Additive manufacturing of MXene electrodes: from rheology-tunable nanoinks to size-scalable integrated electronics

Qiancheng Zhao a, Hao Liu a, Chuan Liu e, Takeo Minari d, Se Hyun Kim c, Xiaowu Tang *ab and Xuying Liu *a
aSchool of Materials Science and Engineering, Zhengzhou Key Laboratory of Flexible Electronic Materials and Thin-Film Technologies, Zhengzhou University, Zhengzhou 450001, China. E-mail: liuxy@zzu.edu.cn; xwtang@zzuli.edu.cn
bCollege of Material and Chemical Engineering, Institute of New Energy Science and Technology, School of Future Hydrogen Energy Technology, Zhengzhou University of Light Industry, Zhengzhou 450001, P. R. China
cSchool of Chemical Engineering, Konkuk University, Seoul 05029, Korea
dPrinted Electronics Group, Research Center for Functional Materials, National Institute for Materials Science (NIMS), Tsukuba, Ibaraki 305-0044, Japan
eState Key Laboratory of Optoelectronic Materials and Technologies, School of Electronics and Information Technology, Sun Yat-sen University, Guangzhou 510275, P. R. China

Received 24th July 2025 , Accepted 11th September 2025

First published on 16th September 2025


Abstract

Additive manufacturing (AM) has emerged as a transformative strategy for translating nanoscale materials into wafer-scale functional architectures, empowering the scalable fabrication of advanced electronics and energy systems. MXenes—a class of two-dimensional transition metal carbides and nitrides—exhibit exceptional electrical conductivity, tunable surface terminations, and mechanical compliance, positioning them as ideal candidates for AM-driven printed electronics. This review uniquely emphasizes the rheological–structural–functional correlations that underpin the evolution of MXene inks from individual nanosheets to architected wafer-scale systems. We dissect critical rheological parameters—including shear-thinning behavior, yield stress, and colloidal stability—and their decisive roles in shaping printability, pattern resolution, and post-print structural fidelity across AM techniques such as direct ink writing, spray coating, and electrohydrodynamic printing. Furthermore, the multiscale influence of rheology on printed micro-/nanostructures and their downstream impacts on electronic, electrochemical, and sensing performance are systematically analyzed. We summarize the latest advances in scalable MXene integration for applications including microsupercapacitors, field effect transistors (FETs), and self-powered biosensors. Finally, we highlight future research directions encompassing machine-learning-assisted ink formulation, hierarchical porous structure design, and eco-conscious processing paradigms. These insights pave the way for intelligent ink systems and hybrid device architectures, propelling MXene electronics toward multifunctional, sustainable, and industry-compatible futures.



Wider impact

This review provides a comprehensive framework that bridges material design, ink rheology, and additive manufacturing technologies to accelerate the deployment of MXene-based printed electrodes in next-generation flexible electronics and energy devices. By establishing the critical coupling between rheological behavior and device performance, the article delivers not only academic insights but also scalable manufacturing guidance applicable to printed transistors, energy storage, and wearable systems. The interdisciplinary perspective offered here addresses a major gap in the current literature and holds significant potential for advancing sustainable, cost-effective, and high-throughput production of multifunctional electronic components. This work is expected to guide future research on green manufacturing, ink formulation optimization, and device-integrated design, contributing to innovations in environmentally conscious electronics and intelligent sensing platforms.

1. Introduction

Two-dimensional transition-metal carbides, nitrides, and carbonitrides (MXenes),1 as an emerging family first reported in 2011 by Naguib et al. at Drexel University via selective removal of the A-layer from MAX phases using hydrofluoric acid, who also introduced the term “MXenes” (Advanced Materials)2,3 have demonstrated remarkable promise in printed electronics due to their exceptionally high electrical conductivity (>9000 S cm−1), tunable surface terminations (–O, –OH, –F, etc.),4 and inherent pseudocapacitive behavior.5,6 Particularly, in the fields of flexible energy storage devices (such as micro-supercapacitors and zinc-ion hybrid capacitors) and sensors,7–10 MXene-based printed electrodes have become a research hotspot due to their high volumetric capacitance (up to 1500 F cm−3) and mechanical flexibility.11,12 Additive manufacturing technologies—such as direct ink writing (DIW) and inkjet printing13—have revolutionized the fabrication of electrochemical energy storage devices by enabling high design freedom and rapid prototyping.14–17 These techniques also offer a new paradigm for the customizable micro-structuring of MXene electrodes, thereby enhancing their electromagnetic, thermal, and energy conversion performances.18 At the core of this advancement lies the synergistic optimization of ink rheology and device performance,19,20 which provides an efficient and sustainable pathway for manufacturing flexible electronics, wearable devices, and miniaturized energy systems.21,22

In recent years, researchers have achieved multiscale control of MXene structures—from nanoscale sheet alignment to macroscale porous architectures—by tuning the viscoelasticity, thixotropy, and shear-thinning behavior of MXene inks.23 For instance, the introduction of rheological modifiers such as xanthan gum has enabled DIW of MXene inks.24 Surface termination engineering with oxygen-containing anions has enabled the fabrication of highly conductive (916 S cm−1), ultra-stable (>1000 cycles) paper-based printed supercapacitors.25 These breakthroughs reveal that rheological tuning not only affects the printability of MXene inks but also directly determines ionic diffusion kinetics (reducing the time constant to 0.8 s) and interfacial charge transfer efficiency.26,27

Moreover, additive manufacturing techniques, such as automated spraying, drop casting, and 3D printing, have enabled the scalable integration of MXene electrodes in high-performance devices,28 including silicon heterojunction solar cells (conversion efficiency >20%),29 GaN photodetectors (responsivity of 64.6 A W−1, dark current reduced by three orders of magnitude),30 and flexible lithium-ion batteries (capacity retention of 98.9% after 400 cycles).31,32 These applications have overcome the limitations of conventional noble metal-based electrodes and interfacial bottlenecks, demonstrating the transformative potential of additive manufacturing for efficient, low-cost, and flexible electrode design.33–36

Additive manufacturing, with its high precision, low-temperature processing, and scalability, facilitates the rational construction of 3D porous architectures and heterointerfaces—such as Ti–O–C bonding in MXene–RHGO composite electrodes—enabling efficient ion transport and stable capacity output in high-loading sodium-ion batteries.37 Combined with in situ nanoparticle integration (e.g., MXene embedded with silver nanoparticles), flexible designs have advanced the application of stretchable pseudo-capacitors in wearable electronics.35 Additionally, this technology supports rapid prototyping and structural optimization, helping MXene and carbon-based electrodes achieve low-cost, high-catalytic-activity thermal energy harvesting in thermocell applications.38

In the field of flexible electronics, additive manufacturing enables performance optimization through the directional assembly of devices, like indium oxide nanowire transistors.39 It also leverages rational material design guided by octahedral net charge descriptors (Qoct) to suppress hydrogen evolution reactions and optimize proton transport pathways, paving the way for high-capacity, highly stable aqueous proton batteries and sustainable electronic devices.40 Overall, additive manufacturing exhibits revolutionary potential in electrode micro/nanostructure design, functional integration, and cross-disciplinary applications.37,39,40

However, MXene-based printed electrodes still face three major challenges: (1) self-restacking of nanosheets significantly reduces the effective surface area (from a theoretical 98 m2 g−1 to an actual ∼32 m2 g−1), limiting capacity performance.41,42 (2) Conventional extrusion-based 3D printing is restricted by ink rheology (shear modulus must be maintained in the range of 103–104 Pa s), hindering the construction of complex 3D conductive networks.11,43 (3) The operating voltage of devices is generally limited (≤0.6 V) due to the coupled effects of electrolyte decomposition and electrode oxidation, resulting in a bottleneck for energy density enhancement.44 To address these issues, emerging integrated structure–function design strategies have shown unique advantages.45 For example, constructing MXene@COF heterostructures can boost NaCl adsorption capacity in capacitive deionization electrodes up to 53.1 mg g−1.46 Directionally aligned MXene electrodes fabricated via micro-continuous liquid interface production (μCLIP) technology have demonstrated an increase in volumetric capacitance compared to traditional structures.47

This review systematically summarizes recent advances in MXene printed electrodes driven by additive manufacturing technologies, with a focus on elucidating the structure–property relationships among ink formulation, rheological behavior, printed architecture, and device performance. Special attention has been given to emerging directions such as multi-material co-printing and 4D printing-enabled shape reconfiguration. Furthermore, the integration of machine learning for ink formulation design is discussed as a pathway toward next-generation printed electronics with high energy density (>50 Wh kg−1) and ultra-flexibility (bending radius <1 mm), providing both theoretical guidance and technical strategies for future development (Fig. 1).48–51


image file: d5mh01421b-f1.tif
Fig. 1 Conceptual diagram for review of MXene-based printed electronics. “Screen printing, bar coating, slot-die coating and laser etching”: Reproduced with permission.52 Copyright 2025, Springer Nature. “Ink(up)”: Reproduced with permission.53 Copyright 2022, Elsevier. “Ink(down)”: Reproduced with permission.54 Copyright 2020, Springer Nature. “3D printing (up)”: Reproduced with permission.31 Copyright 2023, Wiley-VCH. “3D printing (down)”: Reproduced with permission.32 Copyright 2022, Wiley-VCH. “Large area paper transistors”: Reproduced with permission.39 Copyright 2025, Wiley-VCH. “Synthesis of 3D K+-Ti3C2Tx”: Reproduced with permission.55 Copyright 2021, Wiley-VCH. “The MC-fabric humidity sensor fabrication process”: Reproduced with permission.56 Copyright 2022, Elsevier. “High 3D printability”: Reproduced with permission.57 Copyright 2024, Elsevier. “T–S–P electrode”: Reproduced with permission.38 Copyright 2023, Wiley-VCH. “Low-tortuosity electrode fabrication”: Reproduced with permission.58 Copyright 2023, Wiley-VCH. “Integrated bulk–printed ZHCs”: Reproduced with permission.14 Copyright 2024, Wiley-VCH. “Flexibility”: Reproduced with permission.59 Copyright 2022, Springer Nature.

2. Overview of additive manufacturing technologies

2.1 Classification of MXene printing techniques and application scenarios

Additive manufacturing (AM) encompasses a diverse set of printing techniques that enable the layer-by-layer construction of functional materials with tailored microstructures, offering considerable potential for fabricating MXene-based electrodes. The selection of an appropriate AM strategy is intimately tied to the rheological properties of MXene inks and the intended application scenario. Table 1 compares the commonly used printing techniques for MXene inks—inkjet, direct-ink writing, screen printing, aerosol-jet, electrohydrodynamic (EHD) jet, gravure, and flexographic—highlighting their viscosity windows, achievable resolution/line width, layer thickness per pass, throughput, and substrate/annealing requirements. In this section, we categorize mainstream AM techniques employed in MXene printing, including extrusion-based, spraying/coating-based, and laser-assisted methods, highlighting their respective advantages, limitations, and practical implications.
Table 1 Comparison of key features of different printing techniques
Property dimension Printing accuracy Ink compatibility Production efficiency Structural complexity Ref.
Extrusion-based 3D printing Limited precision due to nozzle size Rheology modifiers Slow fabrication Geometry-limited 11, 60–63 and 65
Spraying/coating Low throughput. Low-viscosity, dispersed inks Mass-production Pattern limitation 29, 68–70 and 72
Laser-assisted manufacturing Micrometer-scale circuitry Laser-postprocessed precursor deposits Inefficient scanning 3D-structuring 12, 49, 59 and 73–75
Inkjet printing 10–100 μm scale. <20 mPa s viscosity Moderate throughput 2D-limited 13, 54 and 76–78
Screen printing 50–100 μm scale High-viscosity pastes (1–50 Pa s) Ultra-efficient Layer-dependent 19, 24, 79 and 80


As detailed in Section 3, the feasible ranges of viscosity, yield stress, thixotropy, and wetting behavior map directly onto these extrusion, spraying/coating, and laser-assisted routes, indicating the method selection for specific architectures.

2.1.1 Rheological limitations and structural diversity challenges in extrusion-based 3D printing. Extrusion-based 3D printing is one of the mainstream additive manufacturing techniques currently used for fabricating MXene electrodes.60 However, its application is significantly limited by the rheological properties of MXene inks.61,62 To ensure stable extrusion and shape retention, the ink must possess high viscosity, shear-thinning behavior, and rapid solidification capability. For example, because the ink must simultaneously exhibit high viscosity, pronounced shear-thinning, and rapid gelation, the printable design space is narrow, which limits structural diversity and hinders the fabrication of complex porous lattices or high-precision microfeatures.11 Studies have shown that conventional extrusion-based 3D printing typically yields simple layered or honeycomb-like structures.63–65 In contrast, achieving high-performance MXene electrodes relies on vertically aligned ion diffusion channels and interconnected 3D networks, which impose much higher demands on both ink formulation and printing parameters.5,66 Some researchers have attempted to improve printability by introducing rheology modifiers such as xanthan gum,67 but these additives may compromise the intrinsic electrical conductivity or electrochemical activity of MXene materials.7,20
2.1.2 High-throughput potential of spraying/coating techniques. Spraying techniques offer unique advantages for MXene electrode fabrication, particularly due to their high throughput and ability to rapidly form large-area films.68 For instance, solution-based spraying enables the uniform deposition of MXene thin films on flexible substrates. When combined with templating or masking strategies, this method allows for the patterning of electrodes suitable for the large-scale production of transparent and flexible electronic devices.29,69,70 Spraying can also be integrated with other processes, such as electrospinning or uniaxial stretching, to build multilayer composite film structures that enhance both mechanical flexibility and electrochemical stability.71 Moreover, ref. 26 proposed a “micro-cup” filling strategy based on vertical interface design. This approach significantly improves ion diffusion rates and volumetric capacitance in MXene electrodes, further validating the potential of spraying technologies for high-performance energy storage devices.72
2.1.3 Laser-assisted manufacturing and microstructure control. Laser-assisted additive manufacturing techniques (such as laser sintering and laser-induced graphitization) provide new pathways for the precise control of the microstructure of MXene electrodes.49,73,74 For example, ref. 59 reported a laser-sintering-based direct ink writing technique for MXene, which achieved directional alignment and densification of MXene nanosheets via localized high-temperature treatment, ultimately obtaining micrometer-scale circuits with electrical conductivity as high as 6900 S cm−1.59 Laser technology can also be utilized for the surface functionalization and modification of MXene composites. For instance, laser etching can be used to introduce hierarchical pores or modulate the distribution of surface functional groups in MXene films, thereby enhancing their electrochemically active sites.12 Furthermore, a laser-assisted oxide dispersion strengthening (ODS) strategy has been applied for the microstructure optimization of metal–MXene composite electrodes, addressing issues of thermal stress and deformation in high-resolution printing.75

Having defined the printing method landscape and its rheological constraints, we next discuss how multi-material integration leverages these routes to engineer hybrid functionalities and overcome intrinsic MXene limitations (Fig. 2).


image file: d5mh01421b-f2.tif
Fig. 2 MXene printing technology classification and applicable scenarios. (A) MXene electrodes with designed structures using 3D printing to promote electrolyte permeation and ion diffusion. Reproduced from ref. 11 with permission from American Chemical Society, Copyright 2022. (B) Screen-printing flexible MXene patterns for EMI, Joule heater and piezoresistive sensor devices. Reproduced from ref. 24 with permission from John Wiley and Sons, Copyright 2022. (C) The nematic phase leads to the formation of the Ti3C2Tx architecture by van der Waals forces during solvent evaporation (scale bar: 1 μm). Reproduced from ref. 66. with permission from John Wiley and Sons, Copyright 2025. (D) Schematic representation of the automated spraying apparatus of the large-scale deposition of Ti3C2Tx flakes as the back electrode for SHJ solar cells. Reproduced from ref. 29 with permission from American Chemical Society, Copyright 2022. (E) Schematic showing the fabrication process of the PEG@PAN coaxial fiber film and PM composite film. Reproduced from ref. 71 with permission from John Wiley and Sons, Copyright 2024. (F) Optical image of high-resolution integrated circuits fabricated through direct MXene printing. Scale bar, 10 mm. Inset: Bent MXene circuit. MXene printed “ZJU” and “EMPA” logos on curved surfaces. Scale bar, 20 mm. SEM image of periodically printed MXene lines with a gap of 30 μm. Scale bar, 500 μm. Inset: Line gap of 30 μm. SEM images of MXene lines with different gaps ranging from 3 μm to 25 μm. Scale bar, 50 μm. Reproduced from ref. 59 with permission from Springer Nature, Copyright 2022.

2.2 Multi-material integration and functionalized design

The integration of MXenes with diverse functional materials through additive manufacturing enables the construction of hybrid electrodes with enhanced electrochemical, mechanical, and environmental properties. These multi-material strategies not only address the intrinsic limitations of pristine MXene, such as sheet restacking and oxidation susceptibility, but also introduce new functionalities, such as stretchability, transparency, or multi-mode sensing. In this section, we summarize the composite design principles and structural configurations that support the functional diversification and performance enhancement of MXene-based electrodes.
2.2.1 Composite strategies for MXene with organic/inorganic materials. The combination of MXene with organic materials can significantly improve its flexibility and interfacial compatibility. For example, MXene forms conductive composite electrodes with polypyrrole (PEDOT:PSS) through hydrogen bonding, achieving both high transparency and mechanical stability.81 Composite formation with inorganic materials focuses on enhancing MXene's charge transport capability or chemical stability.82 Examples include the heterostructure design of MXene with covalent organic frameworks (COFs), which boosts the electrode's specific capacity and cycle life via interfacial charge transfer effects.83 Ref. 84 also reported a composite system of MXene with perovskite nanosheets, realizing flexible capacitors with a high dielectric constant via laser additive manufacturing.84 Furthermore, composite strategies combining MXene with carbon fiber, graphene, or metal oxides have been widely explored to address MXene sheet stacking issues and construct hierarchical conductive networks.85–88
2.2.2 Multi-material layout for flexible electronic devices. In flexible electronic devices, the multi-material integration of MXene electrodes must balance conductivity, mechanical compatibility, and functional synergy.89,90 For instance, kirigami patterning of MXene/PDMS composite films enables stretching strains exceeding 300% while maintaining high conductivity.91 Ref. 91 further proposed a “sandwich” structural layout, where MXene electrodes are integrated layer-by-layer with dielectric and encapsulation layers via stencil printing, achieving an ultralow detection limit (66.3 nF kPa−1) and anti-fatigue characteristics in pressure sensors. Additionally, the heterogeneous integration of MXene with thermochromic materials or electromagnetic shielding materials is used to develop multifunctional smart electronic devices;92,93 for example, MXene-based thermochromic patterns can display electromagnetic wave distribution in real time (Table 2).26 In multi-material printing processes, the combination of inkjet printing and laser micromachining has proven effective in resolving interfacial compatibility issues between different materials, advancing the practical application of MXene in integrated wearable energy-sensing systems (Fig. 3).54,76–78
Table 2 Comparison of the advantages and disadvantages of MXene and traditional carbon materials (e.g., graphene, carbon black, and carbon nanotubes) in additive manufacturing
Comparison dimension Electrical conductivity Rheological controllability Film-forming ability Energy storage capacity Transistor mobility (cm2 V−1 s−1) Functionalization Advantages Ref.
MXene 104–105 S cm−1 Excellent dispersion Dense, stackable, flexible. >300 F g−1, 1–40 –O/–OH/–F/–Cl/–S; tunable WF High electrical conductivity; good dispersibility. 6, 20, 23, 24, 43, 59, 94 and 95
Graphene 103–104 S cm−1 Requires modification poor flow Voids/wrinkles reduce density. <150 F g−1 0.1–10 Tunable; functionalization may reduce conductivity and increase cost. Ultrahigh electrical conductivity and mechanical strength. 80, 96 and 97
CNTs 103–104 S cm−1 High aspect ratio; entangling Porous random nets <100 F g−1, 0.2–50 Tunable; generally preserves performance. High-aspect-ratio percolation networks. 79 and 98
Carbon black 100–102 S cm−1 Excellent rheological modifier Moderate film, high resistance ∼50–100 F g−1 Filler only Limited chemical functionalization; properties sensitive to surface oxidation/processing. Low cost, mature 99 and 100



image file: d5mh01421b-f3.tif
Fig. 3 Multi-material integration and functionalized design. (A) Schematic of the fabrication process of a transparent flexible SC device based on Ti3C2Tx/PEDOT:PSS composite electrodes. Reproduced from ref. 81 with permission from John Wiley and Sons, Copyright 2024. (B) Crystal structures of Ca2Nb3O10 and Ca2NaNb4O13. Reproduced from ref. 84 with permission from Elsevier, Copyright 2022. (C) Wet-spinning of MSP fibers. Schematic of the wet-spinning process of the MSP fibers and the molecular structure of PVA, alginate, and TEM images of MXene nanosheets. Forming mechanism of MSP fibers by intercalation and chelation. Optical images of MSP fibers collected using a roller. SEM image of a knotted MSP microfiber. Reproduced from ref. 36 with permission from John Wiley and Sons, Copyright 2025. (D) Schematic illustration of the preparation of MXene–RHGO and the formation of Ti–O–C bonding, and schematic of the diffusion, adsorption, and migration of Na+ at the heterointerfaces of the MXene and RHGO inside the 3D MXene–RHGO composites. Reproduced from ref. 37 with permission from John Wiley and Sons, Copyright 2024. (E) Schematic of a cutting-based kirigami film and a crack-based kirigami film and the crack evolution of h-PDMS/MXene film during stretching. Reproduced from ref. 91 with permission from American Chemical Society, Copyright 2021.

3. Rheology control strategies for MXene inks

This section maps the key rheological parameters of MXene inks to the operating windows of major additive manufacturing routes. In brief, extrusion/direct-ink-writing benefits from yield stress and thixotropy to ensure shape retention; spraying/coating (including screen/inkjet-like flows) favors low-viscosity, wetting-controlled inks for uniform films; and laser-assisted routes decouple flow from post-densification and surface functional control, where rheology mainly guarantees a smooth, defect-free preform. We therefore discuss how viscosity–shear curves, yield stress, and recovery behavior govern print fidelity and defect modes across these routes, and how ink design can be tuned accordingly.

3.1 Influence of rheological parameters on printing quality

The rheological behavior of MXene-based inks plays a decisive role in determining their printability, structural fidelity, and device performance. In this work, we explicitly link each rheological lever to the workable windows of extrusion, spraying/coating, and laser-assisted printing. Given the non-Newtonian nature of these colloidal dispersions, a comprehensive understanding and precise tuning of rheological parameters, including viscosity, shear-thinning index, yield stress, and thixotropy, are essential for optimizing additive manufacturing processes, such as direct ink writing (DIW), screen printing, and electrohydrodynamic printing.101–103 For extrusion/DIW, sufficient yield stress with rapid viscosity recovery stabilizes filaments and suppresses slumping and edge beading. For spraying/coating, a low-viscosity regime with appropriate wetting balances film uniformity against satellite-droplet and coffee-ring defects. For laser-assisted routes, rheology primarily ensures uniform precursor spreading; post-treatment controls densification, work-function tuning, and microstructure evolution.
3.1.1 Viscosity, shear-thinning behavior, and microstructure correlation. The rheological properties of MXene inks directly affect printing precision and microstructure formation.101–103 Research indicates that inks require both appropriate viscosity and shear-thinning behavior to meet the demands of additive manufacturing processes, such as direct ink writing (DIW).14,94,104–108 For example, MXene/xanthan gum hybrid inks enable dynamic viscosity control (in the 1–1000 Pa s range) by adjusting the ratio of MXene to the polysaccharide polymer while also exhibiting significant shear-thinning characteristics (viscosity decreases as shear rate increases). This ensures enhanced ink flowability during extrusion and rapid recovery of high viscosity after deposition to maintain structural stability.11,101,104,109 This shear-thinning behavior is closely linked to the orientation and alignment of MXene sheets under shear force. By controlling the shear rate, sheets can be induced to align directionally along the flow direction, forming an ordered layered microstructure. This subsequently enhances electrode ion transport efficiency (e.g., the rate capability of capacitors).19,110

Furthermore, the incorporation of xanthan gum not only optimizes rheological properties but also suppresses the self-stacking of MXene sheets via hydrogen bonding and the steric hindrance effects of its polymer chains, thereby forming a porous network structure.18,109,111 Similarly, the addition of graphene oxide (GO) can crosslink MXene sheets through π–π interactions, enhancing ink dispersion stability. Simultaneously, by modulating the ink's yield stress (>100 Pa), high-resolution printing can be achieved, avoiding the collapse of microstructures during the printing process.102,112 Practically, these parameters define the extrusion window (nozzle pressure, line width stability, and layer stacking) in DIW-type printing.

Having established how viscosity and shear-thinning shape filament stability and alignment, we next discuss the additive routes that place MXene inks within these printable windows.

3.1.2 Optimization of ink stability by additives. Additives, such as xanthan gum and GO, serve dual roles in MXene inks. First, as rheological modifiers, they enhance ink stability through polymer chain entanglement or inter-sheet interactions. Second, as functional components, they directly participate in constructing the electrode's microstructure.60 For instance, the shear-thinning properties of xanthan gum have been proven to improve the migration capability of Fe–Mn oxide particles in heterogeneous porous media. This mechanism is also applicable to MXene inks, where increasing the zero-shear viscosity at low shear rates (6.3 × 104 mPa s) suppresses MXene sedimentation, ensuring long-term storage stability of the ink.102,109,113 The introduction of GO enhances the ink's oxidation resistance by modulating the interlayer spacing of MXene sheets (confirmed by XRD analysis). Chemical bonding between its oxygen-containing functional groups and MXene surface terminations (e.g., –O and –OH) reduces oxidative side reactions, thereby extending the ink's service life.102,104,114 From a process viewpoint, such stability engineering places MXene inks in the low-viscosity, wetting-controlled window required for uniform spraying/coating.

Beyond empirical tuning, the following subsection summarizes constitutive models that quantify these windows and guide parameter selection.

3.1.3 Theoretical foundations of ink rheology: fluid mechanics models. The rheological behavior of MXene inks during extrusion printing is closely governed by fundamental fluid mechanics, particularly non-Newtonian flow models.115 Although Newtonian fluids exhibit a linear relationship between shear stress (τ) and shear rate (γ˙), expressed as τ = ηγ˙, MXene inks are more accurately described by applying generalized non-Newtonian models due to their complex solid–liquid phase interactions.116 A commonly used model is the Herschel–Bulkley equation,117 which captures both yield stress and shear-thinning behavior:
τ = τ0 + ˙n,
where τ0 is the yield stress (below which the ink behaves like an elastic solid), K is the consistency index, and n is the flow index. For MXene-based inks, typical values of n < 1 indicate pseudoplastic (shear-thinning) behavior, enabling easier extrusion under high shear while maintaining shape fidelity after deposition.118 When n = 1 and τ0 > 0, the model reduces to the Bingham plastic, which is commonly used for paste-like inks in direct ink writing (DIW).

Such models also explain the observed thixotropic behavior, where viscosity decreases under shear and gradually recovers after rest.119 This is essential for printing applications: during extrusion, a low viscosity ensures flowability; post-deposition, high zero-shear viscosity and yield stress prevent the printed filament from collapsing due to gravity or surface tension.120,121

Additionally, shear-induced alignment of MXene nanosheets can be understood through the torque balance between viscous drag, capillary forces, van der Waals interactions, and gravity.122 This physical framework underpins recent strategies that tune ink composition and shear rate to align MXene flakes along the flow direction, forming anisotropic layered architectures with enhanced ion and electron transport.123

In summary, integrating fluid mechanics into ink design offers a quantitative foundation for tailoring MXene ink flow properties and achieving print fidelity and structural control.124 Future development may further benefit from fitting experimental rheology curves to these models, enabling the predictive design of inks for complex 3D architectures.

These models also rationalize why DIW relies on yield-stress and recovery, why spraying/coating prioritizes low-viscosity wetting, and why laser-assisted routes shift control to post-densification. Although constitutive models rationalize steady and transient responses, practical ink design hinges on where a given formulation sits relative to these windows; we therefore outline functionalized/binder-free strategies that place MXene inks in the printable regime while preserving conductivity and stability.

3.2 Functionalized ink design

The development of functionalized MXene inks represents a pivotal strategy for overcoming the limitations associated with conventional formulations, such as poor mechanical resilience, oxidation susceptibility, and the need for insulating polymer binders. This section focuses on recent advances in binder-free ink development and composite ink systems that simultaneously enhance conductivity, mechanical flexibility, and printability.
3.2.1 Development of binder-free inks. Conventional MXene inks often rely on polymer binders (e.g., PVDF) to maintain structural integrity, but the insulating nature of binders significantly reduces electrode conductivity.125 Recently, binder-free ink designs have become a research focus,126 with the core strategy being structural self-support through MXene sheet self-assembly or surface functionalization. For example, oxyanion-terminated Ti3C2Tx MXene inks enhance inter-sheet interactions via electrostatic binding, forming stable gel networks. This enables high-precision printing without additional binders, achieving dried electrode conductivity of up to 916 S cm−1.41,104 Another approach utilizes interfacial interactions between MXene and natural polymers (e.g., chitosan) to form porous aerogel structures through freeze-drying. These exhibit a 20-fold increase in compressive modulus while maintaining high conductivity (>100 S cm−1).18,127
3.2.2 Synergistic enhancement of conductivity and mechanical flexibility. Co-optimizing MXene ink conductivity (∼104 S cm−1) and mechanical flexibility is critical for flexible electronics. Studies have shown that anisotropic microstructures can be engineered by controlling ink composition (e.g., MXene/GO composites) and printing parameters (e.g., extrusion pressure and layer stacking).112 Composite flexible transparent electrodes (MXAg) combining MXene with 1D silver nanowires (AgNWs) demonstrate synergistic improvements. MXene sheets act as a “solder” at AgNW network junctions, while their compliant structure disperses mechanical stress. This enhances bending stability and reduces inter-wire contact resistance, achieving sheet resistance as low as 13.9 Ω sq−1.21 For instance, MXene electrodes printed via micro-continuous liquid interface production (μCLIP) exhibit highly aligned in-plane sheets, showing 30% higher conductivity than disordered structures while retaining >95% capacitance after 1000 bending cycles.110,112 Additionally, incorporating cellulose nanofibrils (CNFs) strengthens inter-sheet bonding via hydrogen bonds, yielding fiber electrodes with 120 MPa tensile strength alongside high conductivity (916 S cm−1) and flexibility (bending radius < 1 mm).41,128
3.2.3 Innovative extension. Recent research explores dynamic crosslinking strategies, such as leveraging coordination between MXene surface terminations (–O and –F) and metal ions (Al3+). Post-printing ion infusion can in situ enhance inter-sheet connections, simultaneously boosting mechanical strength (elastic modulus >50 MPa) and electrochemically active surface area.72,129 Such functionalized ink designs provide new perspectives for MXene printed electrodes in wearable energy devices and flexible sensing applications (Fig. 4).112,130
image file: d5mh01421b-f4.tif
Fig. 4 Influence of rheological parameters on print quality. (A) Schematic illustrating that print-line alignment is governed by a balance of microscale forces during deposition: shear from the flowing ink (Fc1), gravity (Fg), viscous drag on the substrate (Fd), capillary/meniscus force (Fc2), and van der Waals attraction between adjacent printed layers (FvdW). Reproduced from ref. 110 with permission from American Chemical Society, Copyright 2021. (B) C12E9 imparts amphiphilicity to MXene flakes, inducing the formation of a 3D interconnected network structure with high viscosity properties, uniform distribution of flakes, and enhanced ink printing performance. Reproduced from ref. 20 with permission from American Chemical Society, Copyright 2022. (C) storage (G′) and loss (G′′) moduli as a function of oscillation strain for the MXene/C12E9 ink. Ink viscosity progression over time with alternate low (1 s−1) and high shear rates (100 s−1) presenting the appropriate viscosity drop and recovery during and after ink extrusion through a nozzle. Relationship between the shear stress and shear rate of MXene/C12E9 ink fitting the Herschel–Bulkley fluid model. Reproduced from ref. 20 with permission from American Chemical Society, Copyright 2022. (D) Schematic of the proposed EMI shielding mechanism of hierarchical porous architecture with a continuous conductive network. Reproduced from ref. 131 with permission from Elsevier, Copyright 2023. (E) Schematic of the fabrication of MXene fibers with tablet interlocks between the outer COC layer and inner MXene fibers at the interface layer directly through fluidics-assisted thermal drawing, and the preparation of woven textiles based on strong MXene fibers. Reproduced from ref. 128 with permission from John Wiley and Sons, Copyright 2023. (F) Stress–strain curves of CCM fibers at various draw-down ratios. Tensile strength and toughness of CCM fibers. Electrical conductivity of CCM fibers. Reproduced from ref. 128 with permission from John Wiley and Sons, Copyright 2023.

4. Electrode structure design and performance optimization

4.1 Microstructural innovation

The microstructural design of MXene-based printed electrodes plays a critical role in determining ion/electron transport dynamics, electrochemical accessibility, and long-term mechanical stability. Additive manufacturing enables the deliberate construction of 3D architectures and anisotropic alignments, offering opportunities to overcome the intrinsic limitations of MXene, such as restacking and in-plane transport constraints.
4.1.1 3D porous architecture promotes electrolyte permeation and ion diffusion. Designing 3D porous MXene electrode structures via additive manufacturing techniques (e.g., 3D printing) significantly enhances electrolyte wetting efficiency and shortens ion diffusion paths.112 For instance, MXene/graphene oxide aerogels (SMGAs) fabricated using direct ink writing (DIW) technology demonstrate tunable hollow-core structures.6,20 Their porous nature not only reduces electrode density but also enhances ion transport kinetics through vertically aligned micro- and sub-micropores. At the microstructural level, this technique enables layer-by-layer deposition of functional inks into 3D configurations at ambient temperature, constructing conductive scaffolds with periodic features, tunable elastic moduli, or specific topologies.18,43 This effectively improves electrode charge transport efficiency and structural stability. Furthermore, hybrid fibers combining cellulose nanofibrils (CNF) with MXene, formed using the ice-templating method, create directional pore channels. This effectively suppresses MXene sheet restacking, thereby boosting ion diffusion rates and capacitive performance.41,132
4.1.2 Vertically aligned MXene sheets reduce ion transport resistance. The vertical (edge-on) orientation of MXene sheets during printing markedly reduces through-plane ion-transport resistance. Studies have shown that, by tuning ink viscoelasticity to promote shear-/capillary-driven edge-on alignment of MXene sheets, through-thickness, percolating ion-transport channels can be formed.133 For example, in MXene electrodes fabricated via capillary-force-driven DIW, the sheets preferentially align along the printing direction; when combined with surface microchannel confinement, this configuration enables rapid ion diffusion and electronic conductivity up to 916 S cm−1.20,110,134 his alignment-control strategy provides a practical paradigm for designing high-power-density energy-storage devices (Fig. 5).
image file: d5mh01421b-f5.tif
Fig. 5 Innovative MXene microstructure design and multiscale performance correlation. (A) Schematic of low tortuosity and sufficient ion migrations for the 3D Na-MXene electrode. Reproduced from ref. 11 with permission from American Chemical Society, Copyright 2022. (B) Schematic of the wet-spinning process for the preparation of CNF/MXene hybrid fibers. Reproduced from ref. 41 with permission from John Wiley and Sons, Copyright 2023. (C) Schematic and digital images of the fabrication process for the QSSC device. Reproduced from ref. 132 with permission from John Wiley and Sons, Copyright 2023. (D) A photograph showing the hybrid 3D printing approach with the DIW of droplets on the μCLIP-printed surface patterns. (E)–(H) Optical images for the patterned surface of linear grooves (E) before and (F) after droplet deposition, with (H) a zoomed-in image showing homogeneous MXene distribution. (I) Optical images of 3D-printed, complex surface patterning consisting of nonlinear patterns (from left to right: wavy grooves, antenna-shaped circles, cross-linked triangle channels, and an interdigitated structure), and simulated nanoparticle position field within microchannels showing capillary-driven MXene transport. The scale bar represents the particle density distribution (mg m−3). (Scale bars are 2 mm in (E) and (F), 300 μm in (G), and 1 mm in (H)). Reproduced from ref. 110 with permission from American Chemical Society, Copyright 2021.

4.2 Interface engineering and surface modification

Surface and interfacial engineering strategies are essential for enhancing the electrochemical stability, charge transfer efficiency, and environmental durability of MXene-based printed electrodes. Due to the high surface reactivity of MXene nanosheets, especially at undercoordinated transition metal sites, interfacial deterioration such as oxidation, delamination, and structural collapse can significantly impair device performance. This section discusses surface functional group modulation and heterointerface construction as key approaches for achieving chemically robust and electrochemically efficient MXene electrodes.
4.2.1 Influence of surface functional group modulation on electrochemical performance. The chemical states of MXene surface functional groups (e.g., –O, –OH, and –F) directly impact its electrochemical activity and stability.21,33 Studies have revealed that environmental factors like water and oxygen in solutions react with undercoordinated Ti atoms on MXene surfaces, generating degradation products, such as TiO2 and CH4, thereby significantly compromising electrical conductivity and structural integrity.114 Oxygen anion termination modification (e.g., PO43− and SO42−) enhances electrostatic interactions between MXene layers, enabling stable and rheology-tunable inks.43 For instance, oxygen anion-terminated Ti3C2Tx MXene ink exhibits exceptional oxidation resistance without additives, and printed paper-based supercapacitors retain >90% capacity after 10[thin space (1/6-em)]000 bending cycles.25 Additionally, boron (B) doping optimizes the electronic structure of Pd–B/Pd heterometallenes, boosting oxygen reduction reaction activity and offering a new paradigm for MXene surface modification.135
4.2.2 Heterointerface optimization suppresses MXene oxidation. Constructing heterointerfaces between MXene and polymers/carbon materials effectively enhances environmental stability.136 Recently, a metal nano-armor strategy based on seamless heterointerface design has offered a novel approach: sputtering a nanoscale copper layer (≈141 nm) onto MXene films forms a uniform, dense MXene@Cu heterostructure. This interface physically blocks O2/H2O penetration into MXene interlayers, inhibiting oxidation. Experiments have shown that MXene@Cu films maintain 72.0% conductivity after 30-day air exposure, far exceeding pure MXene films (44.3%).137 Meanwhile, molecular-level hybridization of methylcellulose (MC) with MXene significantly suppresses oxidation in aqueous electrolytes by expanding interlayer spacing and strengthening hydrogen bonding while improving wettability and mechanical strength.138 Furthermore, 3D artificial array interface engineering (e.g., MXene/Zn heterostructures) enables dendrite-free metal anodes by regulating zinc deposition kinetics, providing guidance for designing high-stability MXene-based electrodes.14,49,134,139

4.3 Pathways to high-performance electrodes

The realization of high-performance MXene-based electrodes via additive manufacturing relies on the synergistic optimization of ink formulation, structural engineering, and post-processing treatments. This section discusses key strategies for translating material and printing innovations into practical device performance, with a focus on achieving high conductivity, mechanical compliance, and long-term operational stability.
4.3.1 Pre-inserted ion technology enhances cycling stability. Pre-inserting ions (e.g., K+ and Na+) expands MXene interlayer spacing and improves potassium/sodium storage performance. For instance, potassium-preinserted Ti3C2Tx MXene exhibits a specific capacitance of up to 220 F g−1 and exceptional cycling stability (>85% retention after 5000 cycles) in potassium-ion capacitors through solvation structure optimization.55 Similarly, chemical sodium pre-insertion significantly enhances sodium-ion diffusion coefficients and electrode reversibility by strengthening electrostatic shielding effects between MXene layers.140,141 In supercapacitors, the performance improvement of electrophoretically deposited (EPD) composite electrodes stems from electric field-driven 3D porous structure construction and full exposure of active sites.7 Furthermore, EPD technology eliminates the need for charge inducers, enables scalable control of film thickness/porosity, and synergizes with ion pre-insertion (e.g., K+/Na+-expanded interlayers). These strategies collectively enhance cycling stability and ion storage performance through structural optimization, offering efficient and scalable pathways for high-performance energy storage electrode fabrication.142
4.3.2 Asymmetric design and all-MXene integrated systems. Asymmetric designs (e.g., heterostructures and gradient porosity) overcome the energy density limitations of symmetric MXene devices. For example, DIW-printed “micro-cup” structured supercapacitors achieve a volumetric capacitance of 38.6 F cm−3 and reduced time constant (0.3 ms) via vertical interfacial and sandwich geometric design.26 Additionally, all-MXene integrated systems (e.g., synergistic design of MXene-based current collectors and active layers) enable thickness-independent capacitive behavior and rapid charge transport in high-mass-loading electrodes through horizontal/vertical sheet orientation combinations.77,143,144 These approaches provide critical references for developing next-generation high-performance miniaturized energy storage devices (Fig. 6).49,61 Section 4 focuses on printed electrodes their structure, densification routes, sheet and contact resistance, interfacial adhesion, patterning resolution/accuracy, and operational stability whereas Section 5 presents the device-level functions and system integration.
image file: d5mh01421b-f6.tif
Fig. 6 Multifunctionalization of MXene driven by interfacial engineering and surface modification strategies. (A) Viscoelasticity-adjustable p-MXene ink and its interaction mechanisms. The inset contrasts the viscous p-MXene ink and liquid MXene ink at a solid content of 40 mg mL−1. Good adaptability of p-MXene inks for screen-printing EMI shielding and IR anticounterfeiting coatings. Reproduced from ref. 19 with permission from American Chemical Society, Copyright 2022. (B) Schematic showing the in situ antioxidative shielding of oxyanions, the electrostatic interactions among MXene sheets, and the paper supercapacitors with customizable capacity, flexibility, and shape. Reproduced from ref. 25 with permission from John Wiley and Sons, Copyright 2024. (C) Schematic of the synthesis of the Pd–B/Pd HMs. Schematic illustrating the role of Pd–B/Pd HMs when catalyzing the alkaline ORR. Reproduced from ref. 135 with permission from John Wiley and Sons, Copyright 2023. (D) Fabrication of MC/MX hybrid films and the interface interaction. Reproduced from ref. 138 with permission from John Wiley and Sons, Copyright 2022. (E) Mechanisms of MXene array interfaces regulating the growth kinetics and deposition behavior of Zn atoms revealed at the multiscale level. Reproduced from ref. 134 with permission from Springer Nature, Copyright 2023. (F) Illustration of a 3D K+-Ti3C2Tx//AC PIHC device. CV curves of the as-prepared 3D K+-Ti3C2Tx//AC at different scan rates (0.5 to 50 mV s−1). Rate capability of the 3D K+-Ti3C2Tx//AC from 0.05 to 10 A g−1. Ragone plots of the 3D K+-Ti3C2Tx//AC compared with previously published works. Reproduced from ref. 55 with permission from John Wiley and Sons, Copyright 2021.

5. Device performance optimization and emerging applications

This section consolidates the device-level implications of MXene electrodes, including WF/SBH engineering, transport behaviors, circuit metrics, and emerging applications.

5.1 Breakthroughs in energy storage device performance

5.1.1 Supercapacitors: high volumetric capacitance and rapid charge/discharge. MXene printed electrodes significantly enhance supercapacitor volumetric capacitance, charge/discharge rates, and stability (e.g., oxidation-resistant passivation layers) through 3D microstructure design and surface chemistry modulation.76,83 Fiber-like, film-like, and 3D porous MXene-based supercapacitors achieve high flexibility, specific capacitance, and mechanical stability, meeting diverse flexible energy storage demands.145–147 For instance, 3D-printed MXene electrodes with optimized pore structures (5–50 μm diameter) facilitate electrolyte penetration and ion diffusion, delivering a high specific capacitance of 381 F g−1 (at 2 mV s−1) and 95% capacity retention after 10[thin space (1/6-em)]000 cycles.11 Modulating oxygen-containing functional groups on MXene surfaces enhances electrode–electrolyte interfacial charge transfer, boosting an energy density to 15.3 Wh kg−1 at 5 A g−1.12,148

Porous architecture design, conductive network construction, and binder-free strategies further optimize MXene electrodes. Electrophoretic deposition (EPD) leverages a “size-sieving” effect, enabling preferential vertical alignment of large MXene sheets (interlayer spacing expands from 1.358 to 1.375 nm).95 This dramatically enhances ion accessibility and diffusion kinetics, achieving 468 F g−1 gravimetric capacitance and 910 F cm−3 volumetric capacitance.7 Co-depositing MXene with CNTs or rGO constructs 3D conductive networks (19[thin space (1/6-em)]262 S m−1 electronic conductivity), yielding 470 mF cm−2 areal capacitance and <0.5 s relaxation time constant, approaching ideal electric double-layer behavior. Template matching MXene flake dimensions (e.g., 310 nm) prevents restacking and forms uniform interconnected pores, boosting electrochemical performance (474 F g−1 specific capacitance, 98.6% cycling stability). Integrating CNTs enhances conductivity and mechanical strength (107 MPa), enabling flexible supercapacitors with high-rate capability, low impedance, and 10.82 Wh kg−1 energy density.149

5.1.2 Zinc-based batteries: suppressing HER and boosting energy density. MXene electrodes inhibit hydrogen evolution reactions (HER) via interface engineering in zinc-based batteries150 while enhancing energy density. MXene-based HER catalysts also improve performance through surface terminal group modulation151 (e.g., –O termination for optimized H* adsorption)

• Single-atom decoration (e.g., Pt/Ru SACs strengthening metal–support coupling)

• Defect engineering (e.g., Ti/Mo vacancies exposing active sites)

• Multidimensional heterostructures (e.g., MoS2/Nb2C and LDH/MXene).152,153

MXene/defect-engineered carbon fiber composites reduce localized current density on zinc anodes, suppressing dendrite growth and enabling 91% capacity retention after 500 cycles at 1 A g−1.154 Dopamine intercalation and carbonization generate N-doped graphene-like carbon (NGC) within Ti3C2Tx interlayers, forming NGC–Ti3C2Tx heterostructures that significantly enhance ion storage capacity.155,156 SnS nanoparticles anchored on MXene via covalent bonding accelerate charge transfer kinetics, delivering 520 mAh g−1 reversible capacity at 0.5 A g−1 in flexible sodium-ion battery anodes.89 MXene's metallic conductivity and abundant active sites further enable bifunctional ORR/OER in Zn–air batteries, achieving 980 Wh kg−1 energy density.15,157,158

Flexible MXene (Ex-Ti3C2Tx) films with interlayer-expanded (15.04 Å via MgO templating) and defect-enriched structures (Ti3+/Ti2+ sites and Ti–O bridges) serve as cathodes in Li–CO2 batteries. These uniquely stabilize the two-electron product Li2C2O4, preventing its disproportionation to Li2CO3 and significantly enhancing CO2 reduction/evolution reaction (CRR/CER) reversibility and electrochemical performance (Fig. 7).159


image file: d5mh01421b-f7.tif
Fig. 7 Multi-dimensional synergistic mechanism and cross-scale validation of performance breakthroughs in MXene-based energy storage devices. (A) Schematic illustration of the two-electrode Swagelok configuration used for the measurement. Reproduced from ref. 11 with permission from American Chemical Society, Copyright 2022. (B) The areal capacitance of pristine MXene, Na-MXene, and 3D Na-MXene was measured at different current densities (mass loading of 2.5 mg cm−2). Reproduced from ref. 11 with permission from American Chemical Society, Copyright 2022. (C) Differentiation of capacitance contribution from CV curves of non- and 3D Na-MXene at a scanning rate of 5 mV s−1 to calculate both fast kinetic and slow kinetic processes. Reproduced from ref. 11 with permission from American Chemical Society, Copyright 2022. (D) Longterm stability of 3D Na-MXene at a current density of 10 mA cm−2 for 10[thin space (1/6-em)]000 cycles; inset shows the constant current charge/discharge curves of the initial six cycles and last six cycles. Reproduced from ref. 11 with permission from American Chemical Society, Copyright 2022. (E) Illustration of the Swagelok three-electrode setup for evaluating the electrochemical performance of the MXene-based film electrodes. Specific gravimetric capacitances are calculated from the GCD curves. Reproduced from ref. 12 with permission from American Chemical Society, Copyright 2024. (F) Schematic illustration of the fabrication of MXene/SnS2@NCFs. Comparison of the rate capability of MXene/SnS2@NCFs with other reported values. Reproduced from ref. 89 with permission from John Wiley and Sons, Copyright 2023. (G) A schematic illustration of a self-assembled MXene layer on the Zn surface and its functions. Reproduced from ref. 160 with permission from John Wiley and Sons, Copyright 2020. (H) A schematic illustration of Zn plating on Zn surfaces with or without an MXene–mPPy coating layer. Reproduced from ref. 156 with permission from John Wiley and Sons, Copyright 2022. (I) The formation of highly lattice-matched Zn deposits on Ti3C2Cl2. Reproduced from ref. 15 with permission from American Chemical Society, Copyright 2022.

5.2 Flexible electronics and intelligent sensing

5.2.1 Mechanical stability design for wearable devices. Mechanical stability of MXene-printed electrodes has been advanced through hierarchical structural design.21 For example, carbon nanotube-MXene double-network sponge (DNS) electrodes retain 98% electrical conductivity under 80% compressive strain, with >90% capacitance retention after 1000 compression cycles.161 Flexible films combining MXene with N-doped carbon fibers exhibit exceptional fatigue resistance, showing no cracking or delamination after repeated bending (3 mm curvature radius).86 MXene-based hydrogel electrodes achieve self-healing capability (>95% recovery efficiency) via dynamic cross-linked networks, enabling deployment in complex deformation scenarios.162 Bioinspired designs (e.g., serpentine electrodes) and geometric optimization (e.g., arc angle α = 270°) significantly enhance mechanical stability in wearable energy storage devices. These devices achieve 61% elastic tensile strain while maintaining 95% capacity under extreme deformation (2000% compressive cycling).163 Concurrently, nitrogen-doped freestanding electrodes (Py-Ti3C2) prepared via pyridine-assisted solvothermal methods demonstrate superior flexibility and electrochemical performance, further elevating device capabilities.164
5.2.2 Self-powered integrated systems (energy harvesting + wireless communication). MXene-based self-powered systems enable 24/7 operation by integrating energy storage and harvesting modules. For instance, an MXene/perovskite solar cell-zinc-ion battery wristband continuously powers wireless temperature sensors for 8 hours under indoor light (200 lux).154,165,166 MXene triboelectric nanogenerators (TENGs)167,168 treated with surface plasma achieve enhanced charge density (230 μC m−2). When coupled with micro-supercapacitors, they unify mechanical-to-electrical energy conversion and storage, supporting bluetooth low energy (BLE) transmission (>10 m range).169,170 Additionally, integrated systems combining MXene piezoresistive sensors with flexible Li–S batteries enable real-time monitoring of human motion signals with wireless data transmission, demonstrating potential for smart healthcare monitoring.77,171,172

MXene composites leverage exceptional conductivity (e.g., 6900 S cm−1 for Ti3C2Tx) and flexibility, combined with 3D/4D printing structural customization capabilities, to revolutionize flexible electronics and intelligent sensing.173 Printed MXene micro-supercapacitors (MSCs), fabricated using techniques like 3D freeze-printing, integrate seamlessly with flexible solar cells and sensors, enabling self-powered flexible systems for wearable applications.61 Their performance in high-sensitivity sensors is equally remarkable:

• Pressure sensors (6.03 kPa−1 sensitivity and 9 Pa detection limit)

• Wireless plant ethylene sensors (0.084 ppm detection limit)

• Photothermal responsive devices (98% shape recovery in 14 s)

Despite challenges like oxidation resistance and ink dispersion, advances in MXene contact engineering and printing processes are establishing robust material foundations and scalable manufacturing pathways for next-generation flexible intelligent electronics (Fig. 8).5,49


image file: d5mh01421b-f8.tif
Fig. 8 Multifunctional integration of MXene-based flexible electronics and smart sensing. (A) Schematic of different types of 3D compressible electrodes. Reproduced from ref. 161 with permission from John Wiley and Sons, Copyright 2021. (B) Long-cycle performance of DNS under a set strain of 50% (inset is the image of the C-LIB prototype under compression to 50% strain). Capacity retention of DNS under cyclic strain in an in situ measurement (compressing the C-LIB repeatedly for 1000 cycles while simultaneously recording the charge/discharge curves). Reproduced from ref. 161 with permission from John Wiley and Sons, Copyright 2021. (C) Schematic of the integrated multifunctional flexible sensor-supercapacitor device. Reproduced from ref. 162 with permission from John Wiley and Sons, Copyright 2023. (D) Schematic of the detailed fabrication processes of RPH, FPH, and HPH. Reproduced from ref. 162 with permission from John Wiley and Sons, Copyright 2023. (E) Digital photographs demonstrating the integration process of the self-powered smart bracelet. JV curves of the prepared flexible PSC under sunlight (up) and room light (below). Reproduced from ref. 154 with permission from American Chemical Society, Copyright 2021. (F) Schematic of the synthesis of LIG/Ag and the fabrication process of LIG/Ag-based FSCs. Reproduced from ref. 169 with permission from John Wiley and Sons, Copyright 2025. (G) Optical image of the planar all-in-one glucose-detection system. Response of the glucose-detection system used for the continuous monitoring of glucose solutions with different concentrations. Reproduced from ref. 169 with permission from John Wiley and Sons, Copyright 2025. (H) The overall design of the fabric-based smart mask for respiration monitoring. Reproduced from ref. 56 with permission from Elsevier, Copyright 2022.

5.3 Printed MXene electrodes for high-mobility transistors

MXene electrodes (such as Ti3C2Tx) demonstrate significant advantages when used in field-effect transistors (FETs),174 including high conductivity, tunable work function, compatibility with low-temperature solution processing, low contact resistance, bipolar charge transport characteristics, and environmental stability. These properties make them ideal electrode materials for FETs based on organic, oxide, and two-dimensional (2D) semiconductors.175 The following provides a detailed analysis of key aspects, citing relevant literature to support the arguments.
5.3.1 Electrical properties and printing methods of printed MXene electrodes. MXene electrodes achieve high conductivity (typically superior to conventional metal electrodes) through low-temperature solution processing, which is crucial for various printing methods.176 For example, AD-MXene (a modified MXene) exhibits excellent dispersion stability in ethanol, making it suitable for electrohydrodynamic printing while maintaining high conductivity. These electrodes outperform vacuum-deposited gold (Au) and aluminum (Al) electrodes in thin-film transistors, simultaneously offering good environmental stability (attributed to their hydrophobicity).62 Relevant literature indicates that MXene inks can be integrated using methods such as direct writing and screen printing, enabling low-cost and scalable manufacturing.173 As 2D materials, MXenes possess negative zeta potential, high carrier mobility, and tunable surface chemistry. These properties facilitate their widespread use as conductive electrodes in printed electronics, including electronic, optoelectronic devices, and sensors.5,7,177–179 Furthermore, tunable work function, achieved via surface terminations, SAMs, plasma/UV-ozone, and dopants, is the core advantage of MXene electrodes (see Table 3). Through surface functionalization or metal doping (e.g., gold deposition), the work function can be optimized to match semiconductor energy levels, thereby reducing contact resistance and the Schottky barrier.180 For example, work function modulation enables near-zero Schottky barrier heights (∼14 meV) and low contact resistance (∼0.7 kΩ μm), significantly enhancing device performance.181,182
Table 3 Effects of different strategies on work function modulation of MXene electrodes
Strategy category Specific method Work function modulation effect Effect on contact resistance/Schottky barrier Ref.
Metal deposition/doping Au deposition Optimize WF matching Interface optimization 180–182
Surface functionalization/oxidation Surface modification Modulate surface WF 12-fold Rc reduction; ΦB ∼ 44 meV 183
Hetero-doping/compositing PtTe2/WO3 doping Energy tuning Narrowed barrier width 184 and 185
Plasma treatment O2 plasma treatment Enhance WF Optimize interface energy-level alignment 193 and 194
Terminal group modulation –O, –OH, –F surface groups W F tunability: 2–6 eV Defect reduction. 181, 186–190 and 195
Solvent engineering/chemical treatment TFA modification Conductive dispersion Indirectly stabilize WF 94 and 196


5.3.2 Contact resistance and Schottky barrier suppression. Printed MXene electrodes significantly reduce contact resistance and the Schottky barrier through interface engineering, thereby enhancing charge injection efficiency. The examples referenced in the query align with the literature: molecular functionalization methods (e.g., surface oxidation or doping) can reduce contact resistance by up to 12-fold and lower the Schottky barrier to ∼44 meV, enabling high-efficiency charge injection.183 In 2D semiconductor (e.g., MoTe2 or WS2) FETs, MXene electrodes integrated with low-work-function materials (such as PtTe2 or WO3 doping) narrow the Schottky barrier width, thereby improving carrier injection.184,185 Specifically, the reduction in contact resistance correlates with work function matching, which is enabled by MXene's surface tunability. Through doping or functionalization, interface trap states are decreased and energy level alignment is optimized, thereby alleviating injection bottlenecks (Fig. 9).181,186–190 The literature further demonstrates that design rules, such as interface charge transfer, can overcome injection barrier limitations, achieving high carrier mobility (e.g., ∼40 cm2 V−1 s−1 in Ti3C2Tx-MoS2 transistors).191,192
image file: d5mh01421b-f9.tif
Fig. 9 Contact engineering of FET and printing techniques for MXene-based printed electrodes. (A) Schematic of the general reaction procedure for preparing AD-MXenes. Reproduced from ref. 62 with permission from Springer Nature, Copyright 2022. (B) Schematic showing the EHD printing of AD-MXene inks from a nozzle. The inset shows an optical microscopy image of AD-MXene ink drops at the edge of the nozzle tip during the cone-jet mode of the EHD printing process. Chart showing the EHD printing modes as a function of working distance and applied voltage. Reproduced from ref. 62 with permission from Springer Nature, Copyright 2022. (C) Schematic showing the abnormal grain growth of the 2H crystal from the seed. Reproduced from ref. 181 with permission from Springer Nature, Copyright 2023. (D) TLM plots of the 1T′/2H-MoTe2 junction FET with different Vg showing the linear dependence of the total resistance (RW) on L. The error bars result from the averaging of at least five different TLMs. Inset represents an optical microscopy image of TLM patterns for the 2H-MoTe2 FETs with vdW Au/1T contact electrodes (scale bar: 20 μm). Thermionic barrier height (ΦB) values measured at various Vg for hole transfer with 3DPt (blue), vdW Ag/1T′-MoTe2 (purple), and vdW Au/1T′-MoTe2 contacts (red) in 2H-MoTe2 FET. Reproduced from ref. 181 with permission from Springer Nature, Copyright 2023. (E) Schematic of the formation of the Ohmic contact between Ti3C2Tx MXenes and s-CNTs. Reproduced from ref. 182 with permission from the Royal Society of Chemistry, Copyright 2025. (F) Schematic of the MoS2 FET with N-benzylmaleimide functionalized at the contact region (Nbm-ct-MoS2 FET). Reproduced from ref. 183 with permission from American Chemical Society, Copyright 2022. (G) Normalized resistance of three-layer Nbm-ct-MoS2 FETs at Vg = 60 V along with the channel lengths. Reproduced from ref. 183 with permission from American Chemical Society, Copyright 2022. (H) Schottky barrier of monolayer Nbm-ct-MoS2 FETs. Reproduced from ref. 183 with permission from American Chemical Society, Copyright 2022. (I) Schematic of the fabrication process of a PtTe2-contacted MoTe2 FET on a 285 nm SiO2/n++ Si substrate. Cross-sectional HRTEM image (top) and schematic (bottom) of PtTe2-contacted MoTe2 transistors. Reproduced from ref. 186 with permission from American Chemical Society, Copyright 2024. (J) Illustration of the trilayer WS2 device after ozone oxidation. Two kinds of hole injection mechanisms are shown: (i) SB width control and (ii) SB height control. Reproduced from ref. 187 with permission from American Chemical Society, Copyright 2023.
5.3.3 Field-effect mobility and ambipolar transport. Printed MXene electrodes exhibit outstanding performance in high-mobility transistors, particularly by facilitating ambipolar transport (simultaneously supporting electron and hole injection).182,192 In organic FETs (OFETs),193 MXene electrodes functionalized through methods such as oxygen plasma treatment enhance electrode surface energy and work function. This enables balanced n-type and p-type mobilities (approaching 1 cm2 V−1 s−1) and high on/off ratios (>107).194 Consistent with the literature, OFETs demonstrate ambipolar charge transport characteristics where MXene electrodes optimize energy-level alignment and reduce Schottky barriers, thereby supporting ambipolar behavior. Furthermore, work function modulation (e.g., increasing by 0.7 eV) ensures efficient charge injection while avoiding non-ideal transport behavior.197–200 In 2D semiconductor FETs, MXene electrode integration (e.g., with graphene contacts) enables ambipolar carrier injection control, boosting mobility to 80 cm2 V−1 s−1 while reducing the Schottky barrier.96,201,202 These enhancements are attributed to MXene's charge injection efficiency, as well as optimized carrier transport through microstructure design (e.g., patterned electrode arrays).182,192
5.3.4 Environmental stability and surface functionalization. The environmental stability of MXene electrodes is significantly enhanced through surface functionalization, supporting long-term flexible electronics applications.203,204 The literature demonstrates that hydrophobicity and proton-accepting functional groups enable devices to maintain stable electrical characteristics after storage. For instance, Ti3C2Tx MXene achieves improved environmental stability and gas sensing performance via functionalization-controlled thickness and termination groups;195 these devices exhibit long-term stability at 60% humidity (e.g., 97% capacitance retention after 50[thin space (1/6-em)]000 cycles, which is applicable by analogy to transistor contexts). Robust interface barriers in MXene electrodes (e.g., MXene/GaN Schottky diodes) further ensure physicochemical stability, with tunable Schottky barrier heights reaching elevated values.62,205,206 Furthermore, patterning MXene on flexible substrates enables high integration density and conformability, while its matched work function reduces defects, making it suitable for long-term operation (Fig. 10).182,207
image file: d5mh01421b-f10.tif
Fig. 10 Work function tuning mechanisms and surface functionalization for organic field-effect transistors (OFETs). (A) Schematic of the electrodes prepared using the oxygen plasma method. Reproduced from ref. 200 with permission from American Chemical Society, Copyright 2024. (B) Energy level diagram for Au/C8-BTBT, in which SBH stands for Schottky barrier height. Reproduced from ref. 200 with permission from American Chemical Society, Copyright 2024. (C) Schematic of the MoTe2 FET with bottom graphene contacts. Reproduced from ref. 96 with permission from American Chemical Society, Copyright 2024. (D) Measured transfer characteristics of an ambipolar MoTe2 FET with bottom graphene contacts. Reproduced from ref. 96 with permission from American Chemical Society, Copyright 2024. (E) Band diagrams of the device under different gate bias conditions. Reproduced from ref. 96 with permission from American Chemical Society, Copyright 2024. (F) A novel water-free etching strategy for synthesizing few-layered Ti3C2Tx MXenes in deep eutectic solvents at near-ambient temperature. Reproduced from ref. 205 with permission from Elsevier, Copyright 2023. (G) Enhanced interlayer spacing of MXene nanosheets achieved via iodine terminals (calculated using Bragg's law) for optimized NO2 absorption. Reproduced from ref. 206 with permission from Springer Nature, Copyright 2024. (H) General scheme of the MXene/GaN van Hoof structure. Reproduced from ref. 207 with permission from American Chemical Society, Copyright 2024. (I) Band alignment diagram for all studied MXene/GaN van Hoof structures. Abbreviations: VL-vacuum level, CB-conduction band, V-valence band. Reproduced from ref. 207 with permission from American Chemical Society, Copyright 2024. (J) Calculation of the surface barrier height with the fitting curves for both reference GaN and different MXene/GaN structures. The determined values of the built-in electric field and surface barrier height are given in the legends. Reproduced from ref. 207 with permission from American Chemical Society, Copyright 2024.
5.3.5 Solvent engineering and printing processes for synergistic optimization of printed MXene electrodes. MXenes demonstrate significant potential as 2D conductive materials for printed electrodes yet face challenges in organic solvent dispersibility and patterning precision.196 Through trifluoroacetic acid (TFA) treatment, MXene (Ti3C2Tx) achieves stable dispersion in ethanol while maintaining ultrahigh conductivity.94 When integrated with electrohydrodynamic (EHD) printing, MXene inks form electrode patterns with micron-scale resolution (line width ≈180 μm), which are successfully implemented as gate/source–drain electrodes in fully printed transistors. Transistors fabricated with these electrodes exhibit a mobility of 1.38 cm2 V−1 s−1 in p-type semiconducting carbon nanotube (sCNT) channels. Their integration into NAND/NOR logic gates validates MXene electrodes for high-mobility organic integrated circuits.79 In contrast, graphene–carbon nanotube composite inks optimize dispersibility via centrifugal mixing, utilizing screen printing for electrode patterning (line width ≈50 μm). Applied as source–drain electrodes in organic field-effect transistors (OFETs),208 these achieve p-type/n-type mobilities of 0.25/0.13 cm2 V−1 s−1, respectively, further supporting the viability of carbon-based materials in printed integrated circuits (Fig. 11).80
image file: d5mh01421b-f11.tif
Fig. 11 Printed electrodes enabling the patterning and fabrication of logic circuits. (A) Fabrication flow chart of photo-patterned P(CEA-co-HDDA) dielectric film by the spin coating process. Reproduced from ref. 196 with permission from the Royal Society of Chemistry, Copyright 2024. (B) Schematic showing the EHD printing of TFA-MX ink from a nozzle. Reproduced from ref. 94 with permission from John Wiley and Sons, Copyright 2021. (C) Schematics of the bottom-gate staggered TFT and top-gate staggered TFT prepared with EHD-printed TFA-MX electrodes, sCNT active layer, and FPVDF–HFP dielectric layers. Reproduced from ref. 94 with permission from John Wiley and Sons, Copyright 2021. (D) VTC plots and corresponding signal gains of the bottom-gate staggered TFT and the top-gate staggered TFT. Reproduced from ref. 94 with permission from John Wiley and Sons, Copyright 2021. (E) Schematic, OM image (top), and the corresponding circuit diagram (bottom) of the organic complementary inverters prepared with screen-printed Ag/CNT electrodes. Reproduced from ref. 79 with permission from American Chemical Society, Copyright 2022. (F) VTCs and voltage gain plots of the complementary inverter. Reproduced from ref. 79 with permission from American Chemical Society, Copyright 2022. (G) Schematic and OM image of organic complementary inverters prepared using screen-printed graphene–CNT electrodes. Reproduced from ref. 80 with permission from Elsevier, Copyright 2022. (H) OM images of NAND and NOR gates prepared with screen-printed graphene–CNT electrodes. Reproduced from ref. 80 with permission from Elsevier, Copyright 2022. (I) Output voltages of NAND and NOR gates for various input combinations: (0,0), (1,0), (1,1), and (0,1). The bottom of the OM images in (G) and (H) shows the corresponding circuit diagram of each device. Reproduced from ref. 80 with permission from Elsevier, Copyright 2022.

6. Challenges and future directions

6.1 Current technological bottlenecks

6.1.1 Scalability: ink stability and printing precision. The rheological properties of MXene inks (such as viscosity and shear-thinning behavior) directly impact the uniformity and resolution of printed electrodes.63 Existing studies indicate that the agglomeration of MXene nanosheets within the ink significantly reduces printing precision. This is particularly challenging for the additive manufacturing of complex structures (e.g., microchannels and porous frameworks), where maintaining the ink's dynamic stability during continuous printing is difficult.11,79,80 Furthermore, conventional extrusion-based 3D printing is constrained by the stringent requirement to match specific rheological parameters (e.g., yield stress and thixotropy).209 This limitation restricts the types of electrode architectures that can be designed and hinders the scalable production of high-resolution structures and structures with high aspect ratios.11,49
6.1.2 Oxidative degradation during long-term cycling. MXene electrodes are susceptible to oxidation during electrochemical cycling or environmental exposure, resulting in the degradation of their conductivity and capacitive performance.148,210 Experimental evidence indicates that unencapsulated MXene films gradually form oxides like TiO2 in oxygen-containing environments (e.g., aqueous electrolytes or humid air). This oxidation diminishes surface active sites and increases charge transfer resistance.44,127,211 This process is especially pronounced in thick electrodes (>10 μm) due to the extended internal ion diffusion pathways, where localized oxidation exacerbates structural collapse.212,213 Although polymer encapsulation (e.g., using polydopamine) can delay oxidation, the encapsulating layer may compromise MXene's inherent high conductivity and interfacial ion transport efficiency.24,213

6.2 Cutting-edge research directions

6.2.1 Machine learning-assisted optimization of rheological parameters. Machine learning-based multi-objective optimization algorithms offer novel approaches for matching ink rheology with printing parameters.214 For instance, Gaussian process regression models can rapidly predict the nonlinear relationship between MXene ink viscosity and shear rate. This guides the precise doping of additives (e.g., cellulose nanofibers) to tailor yield stress.11,49 Furthermore, deep learning-driven image recognition techniques enable real-time monitoring of inter-layer defects (e.g., cracks or voids) during printing. Integrating this with closed-loop feedback systems allows for the dynamic adjustment of printing parameters (e.g., nozzle speed or pressure), significantly improving printing consistency.215
6.2.2 Biomimetic structure design and multi-scale simulation. Inspired by biological structures (e.g., plant vasculature or insect compound eyes), biomimetic hierarchical channel designs can optimize ion transport kinetics within MXene electrodes.216 For example, 3D Fe3O4/MXene aerogel electrodes, fabricated via freeze-drying combined with mechanical pressing, feature a hierarchical pore structure (micro–meso–macro pores) that shortens ion diffusion paths and enhances mechanical flexibility.42 Multi-scale simulations (e.g., coupling molecular dynamics with finite element analysis) further elucidate the relationship between MXene sheet stacking angles and electrolyte penetration depth. This provides theoretical guidance for the topological optimization of high-areal-capacity electrodes.42,212 MXene can simulate enzyme-substrate specific binding at the molecular level. Future interface designs can leverage hydrophilicity and biocompatibility to enhance bio-recognition capabilities. Simulating antioxidant enzyme mechanisms could improve MXene stability, while developing novel composite materials (e.g., MOF-MXene) will advance high-performance wearable and implantable biosensors toward clinical translation, offering new design paradigms for smart healthcare.217 Flexible neurointerfaces: printed MXene microelectrodes combine low impedance and high charge-injection capacity with conformal contact to soft tissue, enabling stable electrophysiological recording/stimulation on curved surfaces. Prototyping via DIW/screen printing demonstrates durable operation under bending and perspiration exposure.218,219 Soft robotics: MXene-based stretchable conductors maintain resistance stability at high strain and provide integrated sensing/heating for untethered soft actuators. Printed serpentine meshes on elastomers achieve repeatable actuation control and self-sensing in complex motions.220
6.2.3 Green manufacturing and recycling technologies. Green manufacturing technologies focused on sustainability are emerging research hotspots. For example, electrophoretic deposition (EPD) enables large-area fabrication of MXene films without binders, allowing for precise control over film thickness and porosity by adjusting voltage and deposition time.7,221 Additionally, strategies for recycling spent MXene electrodes have been proposed. Gentle acid treatment can delaminate MXene sheets from degraded electrodes, which can then be re-dispersed into functional inks. Regenerated electrodes using this approach can recover over 85% of their original specific capacitance.83,222 Water-triggered self-healing strategies also confer MXene freestanding electrodes with excellent renewability and environmental adaptability. This strategy significantly extends the service life of energy conversion devices, reduces maintenance costs and environmental burdens, and aligns with the trend toward environmentally friendly materials through its green repair process.223

From an industrial standpoint, MXene-printed electronics benefit from R2R-compatible inks with extended shelf-life and controlled oxidation, low-temperature densification (<120 °C) for polymeric substrates, reliability under humidity/thermal cycling with thin-film packaging, and standardized metrology (reporting sheet resistance at a specified dry-film thickness, the minimum printable line width at a defined production web speed, and yield/uniformity). For wearables and smart packaging, biocompatibility/skin-contact safety and recycling routes are also critical. We therefore summarize the target figures-of-merit and qualification tests to enable scale-up (Fig. 12).


image file: d5mh01421b-f12.tif
Fig. 12 MXene-based flexible electronics and intelligent sensing: overview of technical bottlenecks, frontier research directions, and application expansion. (A) Configuration of Sb2S3/Ti3C2Tx//AC/rGO Na+-ASCs. Reproduced from ref. 211 with permission from Springer Nature, Copyright 2021. (B) Schematic of the fabrication procedure and the morphological difference between the pure MXene (MX) film and polydopamine-treated MXene (PDTM) film. Reproduced from ref. 213 with permission from American Chemical Society, Copyright 2020. (C) Schematic of the PL-MXene electrode composed of MXene flakes with a protective PVPh polymer layer. Reproduced from ref. 224 with permission from American Chemical Society, Copyright 2021. (D) Application of the 3D-printing method for fabricating MXenes. Reproduced from ref. 49 with permission from the Royal Society of Chemistry, Copyright 2024. (E) Schematic of the fabrication of the aerogel films. Reproduced from ref. 42 with permission from John Wiley and Sons, Copyright 2022. (F) EPD process for the fabrication of MXene-CF and m-MXene-CF. Reproduced from ref. 225 with permission from Elsevier, Copyright 2022. (G) MXene-based pressure sensor (M-PS), both fabricated via additive manufacturing (3D printing) using home-modified MXene ink, a power-management circuitry, an energy-storing circuitry, data collecting, and wireless data/power transmitting modules. Optical image of wearable MSP2S3. The inset “on” and “off” statuses of the LED, respectively, represent the valley and peak of the pulse signal detected by the MSP2S3 in mode one. Reproduced from ref. 53 with permission from Elsevier, Copyright 2022.

6.3 Expanding emerging application areas

6.3.1 Flexible self-powered sensing systems. The high conductivity and mechanical flexibility of MXene-printed electrodes make them exceptional for flexible piezoresistive/capacitive sensors.226,227 For instance, flexible transparent electrodes (MXAg@PMMA) hybridizing MXene with silver nanowires (AgNWs) exhibit high light transmittance (83.8%), tunable work function (4.5–4.7 eV), and excellent environmental stability. These properties make them suitable for fully solution-processed devices, like flexible quantum dot light-emitting diodes (QLEDs).21,228,229 MXene/silver nanoparticle composite electrodes can be integrated onto electronic skin via spray coating, enabling high-sensitivity monitoring of human motion (e.g., joint bending) with response times under 50 ms.176,224 Further integration with triboelectric nanogenerators (TENGs) allows MXene-based self-powered systems to simultaneously harvest mechanical energy and detect physiological signals, powering wearable devices.31,53 Combining MXene's photothermal conversion capability and conductivity with the thermally induced deformation properties of liquid crystal elastomers (LCEs)230–232 has successfully enabled intelligent bilayer soft actuators capable of self-sensing and closed-loop control.233 This provides a new pathway for the development of future high-precision bionic robots and smart actuators.203
6.3.2 High-voltage miniature energy storage devices. Addressing the bottleneck of low operating voltage (typically ≤0.6 V) in MXene-based supercapacitors, novel asymmetric designs (e.g., MXene//MnO2) combined with ionic liquid gel electrolytes can elevate the operating voltage beyond 1.8 V while mitigating MXene oxidation.27,44 These high-voltage micro-supercapacitors (MSCs) exhibit high energy density (≈25 μWh cm−2) and long cycle stability (>10[thin space (1/6-em)]000 cycles) in applications like flexible displays and miniature robotic actuators.33,61 Furthermore, strategies such as modulating MXene functional groups, introducing BC interlayer expansion, or designing MXene/nitrogen-doped carbon nanosheet composite electrodes have significantly enhanced the energy density and power characteristics of flexible supercapacitors, making them suitable for miniature/flexible energy storage scenarios.35
6.3.3 Environmental remediation and energy conversion. Breakthroughs have been achieved using MXene-printed electrodes in capacitive deionization (CDI) and osmotic energy conversion.234 For example, MXene@COF heterojunction electrodes, with ion selectivity tuned via interfacial engineering, achieved a NaCl adsorption capacity of 53.1 mg g−1 in oxygenated saline solutions, outperforming traditional carbon-based materials in cycle stability.46,235 Additionally, MXene-based membrane reactors can convert low-grade waste heat into electricity, achieving a thermoelectric conversion efficiency (≈1.2%) that is nearly three times higher than that of conventional carbon-based electrodes.38,236 MXene/alginate/PVA composite fibers, featuring a dense structure cross-linked by Ca2+ ions, exhibit exceptional acid–base stability (stable for over 1 month at pH 1–13), high hygroscopicity (moisture absorption of 10.7 g g−1), and outstanding flame retardancy (limiting oxygen index of 54%, capable of completely suppressing smoldering). These properties make them suitable for wastewater treatment, fireproof textiles, and atmospheric water harvesting.36 Researchers have constructed MXene-confined structures that are synergistically catalyzed with copper oxide nanoparticles, achieving highly efficient removal of environmental antibiotic pollutants. This platform also demonstrates significant potential for energy applications based on its structure, electrochemical performance, and reaction mechanisms (Fig. 13).237
image file: d5mh01421b-f13.tif
Fig. 13 Design of self-supporting flexible P-MXene/CPolymer-A film electrodes. (A) Schematic showing the basic preparation process of the P-MXene/CPolymer-A film electrodes. Reproduced from ref. 33 with permission from American Chemical Society, Copyright 2022. (B) Comparison of the cyclic stability of the P-MXene/CPAQ-A electrode with the pristine MXene and P-MXene-A electrodes, tested at a current density of 20 A g−1. The inset shows the GCD curves of these three electrodes before and after 40[thin space (1/6-em)]000 charging/discharging cycles. Reproduced from ref. 33 with permission from American Chemical Society, Copyright 2022.

7. Conclusion

This review provides a comprehensive overview of recent progress in MXene-based printable electrodes, with a particular focus on ink formulation, additive manufacturing strategies, structural design, interfacial engineering, and device-level performance optimization. The unique combination of tunable surface terminations, metallic conductivity, and two-dimensional morphology positions MXenes as a compelling class of materials for various additive manufacturing techniques, including inkjet printing, screen printing, and extrusion-based direct writing. A key emphasis was placed on the rheological engineering of MXene inks, where additives, surfactants, and co-solvents are employed to modulate shear-thinning behavior, yield stress, and ink stability. Such control is pivotal for achieving high-resolution, defect-free patterning, especially on flexible substrates. Simultaneously, the microstructure of printed MXene architectures can be tailored via solvent evaporation kinetics, substrate interactions, and hybrid assembly strategies to optimize electrical conductivity, mechanical compliance, and electrochemically active surface area. In addition, interfacial and surface engineering approaches, such as terminal group modulation, polymer hybridization, and multilayer stacking, have been shown to significantly enhance environmental stability, mechanical integrity, and compatibility with functional interfaces, including organic semiconductors, biosensing platforms, and electrocatalytic layers. Comparative insights with other advanced printable materials (e.g., CNTs, graphene, and transition metal oxides) underscore MXene advantages in terms of conductivity, ink processability, and functional integration while revealing critical challenges, such as susceptibility to oxidation and long-term mechanical fatigue. Despite these limitations, MXene-based printed electronics are rapidly expanding into a broad range of applications, including neuromorphic circuits, wearable and implantable sensors, human–machine interfaces, and energy conversion/storage systems. The convergence of materials innovation, precision additive manufacturing, and artificial intelligence-driven ink design is expected to accelerate the translation of MXene architectures into next-generation electronics with enhanced performance, sustainability, and multifunctionality. Looking forward, future efforts should prioritize the stabilization of MXene inks under ambient conditions, the development of environmentally benign synthesis and patterning protocols, and the scalable integration of MXenes with soft and biocompatible materials for emerging applications in bioelectronics, healthcare, and energy autonomy. Interdisciplinary collaborations bridging materials science, device engineering, and biomedical technologies will be essential to fully unlock the potential of MXene-based printed electronics and transition them from laboratory innovations to commercially viable solutions and compatibility with functional interfaces, including organic semiconductors, biosensing platforms, and electrocatalytic layers. Comparative insights with other advanced printable materials (e.g., CNTs, graphene, and transition metal oxides) underscore MXene advantages in terms of conductivity, ink processability, and functional integration while revealing critical challenges such as susceptibility to oxidation and long-term mechanical fatigue. Despite these limitations, MXene-based printed electronics are rapidly expanding into a broad range of applications, including neuromorphic circuits, wearable and implantable sensors, human–machine interfaces, and energy conversion/storage systems. The convergence of materials innovation, precision additive manufacturing, and artificial intelligence-driven ink design is expected to accelerate the translation of MXene architectures into next-generation electronics with enhanced performance, sustainability, and multifunctionality. Looking forward, future efforts should prioritize the stabilization of MXene inks under ambient conditions, the development of environmentally benign synthesis and patterning protocols, and the scalable integration of MXenes with soft and biocompatible materials for emerging applications in bioelectronics, healthcare, and energy autonomy. Interdisciplinary collaborations bridging materials science, device engineering, and biomedical technologies are essential for fully unlocking the potential of MXene-based printed electronics and transitioning them from laboratory innovations to commercially viable solutions.

Conflicts of interest

There are no conflicts to declare.

Data availability

No primary research results, software or code have been included and no new data were generated or analysed as part of this review.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant No. 52003253) and the Henan Science and Technology Department (Grant No. 222301420004).

References

  1. Y. Gogotsi and B. Anasori, ACS Nano, 2019, 13, 8491–8494 CrossRef CAS .
  2. M. Naguib, M. Kurtoglu, V. Presser, J. Lu, J. Niu, M. Heon, L. Hultman, Y. Gogotsi and M. W. Barsoum, Adv. Mater., 2011, 23, 4248–4253 CrossRef CAS .
  3. M. Naguib, V. N. Mochalin, M. W. Barsoum and Y. Gogotsi, Adv. Mater., 2014, 26, 992–1005 CrossRef CAS PubMed .
  4. H. Ding, Y. Li, M. Li, K. Chen, K. Liang, G. Chen, J. Lu, J. Palisaitis, P. O. Å. Persson, P. Eklund, L. Hultman, S. Du, Z. Chai, Y. Gogotsi and Q. Huang, Science, 2023, 379, 1130–1135 CrossRef CAS PubMed .
  5. Z. Wu, S. Liu, Z. Hao and X. Liu, Adv. Sci., 2023, 10, 2207174 CrossRef CAS .
  6. S. Deng, T. Guo, F. Nüesch, J. Heier and C. Zhang, Adv. Sci., 2023, 10, 2300660 CrossRef CAS PubMed .
  7. M. Namvari and B. K. Chakrabarti, Adv. Colloid Interface Sci., 2024, 331, 103208 CrossRef CAS PubMed .
  8. V. A. O. P. Silva, W. S. Fernandes-Junior, D. P. Rocha, J. S. Stefano, R. A. A. Munoz, J. A. Bonacin and B. C. Janegitz, Biosens. Bioelectron., 2020, 170, 112684 CrossRef CAS PubMed .
  9. M. Chao, L. He, M. Gong, N. Li, X. Li, L. Peng, F. Shi, L. Zhang and P. Wan, ACS Nano, 2021, 15, 9746–9758 CrossRef CAS PubMed .
  10. E. Pomerantseva, F. Bonaccorso, X. Feng, Y. Cui and Y. Gogotsi, Science, 2019, 366, eaan8285 CrossRef CAS .
  11. C. Yang, X. Wu, H. Xia, J. Zhou, Y. Wu, R. Yang, G. Zhou and L. Qiu, ACS Nano, 2022, 16, 2699–2710 CrossRef CAS .
  12. Y. Luo, W. Que, Y. Tang, Y. Kang, X. Bin, Z. Wu, B. Yuliarto, B. Gao, J. Henzie and Y. Yamauchi, ACS Nano, 2024, 18, 11675–11687 CrossRef CAS .
  13. G. Qiu, Y.-Q.-Q. Yi, L. Xie, F. Su, T. Wang, W. Su and Z. Cui, Sci. China Mater., 2024, 67, 205–213 CrossRef CAS .
  14. W. Zhang, C. Zeng, M. Zhang, C. Zhao, D. Chao, G. Zhou and C. Zhang, Angew. Chem., Int. Ed., 2025, 64, e202413728 CrossRef CAS PubMed .
  15. X. Li, M. Li, K. Luo, Y. Hou, P. Li, Q. Yang, Z. Huang, G. Liang, Z. Chen, S. Du, Q. Huang and C. Zhi, ACS Nano, 2022, 16, 813–822 CrossRef CAS PubMed .
  16. A. Chen, H. Wei, Z. Peng, Y. Wang, S. Akinlabi, Z. Guo, F. Gao, S. Duan, X. He, C. Jia and B. B. Xu, Small, 2024, 20, 2404011 CrossRef CAS .
  17. W. Peng, C. Liu, C. Xu, C. Qin, N. Qin, H. Chen, T. Guo and W. Hu, Sci. China Mater., 2024, 67, 2661–2670 CrossRef CAS .
  18. A. Jalali, T. Gupta, V. Pakharenko, Z. B. Rejeb, M. Kheradmandkeysomi, M. Sain and C. B. Park, Carbohydr. Polym., 2025, 352, 123252 CrossRef CAS .
  19. Z. Deng, L. Li, P. Tang, C. Jiao, Z.-Z. Yu, C. M. Koo and H.-B. Zhang, ACS Nano, 2022, 16, 16976–16986 CrossRef CAS .
  20. G. Shi, Y. Zhu, M. Batmunkh, M. Ingram, Y. Huang, Z. Chen, Y. Wei, L. Zhong, X. Peng and Y. L. Zhong, ACS Nano, 2022, 16, 14723–14736 CrossRef CAS .
  21. W. Jiang, S. Lee, K. Zhao, K. Lee, H. Han, J. Oh, H. Lee, H. Kim, C. M. Koo and C. Park, ACS Nano, 2022, 16, 9203–9213 CrossRef CAS .
  22. Q.-T. Lai, X.-H. Zhao, Q.-J. Sun, Z. Tang, X.-G. Tang and V. A. L. Roy, Small, 2023, 19, 2300283 CrossRef CAS .
  23. B. Zhao, J. Wu, Z. Liang, W. Liang, H. Yang, D. Li, W. Qin, M. Peng, Y. Sun and L. Jiang, Adv. Sci., 2022, 9, 2204751 CrossRef CAS .
  24. H. Wu, Y. Xie, Y. Ma, B. Zhang, B. Xia, P. Zhang, W. Qian, D. He, X. Zhang, B.-W. Li and C.-W. Nan, Small, 2022, 18, 2107087 CrossRef CAS .
  25. Y. Zheng, Y. Wang, D. Liu, J. Zhao and Y. Li, Angew. Chem., Int. Ed., 2025, 64, e202415742 CrossRef CAS .
  26. C. Zhang, R. Schneider, M. Jafarpour, F. Nüesch, S. Abdolhosseinzadeh and J. Heier, Small, 2023, 19, 2300357 CrossRef CAS PubMed .
  27. X. Wei, Z. Zhang, B. Wei and Q. Zhang, Small, 2024, 20, 2402422 CrossRef CAS .
  28. A. Osman, H. Liu and J. Lu, Chem. Eng. J., 2024, 484, 149461 CrossRef CAS .
  29. E. Aydin, J. K. El-Demellawi, E. Yarali, F. Aljamaan, S. Sansoni, A. U. Rehman, G. Harrison, J. Kang, A. El Labban, M. De Bastiani, A. Razzaq, E. Van Kerschaver, T. G. Allen, O. F. Mohammed, T. Anthopoulos, H. N. Alshareef and S. De Wolf, ACS Nano, 2022, 16, 2419–2428 CrossRef CAS PubMed .
  30. L. Luo, Y. Huang, K. Cheng, A. Alhassan, M. Alqahtani, L. Tang, Z. Wang and J. Wu, Light: Sci. Appl., 2021, 10, 177 CrossRef CAS PubMed .
  31. X. Hu, Y. Chen, W. Xu, Y. Zhu, D. Kim, Y. Fan, B. Yu and Y. Chen, Small, 2023, 19, 2301604 CrossRef CAS PubMed .
  32. S. Praveen, T. Kim, S. P. Jung and C. W. Lee, Small, 2023, 19, 2205817 CrossRef CAS .
  33. R. Ma, X. Zhang, J. Zhuo, L. Cao, Y. Song, Y. Yin, X. Wang, G. Yang and F. Yi, ACS Nano, 2022, 16, 9713–9727 CrossRef CAS .
  34. L. Wang, Y. Hao, L. Deng, F. Hu, S. Zhao, L. Li and S. Peng, Nat. Commun., 2022, 13, 5785 CrossRef CAS .
  35. Z. Cao, Y.-B. Zhu, K. Chen, Q. Wang, Y. Li, X. Xing, J. Ru, L.-G. Meng, J. Shu, N. Shpigel and L.-F. Chen, Adv. Mater., 2024, 36, 2401271 CrossRef CAS .
  36. K. Wu, Y. Liu, C. Geng and X. Li, Small, 2025, 21, e2411459 CrossRef PubMed .
  37. D. Yu, K. Guo, F. Hou, Y. Zhang, X. Ye, Y. Zhang, P. Ji, U. Khalilov, G. Wang, X. Zhang, K. Wang, Y. Song, X. Zhong, H. Sun, J. Zhu, J. Liang and H. Wang, Small, 2024, 20, 2312167 CrossRef CAS PubMed .
  38. Z. Liu, S. Wei, Z. Hu, M. Zhu, G. Chen and Y. Huang, Small Methods, 2023, 7, 2300190 CrossRef CAS .
  39. M. H. Pazhaya Puthanveettil, J. R. Pradhan, S. K. Mondal and S. Dasgupta, Small Methods, 2025, 2500235 CrossRef CAS .
  40. J. Zhao, N. Ma, Y. Wang, Z. Wang, T. Wang, B. Liang, Y. Zhang, J. Han, C. Zhi and J. Fan, Angew. Chem., Int. Ed., 2025, 64, e202421224 CrossRef CAS PubMed .
  41. Q. Liang, K. Liu, T. Xu, Y. Wang, M. Zhang, Q. Zhao, W. Zhong, X.-M. Cai, Z. Zhao and C. Si, Small, 2024, 20, 2307344 CrossRef CAS .
  42. Y. Luo, Y. Tang, X. Bin, C. Xia and W. Que, Small, 2022, 18, 2204917 CrossRef CAS PubMed .
  43. M. A. S. R. Saadi, A. Maguire, N. T. Pottackal, M. S. H. Thakur, M. M. Ikram, A. J. Hart, P. M. Ajayan and M. M. Rahman, Adv. Mater., 2022, 34, 2108855 CrossRef CAS .
  44. Y. Zhu, J. Ma, P. Das, S. Wang and Z.-S. Wu, Small Methods, 2023, 7, 2201609 CrossRef CAS PubMed .
  45. H. Liu, F. Xing, P. Yu, M. Zhe, X. Duan, M. Liu, Z. Xiang and U. Ritz, Int. J. Biol. Macromol., 2024, 268, 131623 CrossRef CAS .
  46. S. Zhang, X. Xu, X. Liu, Q. Yang, N. Shang, X. Zhao, X. Zang, C. Wang, Z. Wang and J. G. Shapter, Mater. Horiz., 2022, 9, 1708–1716 RSC .
  47. H. Chen, H. Wang and C. Li, Adv. Mater., 2022, 34, 2205723 CrossRef CAS .
  48. F. Ming, H. Liang, G. Huang, Z. Bayhan and H. N. Alshareef, Adv. Mater., 2021, 33, 2004039 CrossRef CAS .
  49. A. Bigham, A. Zarepour, A. Khosravi, S. Iravani and A. Zarrabi, Mater. Horiz., 2024, 11, 6257–6288 RSC .
  50. S. Y. Jeong, Y. Jeon, E. Kim, G. Lee, Y.-W. Oh, C. W. Ahn, E. H. Cho, Y. Lee and K. C. Choi, ACS Nano, 2023, 17, 10353–10364 CrossRef CAS .
  51. X. Yang, T. Lv and J. Qiu, Small, 2023, 19, 2300336 CrossRef CAS .
  52. X. Mi, L. Liu, S. Yang, P. Wu, W. Zhan, X. Ji and J. Liang, Nat. Commun., 2025, 16, 2590 CrossRef CAS .
  53. Q. Yi, X. Pei, P. Das, H. Qin, S. W. Lee and R. Esfandyarpour, Nano Energy, 2022, 101, 107511 CrossRef CAS .
  54. W. Eom, H. Shin, R. B. Ambade, S. H. Lee, K. H. Lee, D. J. Kang and T. H. Han, Nat. Commun., 2020, 11, 2825 CrossRef CAS .
  55. S. Zhao, Z. Liu, G. Xie, X. Guo, Z. Guo, F. Song, G. Li, C. Chen, X. Xie, N. Zhang, B. Sun, S. Guo and G. Wang, Angew. Chem., Int. Ed., 2021, 60, 26246–26253 CrossRef CAS .
  56. H. Xing, X. Li, Y. Lu, Y. Wu, Y. He, Q. Chen, Q. Liu and R. P. S. Han, Sens. Actuators, B, 2022, 361, 131704 CrossRef CAS .
  57. H. J. Jo, M. S. Kang, H. J. Heo, H. J. Jang, R. Park, S. W. Hong, Y. H. Kim and D.-W. Han, Int. J. Biol. Macromol., 2024, 265, 130696 CrossRef CAS PubMed .
  58. Y. Wang, Y. Zhang, D. Cao, T. Ji, H. Ren, G. Wang, Q. Wu and H. Zhu, Small Methods, 2023, 7, e2201344 CrossRef PubMed .
  59. Y. Shao, L. Wei, X. Wu, C. Jiang, Y. Yao, B. Peng, H. Chen, J. Huangfu, Y. Ying, C. J. Zhang and J. Ping, Nat. Commun., 2022, 13, 3223 CrossRef PubMed .
  60. H. Ye, Y. He, T. You and F. Xu, Adv. Funct. Mater., 2025, 35, 2413343 CrossRef CAS .
  61. H. Huang and W. Yang, ACS Nano, 2024, 18, 4651–4682 CrossRef CAS .
  62. T. Y. Ko, H. Ye, G. Murali, S.-Y. Lee, Y. H. Park, J. Lee, J. Lee, D.-J. Yun, Y. Gogotsi and S. J. Kim, Nat. Commun., 2024, 15, 3459 CrossRef CAS .
  63. S. Nouseen and M. Pumera, Adv. Funct. Mater., 2025, 35, 2421987 CrossRef CAS .
  64. B. Liu, C. Zhang, J. Li, G. Zhang, X. Jian and Z. Weng, Sci. China Mater., 2024, 67, 4049–4058 CrossRef CAS .
  65. C. Liu, K. Hao, R. Yu, R. Li, A. Huang, C. Wu, K. Zheng, Y. Yang and X. Zuo, Sci. China Mater., 2024, 67, 3710–3718 CrossRef CAS .
  66. H. H. Sim, J. H. Kim, J. Bae, C. Yoo, D.-S. Kim, J. Pyo and S. K. Seol, Small, 2025, 21, 2409198 CrossRef CAS .
  67. Q. Lu, W. Liu, X. Liu, D. Yu, Z. Song, H. Wang, G. Li and S. Ge, Sci. China Mater., 2023, 66, 3723–3734 CrossRef CAS .
  68. V. Orts Mercadillo, K. C. Chan, M. Caironi, A. Athanassiou, I. A. Kinloch, M. Bissett and P. Cataldi, Adv. Funct. Mater., 2022, 32, 2204772 CrossRef CAS .
  69. L. Meng, W. Wang, B. Xu, J. Qin, K. Zhang and H. Liu, ACS Nano, 2023, 17, 4180–4192 CrossRef CAS PubMed .
  70. Y. Long, Y. Tao, W. Lv and Q.-H. Yang, Acc. Chem. Res., 2024, 57, 2689–2699 CrossRef CAS .
  71. J. Lin, J. Huang, Z. Guo, B. B. Xu, Y. Cao, J. Ren, H. Hou, Y. Xiao, M. Elashiry, Z. M. El-Bahy and Y. Min, Small, 2024, 20, 2402938 CrossRef CAS .
  72. X. Wu, T. Tu, Y. Dai, P. Tang, Y. Zhang, Z. Deng, L. Li, H.-B. Zhang and Z.-Z. Yu, Nano-Micro Lett., 2021, 13, 148 CrossRef CAS .
  73. H. Park, J. J. Park, P.-D. Bui, H. Yoon, C. P. Grigoropoulos, D. Lee and S. H. Ko, Adv. Mater., 2024, 36, 2307586 CrossRef CAS .
  74. J. Ren, Y. Zhang, D. Zhao, Y. Chen, S. Guan, Y. Liu, L. Liu, S. Peng, F. Kong, J. D. Poplawsky, G. Gao, T. Voisin, K. An, Y. M. Wang, K. Y. Xie, T. Zhu and W. Chen, Nature, 2022, 608, 62–68 CrossRef CAS PubMed .
  75. S. Qu, L. Wang, S. Zhang, C. Yang, H. Y. Chia, G. Wu, Z. Hu, J. Ding, W. Yan, Y. Zhang, C. H. Chan, W. Chen, Y. Lu and X. Song, Nat. Commun., 2025, 16, 3234 CrossRef CAS .
  76. C. Zheng, Y. Yao, X. Rui, Y. Feng, D. Yang, H. Pan and Y. Yu, Adv. Mater., 2022, 34, 2204988 CrossRef CAS .
  77. C. Liu, Z. Feng, T. Yin, T. Wan, P. Guan, M. Li, L. Hu, C.-H. Lin, Z. Han, H. Xu, W. Cheng, T. Wu, G. Liu, Y. Zhou, S. Peng, C. Wang and D. Chu, Adv. Mater., 2024, 36, 2403791 CrossRef CAS .
  78. D. Wu, B. Yao, S. Wu, H. Hingorani, Q. Cui, M. Hua, I. Frenkel, Y. Du, T. K. Hsiai and X. He, Adv. Mater., 2022, 34, 2201772 CrossRef CAS .
  79. X. Tang, K. Wu, X. Qi, H.-J. Kwon, R. Wang, Z. Li, H. Ye, J. Hong, H. H. Choi and H. Kong, ACS Appl. Nano Mater., 2022, 5, 4801–4811 CrossRef CAS .
  80. K. Wu, X. Tang, E. J. An, Y. H. Yun, S.-J. Kim, H. C. Moon, H. Kong, S. H. Kim and Y. J. Jeong, Org. Electron., 2022, 108, 106603 CrossRef CAS .
  81. S. Ren, X. Pan, Y. Zhang, J. Xu, Z. Liu, X. Zhang, X. Li, X. Gao, Y. Zhong, S. Chen and S.-D. Wang, Small, 2024, 20, 2401346 CrossRef CAS .
  82. J. Li, J. Hao, R. Wang, Q. Yuan, T. Wang, L. Pan, J. Li and C. Wang, Battery Energy, 2024, 3, 20230033 CrossRef CAS .
  83. L. Děkanovský, J. Azadmanjiri, M. Havlík, P. Bhupender, J. Šturala, V. Mazánek, A. Michalcová, L. Zeng, E. Olsson, B. Khezri and Z. Sofer, Small Methods, 2023, 7, 2201329 CrossRef PubMed .
  84. P. Zhang, Y. Fu, X. Zhang, X. Zhang, B.-W. Li and C.-W. Nan, Sci. Bull., 2022, 67, 2541–2549 CrossRef CAS PubMed .
  85. H. Zhan, B. Wen, B. Tian, K. Zheng, Q. Li and W. Wu, Small, 2024, 20, 2400740 CrossRef CAS .
  86. J. Wang, D. Jiang, Y. Du, M. Zhang, Y. Sun, M. Jiang, J. Xu and J. Liu, Small, 2023, 19, 2303043 CrossRef CAS .
  87. X. Xia, J. Yang, Y. Liu, J. Zhang, J. Shang, B. Liu, S. Li and W. Li, Adv. Sci., 2023, 10, 2204875 CrossRef CAS PubMed .
  88. I. M. Nugraha, J. Olchowka, C. Brochon, D. Flahaut, M. Bousquet, B. Cabannes-Boue, R. Bianchini Nuernberg, É. Cloutet and L. Croguennec, Adv. Sci., 2024, 11, 2409403 CrossRef CAS .
  89. M. Wang, D. Li, H. Xu, L. Wang, Y. Li, G. Li, J. Li and W. Han, Small, 2024, 20, 2305530 CrossRef CAS .
  90. J. Wang, Q. Xu, J. Liu, W. Kong and L. Shi, Small, 2023, 19, 2301709 CrossRef CAS .
  91. W. Chen, L.-X. Liu, H.-B. Zhang and Z.-Z. Yu, ACS Nano, 2021, 15, 7668–7681 CrossRef CAS .
  92. F. Bian, R. Huang, X. Li, J. Hu and S. Lin, Adv. Sci., 2024, 11, 2307482 CrossRef CAS .
  93. Z. Han, Y. Niu, X. Shi, D. Pan, H. Liu, H. Qiu, W. Chen, B. B. Xu, Z. M. El-Bahy, H. Hou, E. R. Elsharkawy, M. A. Amin, C. Liu and Z. Guo, Nano-Micro Lett., 2024, 16, 195 CrossRef CAS .
  94. X. Tang, G. Murali, H. Lee, S. Park, S. Lee, S. M. Oh, J. Lee, T. Y. Ko, C. M. Koo, Y. J. Jeong, T. K. An, I. In and S. H. Kim, Adv. Funct. Mater., 2021, 31, 2010897 CrossRef CAS .
  95. S. Lei, T. Yang, C. Tian, T. Yan, X. Li, W. Song, W. Liu, L. Yan and Y. Zhao, Desalination, 2023, 574, 117248 CrossRef .
  96. J. Cai, P. Wu, R. Tripathi, J. Kong, Z. Chen and J. Appenzeller, ACS Nano, 2024, 18, 23489–23496 CrossRef CAS PubMed .
  97. I.-J. Park, T. I. Kim, I.-T. Cho, C.-W. Song, J.-W. Yang, H. Park, W.-S. Cheong, S. G. Im, J.-H. Lee and S.-Y. Choi, Nano Res., 2018, 11, 274–286 CrossRef CAS .
  98. K. Liu, Y. Bian, J. Kuang, Q. Li, Y. Liu, W. Shi, Z. Zhao, X. Huang, Z. Zhu, Y. Guo and Y. Liu, Giant, 2021, 7, 100060 CrossRef CAS .
  99. T. Larsen, J. D. C. Christiansen, J. R. Royer, F. H. J. Laidlaw, W. C. K. Poon, T. Larsen and S. J. Andreasen, Phys. Rev. Appl., 2024, 22, 034023 CrossRef CAS .
  100. H. Singh, M. Srivastava, N. A. Salih, P. K. Singh, M. Z. A. Yahya, S. N. F. Yusuf, M. Diantoro, F. A. Latif, N. Jain, R. C. Singh and S. Rawat, Ionics, 2025, 31, 3807–3815 CrossRef CAS .
  101. X. He and Q. Lu, Adv. Colloid Interface Sci., 2024, 324, 103086 CrossRef CAS .
  102. Q. Wang, X. Zhang, J. Tian, C. Zheng, M. R. Khan, J. Guo, W. Zhu, Y. Jin, H. Xiao, J. Song and O. J. Rojas, Carbohydr. Polym., 2023, 315, 121000 CrossRef CAS PubMed .
  103. J. Kim and J. Lee, Small, 2022, 18, 2108069 CrossRef CAS .
  104. X. Shi, Z. Yu, Z. Liu, N. Cao, L. Zhu, Y. Liu, K. Zhao, T. Shi, L. Yin and Z. Fan, Angew. Chem., Int. Ed., 2025, 64, e202418420 CrossRef CAS .
  105. L. Y. Zheng, D. Li and L. J. Wang, Compr. Rev. Food Sci. Food Saf., 2025, 24, e70125 CrossRef CAS PubMed .
  106. M. Bhattacharyya, A. Haridas Cp, M. Kaushal and T. Mondal, Small Methods, 2024, e2401741,  DOI:10.1002/smtd.202401741 .
  107. A. T. Tyowua, D. Harbottle and B. P. Binks, Adv. Colloid Interface Sci., 2024, 332, 103274 CrossRef CAS PubMed .
  108. J. B. Kim, H. Y. Lee, C. Chae, S. Y. Lee and S. H. Kim, Adv. Mater., 2024, 36, e2307917 CrossRef PubMed .
  109. S. Wang, C. Yang, X. Li, H. Jia, S. Liu, X. Liu, T. Minari and Q. Sun, J. Mater. Chem. C, 2022, 10, 6196–6221 RSC .
  110. S. Jambhulkar, S. Liu, P. Vala, W. Xu, D. Ravichandran, Y. Zhu, K. Bi, Q. Nian, X. Chen and K. Song, ACS Nano, 2021, 15, 12057–12068 CrossRef CAS PubMed .
  111. R. Nigmatullin, M. A. Johns and S. J. Eichhorn, Carbohydr. Polym., 2020, 250, 116953 CrossRef CAS PubMed .
  112. X. Liu, W. Ma, Z. Qiu, T. Yang, J. Wang, X. Ji and Y. Huang, ACS Nano, 2023, 17, 8420–8432 CrossRef CAS PubMed .
  113. Z. Zhang, L. Sun, F. Chen, X. Liu, X. Huo, X. Pan and C. Feng, Carbohydr. Polym., 2024, 345, 122598 CrossRef CAS .
  114. K. P. Marquez, K. M. D. Sisican, R. P. Ibabao, R. A. J. Malenab, M. A. N. Judicpa, L. Henderson, J. Zhang, K. A. S. Usman and J. M. Razal, Small Sci., 2024, 4, 2400150 CrossRef CAS PubMed .
  115. Y. Kim, E. Kim, D. Kim, C. W. Ahn, B. S. Kim, K. H. Ahn, Y. Lee and J. D. Park, Korea-Aust. Rheol. J., 2023, 35, 117–125 CrossRef .
  116. Z. Aghayar, M. Malaki and Y. Zhang, Nanomaterials, 2022, 12(23), 4346 CrossRef CAS .
  117. S. Lovato, G. H. Keetels, S. L. Toxopeus and J. W. Settels, J. Non-Newtonian Fluid Mech., 2022, 301, 104729 CrossRef CAS .
  118. Y. Li, C. Xu and L. Zhao, J. Fluid Mech., 2023, 970, A12 CrossRef CAS .
  119. Ö. Biricik and A. Mardani, Constr. Build. Mater., 2022, 339, 127688 CrossRef .
  120. M. B. Woods, Auburn University, 2023.
  121. Y. Liu, Z. Wang, Y. Pan, T. Liu, Y. Zhang, Q. Zhang, L. Mo, Q. Luo and C. Ma, Sci. China Mater., 2024, 67, 2600–2610 CrossRef CAS .
  122. G. Guo Liang, S. Agarwala and W. Y. Yeong, Adv. Mater. Interfaces, 2019, 6, 1801318 CrossRef .
  123. P. Wei, C. Cipriani, C.-M. Hsieh, K. Kamani, S. Rogers and E. Pentzer, J. Appl. Phys., 2023, 134, 100701 CrossRef CAS .
  124. G. Zhou, M.-C. Li, C. Liu, Q. Wu and C. Mei, Adv. Funct. Mater., 2022, 32, 2109593 CrossRef CAS .
  125. V. Thirumal, R. Palanisamy, J.-H. Kim, B. Babu and K. Yoo, Crystals, 2024, 14, 261 CrossRef CAS .
  126. K. A. Deo, M. K. Jaiswal, S. Abasi, G. Lokhande, S. Bhunia, T.-U. Nguyen, M. Namkoong, K. Darvesh, A. Guiseppi-Elie, L. Tian and A. K. Gaharwar, ACS Nano, 2022, 16, 8798–8811 CrossRef CAS .
  127. X. Huang, J. Huang, G. Zhou, Y. Wei, P. Wu, A. Dong and D. Yang, Small, 2022, 18, 2200829 CrossRef CAS PubMed .
  128. T. Zhou, C. Cao, S. Yuan, Z. Wang, Q. Zhu, H. Zhang, J. Yan, F. Liu, T. Xiong, Q. Cheng and L. Wei, Adv. Mater., 2023, 35, 2305807 CrossRef CAS .
  129. J. Zou, X. Jing, S. Li, Y. Chen, Y. Liu, P.-Y. Feng and X.-F. Peng, Carbohydr. Polym., 2025, 348, 122849 CrossRef CAS PubMed .
  130. K. Li, J. Zhao, A. Zhussupbekova, C. E. Shuck, L. Hughes, Y. Dong, S. Barwich, S. Vaesen, I. V. Shvets, M. Möbius, W. Schmitt, Y. Gogotsi and V. Nicolosi, Nat. Commun., 2022, 13, 6884 CrossRef CAS .
  131. L. Chen, T. Mai, X.-X. Ji, P.-L. Wang, M.-Y. Qi, Q. Liu, Y. Ding and M.-G. Ma, Chem. Eng. J., 2023, 476, 146652 CrossRef CAS .
  132. G. Zhou, M.-C. Li, C. Liu, C. Liu, Z. Li and C. Mei, Adv. Sci., 2023, 10, 2206320 CrossRef CAS PubMed .
  133. L. Huang, H. Wu, L. Ding, J. Caro and H. Wang, Angew. Chem., Int. Ed., 2024, 63, e202314638 CrossRef CAS PubMed .
  134. J. Ruan, D. Ma, K. Ouyang, S. Shen, M. Yang, Y. Wang, J. Zhao, H. Mi and P. Zhang, Nano-Micro Lett., 2023, 15, 37 CrossRef CAS .
  135. S. Liu, H. Zhang, T. Ren, H. Yu, K. Deng, Z. Wang, Y. Xu, L. Wang and H. Wang, Small, 2023, 19, 2306014 CrossRef CAS .
  136. J. Choi, M. S. Oh, A. Cho, J. Ryu, Y.-J. Kim, H. Kang, S.-Y. Cho, S. G. Im, S. J. Kim and H.-T. Jung, ACS Nano, 2023, 17, 10898–10905 CrossRef CAS PubMed .
  137. W. Qian, Y. Si, P. Chen, C. Tian, Z. Wang, P. Li, S. Li and D. He, Small, 2024, 20, 2403149 CrossRef CAS PubMed .
  138. S. Chen, J.-H. Ciou, F. Yu, J. Chen, J. Lv and P. S. Lee, Adv. Mater., 2022, 34, 2200660 CrossRef CAS .
  139. L. Xue, Y. Liu, Z. Chen, J. Zhang, Z. Luo and L. Zhang, Small, 2025, 21, 2412496 CrossRef CAS .
  140. A. Brady, K. Liang, V. Q. Vuong, R. Sacci, K. Prenger, M. Thompson, R. Matsumoto, P. Cummings, S. Irle, H.-W. Wang and M. Naguib, ACS Nano, 2021, 15, 2994–3003 CrossRef CAS .
  141. Y. Luo, J. V. Handy, T. Das, J. D. Ponis, R. Albers, Y.-H. Chiang, M. Pharr, B. J. Schultz, L. Gobbato, D. C. Brown, S. Chakraborty and S. Banerjee, Nat. Mater., 2024, 23, 960–968 CrossRef CAS .
  142. K. Prenger, Y. Sun, K. Ganeshan, A. Al-Temimy, K. Liang, C. Dun, J. J. Urban, J. Xiao, T. Petit, A. C. T. van Duin, D.-E. Jiang and M. Naguib, ACS Appl. Energy Mater., 2022, 5, 9373–9382 CrossRef CAS .
  143. Q. Chang, X. Fu, J. Gao, Z. Zhang, X. Liu, C. Huang and Y. Li, Adv. Mater., 2023, 35, e2305317 CrossRef PubMed .
  144. H. Tetik, J. Orangi, G. Yang, K. Zhao, S. B. Mujib, G. Singh, M. Beidaghi and D. Lin, Adv. Mater., 2022, 34, 2104980 CrossRef CAS PubMed .
  145. Y. Zhou, L. Yin, S. Xiang, S. Yu, H. M. Johnson, S. Wang, J. Yin, J. Zhao, Y. Luo and P. K. Chu, Adv. Sci., 2024, 11, 2304874 CrossRef CAS PubMed .
  146. Z. A. Akbar, Y. T. Malik, D.-H. Kim, S. Cho, S.-Y. Jang and J.-W. Jeon, Small, 2022, 18, 2106937 CrossRef CAS .
  147. S. Khademolqorani, S. N. Banitaba, A. Gupta, N. Poursharifi, A. A. Ghaffari, V. V. Jadhav, W. U. Arifeen, M. Singh, M. Borah, E. Chamanehpour and Y. K. Mishra, Small, 2024, 20, 2309572 CrossRef CAS .
  148. Y. Guan, Y. Cong, R. Zhao, K. Li, X. Li, H. Zhu, Q. Zhang, Z. Dong and N. Yang, Small, 2023, 19, 2301276 CrossRef CAS PubMed .
  149. K. Lv, J. Zhang, X. Zhao, N. Kong, J. Tao and J. Zhou, Small, 2022, 18, 2202203 CrossRef CAS .
  150. J. Luxa, P. Kupka, F. Lipilin, J. Šturala, A. Subramani, P. Lazar and Z. Sofer, ACS Catal., 2024, 14, 15336–15347 CrossRef CAS PubMed .
  151. H. Li, Y. Chen and Q. Tang, ChemPhysChem, 2024, 25, e202400255 CrossRef CAS PubMed .
  152. L. He, H. Zhuang, Q. Fan, P. Yu, S. Wang, Y. Pang, K. Chen and K. Liang, Mater. Horiz., 2024, 11, 4239–4255 RSC .
  153. C. Tsounis, P. V. Kumar, H. Masood, R. P. Kulkarni, G. S. Gautam, C. R. Müller, R. Amal and D. A. Kuznetsov, Angew. Chem., Int. Ed., 2023, 62, e202210828 CrossRef CAS PubMed .
  154. J. Zhao, Z. Xu, Z. Zhou, S. Xi, Y. Xia, Q. Zhang, L. Huang, L. Mei, Y. Jiang, J. Gao, Z. Zeng and C. Tan, ACS Nano, 2021, 15, 10597–10608 CrossRef CAS PubMed .
  155. K. Liang, T. Wu, S. Misra, C. Dun, S. Husmann, K. Prenger, J. J. Urban, V. Presser, R. R. Unocic, D.-E. Jiang and M. Naguib, Adv. Sci., 2024, 11, 2402708 CrossRef CAS .
  156. Y. Zhang, Z. Cao, S. Liu, Z. Du, Y. Cui, J. Gu, Y. Shi, B. Li and S. Yang, Adv. Energy Mater., 2022, 12, 2103979 CrossRef CAS .
  157. M. Yang, X. Shu, W. Pan and J. Zhang, Small, 2021, 17, 2006773 CrossRef CAS .
  158. X. Wang, L. Xu, C. Zhou, N. H. Wong, J. Sunarso, R. Ran, W. Zhou and Z. Shao, Small Sci., 2023, 3, 2300066 CrossRef CAS PubMed .
  159. X. Li, X. Wang, M. Yang, H. Meng, J. Yuan, Q. Yi, Z. Cao, K. Hou, K. Qi, L. Gao, J. Cheng, B. Wang and J. Wang, Adv. Mater., 2025, 37, 2500064 CrossRef CAS .
  160. N. Zhang, S. Huang, Z. Yuan, J. Zhu, Z. Zhao and Z. Niu, Angew. Chem., 2021, 133, 2897–2901 CrossRef .
  161. Z. Wang, Y. Wang, Y. Chen, H. Wu, Y. Wu, X. Zhao, R. P. S. Han and A. Cao, Small, 2021, 17, 2100911 CrossRef CAS PubMed .
  162. M. Xu, J. Zhu, J. Xie, Y. Mao and W. Hu, Small, 2024, 20, 2305448 CrossRef CAS .
  163. Y. Sun and W. G. Chong, Mater. Horiz., 2023, 10, 2373–2397 RSC .
  164. Y. Xie, G. Chen, Y. Tang, Z. Wang, J. Zhou, Z. Bi, X. Xuan, J. Zou, A. Zhang and C. Yang, Small, 2024, 20, 2405817 CrossRef CAS PubMed .
  165. W. Wu, J. Yuan, S. Dong and J. Hao, ACS Cent. Sci., 2021, 7, 1611–1621 CrossRef CAS .
  166. X. Zhang, C. Jia, J. Zhang, L. Zhang and X. Liu, Adv. Sci., 2024, 11, 2305201 CrossRef CAS .
  167. Z. Zhang, L. Gong, R. Luan, Y. Feng, J. Cao and C. Zhang, Adv. Sci., 2024, 11, 2305460 CrossRef CAS .
  168. G. Weng, X. Yang, Z. Wang, Y. Xu and R. Liu, Small, 2023, 19, 2303949 CrossRef CAS .
  169. L. Kang, J. Jiang, S. Liu, J. Ai, J. Hong, C. Tang, S. C. Jun, Y. Yamauchi and J. Zhang, Small, 2025, 21, 2412044 CrossRef CAS .
  170. K. Yang, B. Li, Z. Ma, J. Xu, D. Wang, Z. Zeng and D. Ho, Angew. Chem., Int. Ed., 2025, 64, e202415000 CrossRef CAS PubMed .
  171. Y. Dou, J. Su, S. Chen, T. Li, L. Wang, X. Ding, S. Song and C. Fan, Chem. Commun., 2022, 58, 6108–6111 RSC .
  172. Z. Zhao, K. Xia, Y. Hou, Z. Ye and L. Jianguo, Chem. Soc. Rev., 2021, 50, 12702–12743 RSC .
  173. J. Zhang, S. Liu, X. Wang, X. Zhang, X. Hu, L. Zhang, Q. Sun and X. Liu, Mater. Horiz., 2024, 11, 2483–2493 RSC .
  174. C. Liu, C. Gao, W. Huang, M. Lian, C. Xu, H. Chen, T. Guo and W. Hu, Sci. China Mater., 2024, 67, 1500–1508 CrossRef CAS .
  175. C. Chen, J. Yang, W. Zhou, X. Hu, T. Guo and S. Zhang, Sci. China Mater., 2024, 67, 1661–1667 CrossRef CAS .
  176. G. Wang, L. Meng, X. Ji, X. Liu, J. Liang and S. Liu, Bio-Des. Manuf., 2024, 7, 463–475 CrossRef CAS .
  177. L. Huang, D. Zhao, X. Yan, X. Liu, Q. Sun, H. Yang, X. Liu and H. Jia, Adv. Electron. Mater., 2025, 11, 2400474 CrossRef CAS .
  178. X. Hu, N. Mao, X. Yan, L. Huang, X. Liu, H. Yang, Q. Sun, X. Liu and H. Jia, Chem. Res. Chin. Univ., 2024, 40, 824–841 CrossRef CAS .
  179. X. Hu, J. Ding, H. Chen, Y. Hou, H. Li, Y. Sun and X. Liu, Sci. Bull., 2024, 69, 2675–2678 CrossRef CAS .
  180. Z. Hao, Z. Wu, S. Liu, X. Tang, J. Chen and X. Liu, J. Mater. