper's appearance changes from white to black. Under varying illumination intensities, as depicted in Fig. 24c(ii), the CNT/paper device exhibited an increase in voltage and current from 0.47 V and 15 μA in darkness to 0.6 V and 22 μA under 500 W of solar irradiation.
Harvesting energy from human body heat presents challenges due to the small temperature gradients involved.373 Liu et al.374 utilized a porous carbon membrane and DI water (Fig. 24d) to generate electricity from body heat for directly powering a wearable wristband. Driven by evaporation under a low thermal gradient, this device highlights the potential for low-temperature thermal energy harvesting and self-sustaining wearable sensor systems. Their device achieved an open-circuit voltage of 0.89 V under a temperature difference of only 4.2 °C, sufficient to directly power electronic devices.
These advancements underscore the immense potential of evaporation-driven electrokinetic generators, offering promising pathways toward scalable, efficient, and multifunctional energy harvesting solutions.
Human daily activities involve many mechanical motions, providing an opportunity for energy harvesting. Li et al.375 demonstrated a flexible microfluidic nanogenerator (MFNG) that utilizes streaming potential to convert ambient mechanical energy into electricity (Fig. 25a(i)). The MFNG consists of ten microchannels, each with a width of 100 μm. When a force is applied to one end of the reservoir, the electrolyte flows through the microchannel under pressure, generating a streaming potential. When the force is removed, the electrolyte backflows, producing an opposite streaming potential. They investigated the relationship between the generated voltage/current and the applied force. As shown in Fig. 25a(ii), both voltage and current increase almost linearly with the applied force, suggesting that using larger force sources (e.g., footsteps) could further enhance the output. A simple finger press can generate a pulse voltage of ∼1.5 V and a current of ∼1 μA across a total effective cross-sectional area of 1 mm2 in the microchannel, which is sufficient to power a commercial liquid crystal display (LCD) (Fig. 25a(iii)). The output performance of MFNGs can be further optimized by adjusting parameters such as fluid type, solution concentration, microchannel surface properties, and geometric dimensions. This approach presents a promising strategy for developing power supply systems that harness mechanical energy from the human body.
![]() | ||
Fig. 25 Applications of a microfluidic energy generator. a) (i) Schematic of a flexible MFNG which consists of ten microchannels. The width of each channel is 100 μm. a) (ii) The voltage generated by the MFNG at different external forces. a) (iii) Photo of the LCD powered by the MFNG directly. Reproduced with permission.375 Copyright 2016, Elsevier. b) (i) Schematic of the microfluidic nanogenerator with a porous patterned structure in the microchannel, where the microfluidic nanogenerator supplied energy to a single NW based pH sensor. b) (ii) Voltage generated from the microfluidic when a 1 μM KCl solution flows through the device at 0.4 mL min−1. Reproduced with permission.31 Copyright 2015, Wiley. |
Additionally, microfluidic systems are widely employed in biomedical diagnostics. The streaming potential effect can be utilized to convert the mechanical energy of fluid flow into electricity, powering other microfluidic devices. Zhang et al.31 fabricated a microfluidic energy generator using soft lithography. They incorporated a patterned micropillar array in the channel, forming a quasi-porous structure, to increase the contact area between the liquid solution and channel walls. Compared to microchannels without internal structures, this design significantly enhanced power output by increasing the liquid–solid interfacial area. Furthermore, they developed a self-powered fluidic sensing system, where the microfluidic energy generator supplied energy to a single nanowire (NW)-based pH sensor (Fig. 25b(i)). As illustrated in Fig. 25b(ii), when a 1 μM KCl solution flows through the device at 0.4 mL min−1, it generates a continuous voltage of ∼160 mV and a current of 1.75 nA.
The output electrical signal of EKEC exhibits high sensitivity to fluid motion, making them ideal for monitoring dynamic fluid properties. Newaz et al.377 developed a graphene field-effect transistor (GraFET) to detect ionic concentration and flow rate in liquids. They fabricated microchannels with dimensions of 80 μm in height and 50 μm in width using soft lithography on PDMS and the transistor was sealed within the microchannel as a probe. When fluid flowed through the channel, the electrical signal of the GraFET increased linearly with the local flow velocity. Additionally, at a constant flow rate, the output electrical signal exhibited an exponential decay with increasing ion concentration. This graphene-based device demonstrated the capability to detect flow rates as low as 70 nL min−1 and ion concentrations down to 40 nM.
