摘要
This perspective delves into the advanced technologies of highly sensitive and selective micro-electro-mechanical systems semiconductor hydrogen sensors. It explores the interplay between sensing material electrodynamics, separation material statistical mechanics, as well as chip thermodynamics, and systematically charts the evolution of multilayered, innovative integration techniques among sensing materials, separation membranes, and low-power chips, proposing viable technological pathways. Hydrogen, celebrated as the "most promising new energy carrier of the 21st century," stands out as a clean, secondary energy source with a high combustion heat value and broad availability. Its global adoption is accelerated by rapid advances in hydrogen production, storage, and fuel cell technologies. Yet, hydrogen presents distinct challenges: it is odorless, tasteless, and colorless, has a wide flammable range in air (4–75 vol% volume concentration), a low ignition energy requirement (just 0.017 mJ for a 23 vol% hydrogen–air mix), and a rapid flame spread rate (exceeding 970 m/s at 29.6 vol%). These properties increase the risk of leaks, fires, and explosions, especially as hydrogen readily interacts with materials during storage and use, leading to hydrogen embrittlement and the failure of protective layers in storage and transport systems. Comparing the injury and fatality rates of 10.87% and 2.65%, respectively, for natural gas incidents, it is noted that while the injury rates are nearly identical, the fatality rate for hydrogen incidents is twice as high as that of natural gas incidents (Figure 1A). Thus, ensuring the safety of life and property hinges on the accurate, timely, effective, and comprehensive monitoring of hydrogen concentrations. Developing highly sensitive, selective, and user-friendly hydrogen sensors is vital for the safe, widespread utilization, and growth of hydrogen energy. Current hydrogen sensors are classified into various types like optical, acoustic, thermal conductivity, catalytic, work function, resistive, and electrochemical, based on their operating principles and testing methods (Figure 1B). These sensors detect hydrogen by observing changes in parameters such as refractive index, spectral absorption, temperature, mass structure, and electrical properties, triggered by the interaction with hydrogen gas.1 They convert these changes into optical or electrical signals using digital-analog conversion. Each type presents unique benefits and limitations (Table 1): optical sensors, for instance, offer high specificity and fast signal transmission but are constrained by their size and cost, limiting their deployment and monitoring precision. Acoustic, thermal conductivity, and work function sensors are compact and energy-efficient but struggle with long-term stability, selectivity, and response time. On the other hand, traditional catalytic, electrochemical, and resistive sensors are known for their higher sensitivity, frequency of sampling, simple structure, and ease of signal processing. However, their increased power consumption and dependency on electricity line installations restrict the flexibility in monitoring and patrol route expansion. Additionally, these sensors often fall short in complex environments or when handling mixed-gas atmospheres, particularly in terms of detection range, response speed, interference resistance, and environmental adaptability. Given that no existing hydrogen sensor type fully satisfies all application demands, there remains significant potential for the development of innovative hydrogen sensor technologies in the future. Compared to conventional sensors, resistive micro-electro-mechanical systems (MEMS) semiconductor gas sensors offer advantages such as low power consumption, strong selectivity, easy scalability, high integration, and a wide dynamic measurement range.2 They effectively address the limitations of traditional detection methods, which often involve exclusivity, high energy consumption, and inefficiency, making them an optimal compromise of cost, market, and overall features. The introduction of new-fashioned sensitive nanomaterials and advanced microstructure chips, along with their cross-scale integration, significantly enhances the adaptability of monitoring elements, expands the spatiotemporal coverage of detection, and improves the accuracy of monitoring results and risk event identification in complex operational environments. Currently, the key focus areas in the development of resistive MEMS semiconductor hydrogen gas sensors include (i) the design and enhancement of sensitive materials, (ii) the manufacturing of reliable microchips and their on-chip compatible integration, and (iii) the decoupling and modulation of sensitization mechanisms under multiple external fields. Palladium (Pd)-based metals have been favored as the ideal choice due to their efficient and selective absorption and dissociation of hydrogen gas at room temperature.3 It is widely accepted that there are two resistance modulation mechanisms in Pd hydrogen sensors.4 The first mechanism involves hydrogen atoms entering the Pd lattice, acting as electron scatterers, which increases the bulk resistance of the material, as shown in Figure 2A. The second mechanism occurs when hydrogen atoms enter the Pd lattice, causing lattice distortion and structural volume expansion, leading to the formation of new electrical contact points between Pd micro-/nanostructures. This results in a decrease in resistivity, meaning the sensor exhibits a high-resistance open circuit in a hydrogen-free environment and a low-resistance closed circuit in the presence of hydrogen. By converting the volume signal into an electrical signal, the sensor achieves rapid and high responsiveness to hydrogen, as depicted in Figure 2B. Hydrogen-induced electron scattering-dominated hydrogen sensors feature relatively simple fabrication and flexible material morphology. In 1992, Hughes and Schubert6 pioneered the use of ultrathin Pd-Ni films for hydrogen sensors, capitalizing on the adaptable material shapes. These sensors, particularly the 50 nm Pd-Ni films with 8%–20% Ni, detected H2 in 10 s at 4% concentration and 20 s at 1%, with a detection limit under 0.1%. Although cost-effective and sensitive, their response time of 10–20 s falls short of the Department of Energy's recent benchmarks by an order of magnitude. To improve response times, Wang and colleagues7 crafted sensors with Pd nanoparticles (NPs) on nanoporous alumina, achieving a 1 s response and 500 ppm detection limit for 1% H2. Another approach involves Pd nanostructures at the electrical percolation threshold, which exhibit decreased resistance in the presence of H2, contrary to the typical increase seen in Pd resistors. Kaltenpoth et al.8 noticed this phenomenon in ultrathin films within silicon channels and Xu et al.9 later enhanced this with Pd island films on siloxane-coated glass, achieving 70 ms response times and 25 ppm detection limits. Ramanathan et al.10 found that only a narrow Pd thickness range (2–3 nm) enables this rapid detection. The "fractured-wire" resistive sensors maintain Hughes' design simplicity while aiming for quicker response and recovery. Since 2000, Penner and colleagues11-14 refined single Pd nanowire (NW) sensors, enhancing detection sensitivity and speed. Notably, Offermans15 and Yang12 reported ppm level detection and a few seconds response at several percentage H2 concentrations with nonfracturing Pd NWs. Advancements in nanoscience have led to various Pd nanostructures like thin films, NWs, nanotubes, and networks for H2 sensing. Their design enhances detection through greater surface area-to-volume ratios and nanoscale effects. Table 2 summarizes the recent and representative Pd-based H2 sensors.16-27 Despite significant research and application in hydrogen sensors, Pd metal materials still face unresolved challenges, particularly in balancing sensitivity and the range of detectable concentrations. Present sensors experience diminished resolution and slower response rates at hydrogen levels above 2 vol%, attributed to Pd's inherent properties (Figure 2C,D).5, 28-30 At lower concentrations (below 1 vol%), Pd forms an α-phase hydride (PdHx) with hydrogen atom fraction x < 0.02. This changes to a mixed α and β-phase (0.02 < x < 0.058) as the concentration rises. At concentrations over 2 vol%, a significant influx of hydrogen atoms into the Pd lattice causes severe lattice distortion, turning PdHx entirely into the β-phase (x > 0.58) and leading to about 3.5% lattice expansion, and even brittleness and flaking.31-33 This results in severe sensor resistance hysteresis or failure. Hydrogen saturation in Pd also causes resistance saturation, making it difficult to distinguish higher gas concentrations. The sensor's response time relies on dynamic equilibrium processes like surface adsorption/desorption, diffusion, and phase transition.34, 35 Adsorption and phase transition, with high activation energies, are critical in the Pd-hydrogen reaction, significantly affecting the sensor's response time.36 Typically, hydrogen's explosive limit is above 4 vol%, so a hydrogen sensor that retains high sensitivity and rapid response in high-concentration hydrogen environments is truly valuable for practical applications. Enhancing surface-selective adsorption and minimizing phase transitions in Pd-based materials are key strategies. Recent advancements in nanotechnology have led to numerous studies demonstrating that nanostructured Pd can offer more active sites for Pd-hydrogen reactions, thus improving the efficiency of hydrogen atom adsorption on metal surfaces. Notably, in ultra-small Pd NPs, NWs, and nanofilms, there's an increase in hydrogen atom solubility in the α-phase PdHx and a decrease in the β-phase, effectively narrowing the coexistence range of these phases.16 Building upon this, alloying Pd nanostructures can further inhibit excessive hydrogen dissolution and lower lattice constants, thereby expanding the measurement range of sensors. Some research has already achieved linear responses for hydrogen gas concentrations as high as 10 vol%.37, 38 However, this strategy often comes at the expense of sensitivity, with response values dropping below 10% at 4 vol% hydrogen concentration. Additionally, the wet chemical synthesis method, commonly used for Pd alloying at the nanoscale, poses challenges in terms of randomness, reproducibility, and consistency, impeding commercial mass production. Striking a balance between anti-interference capabilities and response speed is another major challenge in hydrogen sensor development, especially in chemical scenes. Trace CO levels, usually around 0.2 ppm in the atmosphere, increase near urban or industrial areas, along with other molecules with high polarity like H2S, SO2, and NO2, often posing a risk of poisoning Pd-based materials.39, 40 On the other hand, at H2 concentrations below 1%, α-PdHx resistance response relies on hydrogen adsorption. Yet, oxygen from the air may also bind to Pd active sites, thus obstructing these sites and slowing down hydrogen adsorption. To counteract these problems, various strategies have been adopted to minimize cross-sensitivity and enhance device tolerance (Figure 3A).41 One effective approach involves using physical barriers such as fluorinated ethylene propylene, polyimide, or polysulfone hydrogen-permeable membranes to prevent toxic substances from contacting sensitive materials.42, 43 These barriers are either applied directly to the material surface or positioned before diffusion pores. Research has also focused on inorganic porous membranes like zeolite molecular sieves,44 graphene,45 and organic framework membranes46 to create composite sensitive materials. These materials use adjustable pore sizes to filter out impurities, allowing hydrogen to concentrate on the permeation side, thereby achieving preliminary separation in mixed atmospheres for highly specific responses. Recently, Koo et al.47 used ZIF-8 as a selective filter on Pd-based sensors, enhancing the detection of hydrogen gas. They first created Pd NWs through a lithographic process, then coated these wires with ZIF-8, forming Pd NWs@ZIF-8. On account of ZIF-8's small pores (just 0.34 nm wide), hydrogen molecules, being smaller (0.289 nm in diameter), passed through quickly, while larger oxygen molecules (0.346 nm) were filtered out. This approach resulted in a significant reduction of oxygen's interference with the hydrogen sensing of the Pd NWs. Despite a slight reduction in overall sensitivity, the presence of ZIF-8 drastically improved the sensor's speed: for detecting 1% hydrogen, the response time dropped from 164 to just 7 s, and recovery time fell from 229 s to 10 s. In addition, Kalidindi et al. reported Schiff-base-linked TpPa-SO3H covalent organic framework (COF) hybridized Pd NPs as a selective H2 detector.48 TpPa-SO3H COF features extended conjugated π-electron systems that support effective signal transmission throughout its structure. When tested, the exfoliated Pd@TpPa-SO3H composite demonstrated a 29 ± 1% change in resistance—corresponding to a negative sigmoidal response—at 120°C in the presence of 1% hydrogen gas. The sensor showed impressive response and recovery times, clocking in at 5.3 ± 0.5 and 3.1 ± 1 s, respectively. Its ability to consistently and efficiently recover (>99%) across hydrogen concentrations from 0.2% to 1% marks this material as a capable, reusable hydrogen sensor. Additionally, interferents with larger kinetic diameters, such as SO2 (0.36 nm), NO2 (0.33 nm), and CH4 (0.38 nm), can be effectively screened by the COF sheath. Due to the hydrophilic nature of −SO3H groups, there was a notable increase in the response and recovery times when the humidity increased, thus showing humidity-dependent hydrogen-sensing characteristics of such imine-based COF filtration layer hybridized Pd NPs. Although various hydrogen separation membranes are showcasing effective separation and purification capabilities, a prevalent issue remains in the context of sensor applications: How do we develop a hydrogen separation membrane that simultaneously possesses high selectivity, sufficient stability, good reproducibility, and meets the required standards for permeation rates? Polymer membranes, known for their simplicity, cost-effectiveness, and compatibility with traditional integrated circuit processes, face limitations in separation performance. Increased selectivity often leads to reduced permeability, known as the Robeson upper bound. Long-term stability is also compromised by plasticization and aging. In contrast, inorganic membranes offer high resistance to temperature and pressure but suffer from poor film-forming properties and a stark trade-off between permeability and selectivity. Organic framework membranes stand out for their chemically and structurally diverse units and precise pore size control, surpassing the Robeson limit with their high porosity and ultra-thin nature (Figure 3B,C).41 Yet, challenges remain in material synthesis integration with electronic device fabrication, maintaining stability in complex monitoring environments, and ensuring consistent performance. Physical methods like drop-casting and liquid-phase epitaxy, though simple, struggle with substrate adhesion, while chemical methods like in situ hydrothermal synthesis offer better adhesion but are more complex and risk contaminating and damaging device bases. Stability issues also persist, with many organic framework materials susceptible to degradation under extreme environmental conditions.49, 50 MEMS gas sensor chips are predominantly structured with micro hot-films and micro hot-bridges, employing bulk silicon etching for their fabrication (Figure 4A).51-56 This choice is driven by the need for metal-oxide semiconductor gas-sensitive materials to operate at high temperatures (100−350°C), where silicon-based substrates effectively minimize thermal conductivity.57 Advanced micron-scale design further reduces heat loss from radiation and convection, achieving low power consumption at the milliwatt level. The efficiency of these sensors depends on micro hot-plate heaters and multilayer composite dielectric films. Platinum (Pt) films are favored as heater electrodes in these systems due to their high electrical conductivity and excellent temperature sensitivity. However, challenges with adhesion on the smooth silicon oxide surface led to the use of titanium (Ti) as an intermediate layer, necessitating complex manufacturing processes and leading to issues in thermal expansion compatibility under high temperatures. Pd-based MEMS hydrogen sensors, designed for low-temperature operation but optimal at 70 − 90°C, often use continuous or intermittent pulse heating (120−150°C) to enhance response time and reduce effects of humidity, although this can introduce thermal instability. Addressing this, key development areas include optimizing film wettability and thermal stress compatibility, designing efficient heating structures, and researching cost-effective methods to further reduce power usage, increase integration, and refine processing precision. Alumina and aluminum nitride, used in place of SiON (silicon oxide, silicon nitride) dielectric films, offer superior thermal conductivity. This enhancement reduces thermal zone gradients and improves thermal equilibrium rates, significantly mitigating reliability issues like temperature drift and thermal expansion mismatch.58 The excellent bonding and higher dielectric constant of these ceramic substrates improve metal film quality and the integration of sensitive nanomaterials, facilitating simpler film growth or laser micromachining processes.59 Recent research indicates that high-adhesion alumina substrates can effectively suppress β-phase transitions in Pd nanostructures at high hydrogen concentrations (>4 vol%), due to induced stress at the substrate contact area.5 This advance in materials science holds promise for broadening the application of low-power, Pd-based MEMS sensor chips. In addition, the core electrode area of MEMS devices is often <1 mm2, achieving controlled and localized assembly of sensitive materials on microstructured substrates become increasingly difficult. This results in poor quality control and hampers the device ability to capitalize on the stable and consistent process advantages inherent in micromachining technology.60 Moreover, effectively harnessing the intrinsic properties of nanomaterials, such as surface and size effects, and the synergistic traits from cross-scale material/device coupling, also becomes challenging. Developing methods for seamlessly integrating these sensitive materials with MEMS processes is critical for creating commercially viable, high-performance miniaturized semiconductor gas sensors. Traditional processes like optical lithography and electron beam lithography, while MEMS-compatible, and often incur higher costs in crafting complex three-dimensional structures, primarily due to limitations in film characteristics and process coupling. Printing processes offer cost advantages but suffer from low pattern resolution, and the introduction of third-phase materials can disrupt the microstructure of sensitive materials, leading to contamination (Figure 4B).51-53 Block copolymer (BCP) self-assembly technology offers a promising avenue for fabricating 10 nm-scale high-precision sensitive material structures on chips.61 BCPs, made of two or more distinct homopolymers bonded by covalent links, undergo microphase separation due to their thermodynamic incompatibility. This separation forms large-area, high-density, periodic nanostructures, with morphology dictated by the homopolymers' volume fraction and resolution by the Flory–Huggins interaction parameter and overall polymerization degree.62 In lithography, continuous lamellar and cylindrical phases facilitate the creation of high-contrast, orderly nanostructures such as NWs or dot arrays through selective etching, pushing beyond the diffraction limit of traditional methods.63 Techniques combining dry or wet etching, infiltration deposition, and mechanical stripping allow for high-fidelity transfer of BCP-formed nanostructures to substrates. Integrating BCP assembly with top-down strategies like boundary epitaxy and field-induced assembly is key to uniform structural dimensions and maintaining defect density within acceptable semiconductor industry standards (~5% cumulative fluctuation deviation). Laser-guided assembly, particularly laser-localized annealing, aligns well with MEMS processes, offering precision and efficiency (Figure 4C).54, 64 During slow direct-writing, laser radiation creates sharp temperature gradients, rapidly transforming BCPs into a glassy state and confining microphase separation, thus reducing structural defects.55 This technique also enables deeper crystal domain orientation control, facilitating the production of near-perfect, high-precision sensitive material layers. This approach holds great promise for delivering consistent, reliable patterned solutions for "in situ rapid printing" on MEMS chips, enhancing batch-to-batch consistency.56 Research in sensor mechanisms focuses on decoupling the electrodynamics (electric transport) of material sensitization, influenced by multiple external fields like light, electricity, and heat, from the thermodynamics of target molecule adsorption-separation (gas transport). The performance of sensors largely depends on the state density of sensitive materials and their internal solid periodic potential fields, encompassing crystal defects, band structures, Debye lengths, and interface barriers. Modulating these factors affects the emission and transport of free electrons near the Fermi surface, crucial for the gas-electric response.65 Understanding the limited carrier transfer mechanism in sensitive materials is essential for decoupling sensor mechanisms. The conductive model that describes carrier transport bridges the gap between microscopic gas-sensitive mechanisms and macroscopic performance. Over the past decades, three main models have emerged for resistive gas-sensitive materials: the oxygen ionization model for N-type semiconductors, the oxygen vacancy model for P-type semiconductors, and the quantum tunneling model for nano-metal systems.66 The first two models, grounded in thermionic emission theory, agree that gas-phase substances interacting with material surfaces play a vital role in conductivity. They emphasize that only the surface's free electron concentration is altered, while electrons at fixed energy levels don't directly contribute to conductivity. Electron transport is influenced by the material's microscopic morphology and external factors like bias voltage and field coupling. The third model, applicable to small metal NP systems, suggests electrons move via tunneling and percolation, not classical Ohmic transport.67 At low temperatures (<100 K), when the conductivity of NPs (G) linearly correlates with ln(G)-T−1/2, Coulomb blockade effects create gaps in state density at the Fermi level, localizing metal energy levels. Mott's68, 69 variable range hopping model becomes the primary electron transport mechanism. As temperatures rise and conductivity correlates with ln(G)-T−1, electron tunneling barriers thin, allowing for near-neighbor hopping. Klafter and Sheng's70 thermally activated tunneling KS model becomes the predominant electron transport mechanism. For resistive Pd-based metal hydrogen sensors, thermally activated tunneling is the key transport mechanism. However, this model doesn't fully account for other external factors, geometric forms, or material chemistry affecting grain boundary barriers. Thus, the classical quantum transport model has limitations in identifying descriptors that link material properties with gas-electric response, impeding the exploration of the relationship between key species characteristics and sensing activity. Integrating defect chemistry theories from the oxygen ionization/vacancy models and developing a more universal multifield critical transition conductive model is necessary for a deeper understanding of material sensitization mechanisms and guiding enhancements in sensing activity and selectivity. Gas transport promotion mechanisms are also crucial for enhancing the performance of sensing systems, encompassing two spatiotemporal scales: gas molecule surface adsorption based on quantum mechanics and gas molecule permeation-diffusion within porous membranes based on molecular mechanics and statistical thermodynamics. To uncover the essence of selective enhancement of target gases in separation-sensing material systems, it is essential not only to experimentally measure adsorption isotherms, penetration curves, functional groups, valence changes, and cross-sensitivity but also to use multiscale theoretical simulations for analyzing interactions between gases, sensitive materials and separation membranes at electronic-atomic-molecular levels (Figure 5).74, 75 At the subnanometer scale, density functional theory (DFT) is a widely used method internationally.71 On the one hand, DFT calculations can be used for sensitive materials to obtain their adsorption energy for target molecules, assessing the ease of gas molecule adsorption on the material surface.72 The evolution of the electronic structure (energy bands, state density) of sensitive materials before and after gas molecule adsorption can be calculated to understand the electron transfer between gas and solid. Additionally, calculating the transition states of gas molecules on the surface of gas-sensitive materials further reveals diffusion barriers, redox intermediates, and evolutionary paths of gas molecules. On the other hand, DFT calculations can study the atomic charges and resulting electrostatic potentials in skeleton materials doped within separation membranes, greatly aiding the explanation of gas adsorption, separation, and diffusion behaviors in the membranes. Microscopic energy decomposition can also identify the proportions of various interactions (such as dispersion, electrostatic, and orbital interactions), offering isolation of factors and system parameter alteration capabilities.76, 75 This is suitable for revealing the synergistic effects of single or multiple factors, thereby determining the essence of the adsorption action. Overall, quantum mechanics methods are most applicable to femtosecond/picosecond and subnanometer spatiotemporal scale studies. However, examining dynamic processes on the nanosecond/second and micron-nanometer spatiotemporal scales typically requires considerable computational resources and time. In the study of gas equilibrium and transport at the mesoscopic level, molecular simulation methods, integrating molecular mechanics and statistical thermodynamics, are utilized. These methods analyze macroscopic data like adsorption isotherms, selectivity, and heat of adsorption.77 The primary techniques are molecular dynamics (MD) and Monte Carlo (MC) simulations.78 MD excels in scenarios with long time frames and near thermodynamic equilibrium, effectively simulating gas diffusion paths, coefficients, and structural changes during adsorption. However, its reliance on statistical thermodynamics demands significant computational resources. MC differs notably from MD. In MD, gas molecule motion adheres to Newton's laws, guided by a potential function. MC, on the other hand, maintains fixed chemical potential, temperature, and volume, allowing for fluctuating particle numbers.79 Each MC step involves moving a randomly selected particle, accepting the move if it lowers energy, thereby reaching equilibrium. The grand canonical MC (GCMC) method, often used for gas adsorption and separation analysis, calculates internal energy, adsorption amounts, and free energy with lower computational needs, effectively replicating experimental data.80, 81 Nevertheless, MC methods based on force fields face challenges in accurately representing special adsorption sites, like unsaturated metal sites.82 To address this, a combination of force field methods and DFT is required. This approach begins with determining the potential energy curve between gas molecules and active sites, followed by incorporating fitted force field parameters into MD simulations to obtain adsorption isotherms at crucial active sites.83 Advancements in the domain of hydrogen-sensing technologies within the realm of new energy applications are poised for transformative growth, propelled by the amalgamation of cutting-edge sensitive material systems and state-of-the-art separation membranes. These innovations, when harmoniously integrated with the precision engineering of MEMS sensor chips, set the stage for a profound leap in performance and reliability. Moreover, the meticulous dissection of the underlying sensing mechanisms, informed by the most recent scientific insights, will enhance our ability to devise robust and sensitive detectors. Reflecting on the current trajectory and the intricate challenges that permeate this field of research, we can distill a set of strategic directions to guide future explorations in this specialized field. Compared to multi-element alloys and high-entropy alloy systems, bimetallic Pd alloys exhibit superior mechanical properties, thermal stability, and resistance to poisoning. Common alloys include PdNi, PdCu, PdAg, PdAu, and PdIr.84 At low hydrogen partial pressures, Ag, Au, and Ir increase the lattice parameter of Pd, reducing the energy required for hydrogen occupancy, which promotes hydrogen dissolution. At high hydrogen partial pressures, Ag, Au, and Ir fill some of the vacant 4d electron states of Pd, reducing the number of ele