Similarly, the drawing potential generated by droplets moving across graphene surfaces exhibits a strong linear correlation with droplet moving velocity and direction, making it a promising candidate for droplet motion sensing. Yin et al.378 further leveraged this phenomenon by designing a system that detects brush strokes using two pairs of orthogonally placed electrodes. As illustrated in Fig. 26a(i), a square graphene sheet was integrated with two electrode pairs (E1+/E1− and E2+/E2−) positioned perpendicularly on four edges. When a brush was immersed with a 0.01 M NaCl solution and wrote on the graphene sheet, the voltage V1 between E1+/E1− responded positively to horizontal strokes from left to right. And the voltage V2 between E2+/E2− responded positively to vertical strokes from top to bottom. By analyzing the voltage signals detected between the electrodes, fundamental brushstroke patterns could be easily recognized. (Fig. 26a(ii)). Furthermore, more detailed information from electrical signals, such as brush tip sharpness, writing speed, and applied pressure, remains an area for further exploration.
![]() | ||
Fig. 26 Applications for self-powered devices. a) (i) Photo of handwriting with a brush on graphene between four electrodes: E1+, E1−, E2+ and E2−. a) (ii) The stroke directions by the drawing potential between electrodes can be sensed based on the generated voltage signal. Reproduced with permission.274 Copyright 2014, Springer Nature. b) Schematic of the transpiration energy power generator for a water seepage early warning system. Reproduced with permission.382 Copyright 2023, American Chemical Society. c) (i) The units “0” and “1” can be integrated into the textile by random permutation and combination. Four information units are integrated into a piece of textile. c) (ii) Schematic of an array of information storage units. c) (iii) Based on different voltage signals, the units are showing the logo of “BIT”. Reproduced with permission.383 Copyright 2017, Elsevier. d) Generated streaming current with different objects and the changed position of fingers. Reproduced with permission.354 Copyright 2018, Elsevier. |
Beyond directly sensing fluid velocity, energy sources that induce velocity changes can also be detected. Kong et al.334 developed a flexible self-powered pressure sensor based on graphene/PDMS microchannels, utilizing the principle that energy conversion at the liquid–solid interface correlates positively with pressure. As pressure increased, deformation and displacement of droplets within microchannel were enhanced, leading to greater motion speed and higher voltage generation. Due to the flexible mechanical properties of graphene/PDMS films and the low Young's modulus of liquid droplets,27 the sensor remained highly responsive to external stimuli even under low pressures. They affixed the flexible device to a human wrist to monitor body posture. When the extended wrist transitioned into a bent state, the droplet deformation and displacement within the flexible device induced a rapid voltage response. Their testing demonstrated a response and recovery time of 36.4 ms and 239.2 ms, respectively, under a 2.138 kPa load/unload cycle. Furthermore, under a 1.895 kPa pressure cycle, the sensor maintained a stable signal over 1000 cycles, further affirming its potential for wearable applications.
The sensitivity of self-powered electrical devices to temperature and humidity enables their application in environmental sensing. Water infiltration-induced geological disasters often result in severe economic loss, casualties, and environmental damage.379–381 To address this issue, a seepage early warning system based on EKEC was proposed (Fig. 26b).382 This system employed MoS2-functionalized filter paper as the sensing material, capable of sensing water seepage as low as 2 μL. When the filter paper was exposed to water, spontaneous counterion adsorption occurred in the wet region, forming an EDL at the electrolyte–solid interface. This process generated a potential difference between the wet and dry regions of the paper. Under the influence of a humidity gradient, water gradually migrated from the wet region to the dry region, carrying adsorbed ions and generating an electrical current. As evaporation progressed and moisture decreased, the output voltage and current correspondingly declined until disappearing completely upon full evaporation. Due to the strong hydrophilicity of filter paper, the system achieved a detection resolution on the order of seconds. When a droplet was introduced at the negative electrode of the filter paper, the voltage rapidly increased from 0 mV to 110 mV within 10 s. Additionally, the system demonstrated stable multi-cycle alarm functionality. After 10 consecutive wet–dry cycles, the current decreased only from 18.3 μA to 14.7 μA, showing that 80% of current was retained.
Besides environmental monitoring, self-powered electrical technology can also be used to detect the humidity emitted by the human body to promote the advancement of human health.384 Respiration and perspiration can serve as indicators of physiological conditions and can be monitored through periodic humidity variation. Zhao et al.283 developed a breath sensor utilizing GO films to monitor breathing status. While humans are under resting conditions, with an average breathing rate of 13 times per minute (RH ∼21%), the sensor generated an output voltage of ∼18 mV. After one hour of jogging, the breathing rate increased to 28 times per minute (RH ∼48%), and the sensor's voltage output rose to 30 mV. The frequency and amplitude of the pulsed voltage signals directly reflected breathing activity. With a similar framework, Li et al.385 replaced GO with bio-compatible cellulose nanofibers and fabricated a smart mask capable of monitoring human breathing activities. The voltage pulses and amplitudes generated by the sensor closely correlated with the respiratory rate and intensity. Additionally, this bio-nanofiber-based sensor exhibited high biocompatibility with human skin cells and could fully degrade in natural soil within 35 days, demonstrating its potential as a next-generation respiration sensor.
Due to sweat and moisture evaporation, the air surrounding human fingers generally exhibits elevated humidity levels. When fingers approach or touch objects, RH fluctuations occur in the vicinity. This property enables the development of self-powered touch sensors utilizing the high humidity sensitivity of electrical devices. Shen et al.386 fabricated a tactile sensor based on TiO2 nanowire networks that responded to the presence of a human finger by generating a pulse voltage of 150 mV upon contact. Cheng et al.354 further advanced this concept by developing a rGO-based tactile sensor that responded to finger proximity without direct contact. As illustrated in Fig. 26d, when the finger approached the device at distances of 1, 2, and 3 mm, the sensor generated corresponding current densities of 10, 8, and 4.5 mA cm−2, respectively. This sensor could be utilized for real-time finger tracking applications.
Self-powered electrical technology also holds promise for information storage applications. Liang et al.383 developed a fiber/rGO based moisture energy harvester to generate electrical signals (Fig. 26c(i)). The device exhibited distinct voltage responses depending on whether it was laser-treated. The untreated units produce 0 mV output in response to humidity changes, representing the “0” (off) state. In contrast, the laser-treated units exhibit higher response voltages, representing the “1” (on) state. These flexible sensor units could be integrated into arrays and embedded within wearable fabric. Each sensor unit was positioned at different locations within the fabric and could be programmed to represent either “0” or “1” states, enabling the fabric to transmit digital information, thus offering potential applications in wearable data storage and communication systems. As shown in Fig. 26c(ii), an array of 136 information units is woven into a 3 cm × 7 cm textile, with each unit programable to “0” or “1”. Therefore, this fabric can convey 2136 different kinds of information. Each unit in the network is assigned a coordinate, presenting their position in the fabric. This highly integrated fabric retained excellent mechanical flexibility, allowing it to be conveniently bent to be attached to a person's wrist, or integrated to a face mask as a new type of wearable electronic tag. Furthermore, electrochemical treatment could modulate unit voltage, further enhancing the diversity of electronic tag information expression. For example, Fig. 26c(iii) shows a pattern expression of the “BIT” shape signal based on different voltage responses.
First, the power density and energy conversion efficiency remain relatively low. The performance of electrokinetic generators is largely governed by interfacial effects, as charge generation primarily occurs at the interface between fluids and solid surfaces. However, in situ characterization techniques to explain the potential charge generation and transfer processes at the atomic level in liquid–solid phase interactions remain scarce, hindering a deeper understanding of these phenomena.
Enhancing surface charge density through material optimization and surface morphology modification can improve performance to some extent but charge density cannot be increased indefinitely. Alternatively, improving interfacial slip length can effectively enhance device output. The liquid–liquid interface on SLIPS faces durability challenges due to the depletion of lubrication liquid. Exploring the liquid–polyelectrolyte interface holds great potential for enhancing electrokinetic performance by facilitating higher slip387 lengths and sustainable energy harvesting.
Additionally, the synthesis of materials and the design of devices still lack standardization. There is a need to establish a reliable connection between material properties and output power. Developing standardized fabrication protocols and characterization methods will be crucial for optimizing material performance and ensuring the scalability of electrokinetic energy harvesting technologies.
Finally, most reported devices remain at the laboratory demonstration stage, with significant scale-up challenges hindering real-world applications. The scalability of microfluidic devices remains uncertain, as issues such as flow blockage and charged droplet loss can arise upon upscaling.388 Additionally, real-world operating conditions are more diverse and complex. For example, biofouling and surface contamination can accumulate on the device, leading to performance degradation.212 Therefore, developing electrokinetic systems with stable self-cleaning capabilities could enhance their adaptability across various application environments.
Given the growing demand for sustainable and renewable energy solutions, electrokinetic energy harvesting presents a promising avenue for future advancements in energy generation. The ability to convert ambient mechanical energy, such as fluid flow or droplet impact, into usable electrical power offers a significant step toward reducing reliance on fossil fuels. As research continues to address the current challenges, the potential applications of electrokinetic systems are vast, spanning from self-powered sensors and portable devices to large-scale energy harvesting in industrial or environmental settings. Furthermore, with advancements in materials science, microfluidics, and system integration, electrokinetic energy harvesting could play a key role in the development of autonomous, self-sustaining devices that operate efficiently in diverse, real-world conditions. Continued exploration of this field holds the promise of unlocking new paradigms in sustainable energy technologies, ultimately contributing to the broader goal of expanding energy harvesting strategies and minimizing environmental impact.
REWOD technology relies on capacitance changes caused by droplet deformation on dielectric layers, with a theoretical power density reaching the order of 104 W m−2. It is particularly suitable for energy harvesting and signal monitoring in high-frequency pressure environments such as shoe soles, oil pumps, and vibration devices. However, this technology heavily depends on high-permittivity metal oxide coatings and complex micro-nano fabrication processes, resulting in high per-device costs ($11.5–55.5 per cm2). Additionally, critical issues like synchronous control of multiple droplets remain unresolved, severely limiting its large-scale application.
In contrast, TENGs convert energy through contact electrification and electrostatic induction, offering advantages such as readily available materials, mature fabrication processes, and low manufacturing costs ($0.3–1.3 per cm2). They are suitable for large-area preparation and are widely used in contact- and friction-driven wearable textile energy harvesting, droplet and wave impact energy harvesting, as well as self-powered sensors for biological and chemical detection and environmental monitoring. However, their performance is susceptible to humidity, and the durability of friction materials poses a significant challenge to long-term stable operation.
EKEC technology generates potential through fluid shear-induced interfacial charge dragging effects, enabling adaptation to complex fluid environments such as waves, moisture, and evaporation. It shows potential in energy harvesting and self-powered sensing for scenarios involving droplets, waves, and surface wetting. However, its energy conversion efficiency is relatively low (approximately 1–4%, though the theoretical upper limit is around 40%). Additionally, the fabrication of microchannel structures is complex and costly ($16–66 per cm2), and the lack of standardized synthesis pathways and device design processes results in low overall technological maturity and scalability, hindering industrial progress.
From a developmental perspective, in-depth research on interfacial charge transfer mechanisms and the development of novel functional materials will be key to improving the performance metrics of these three technologies. Future studies should focus on addressing industrial bottlenecks such as scalable fabrication processes, environmental stability, and system integration to advance these technologies from the laboratory to practical applications.
Despite these promising developments, several challenges persist. As detailed in the preceding sections, scaling up the output power of REWOD requires precise synchronization and control of multiple droplets. Similarly, long-term stability, environmental adaptability, and methods to continuously boost energy output are the main challenges faced in TENGs and need further development. And in electrokinetic systems, challenges such as limited atomic-level characterization, energy conversion inefficiencies, and durability concerns continue to hinder large-scale implementation. Moreover, some technologies demand costly materials, complicated fabrication processes or specialized coatings for surface modifications, complicating large-scale manufacturing and commercial uptake. Achieving higher output power and stable performance under variable environmental factors such as salinity, fluid velocity, or pH requires deeper insight into interfacial physics and material optimization. Additionally, integrating these energy harvesters with on-chip signal conditioning, energy storage, and wireless communication modules poses a key engineering challenge on the path to fully autonomous systems.
Moving forward, research should focus on optimizing device architectures to maximize power density while preserving mechanical flexibility, biocompatibility, and environmental resilience. Novel functional materials such as bioinspired self-healing polymers or hybrid nanocomposites could address durability concerns, while scalable fabrication methods (e.g., roll-to-roll or 3D printing) may reduce production costs and complexity. Beyond the laboratory, the synergy of interfacial EHD-based harvesters with advanced computing, machine learning, and internet IoT platforms can unlock transformative solutions for wearable health monitoring, environmental sensing, and industrial automation. With continued innovation and interdisciplinary collaboration, these fluidic energy-harvesting technologies are poised to transition from proof-of-concept demonstrations to market-ready devices, offering sustainable, decentralized power sources that address real-world energy challenges.
Footnote |
† These authors contribute equally to the paper. |
This journal is © The Royal Society of Chemistry 2025 |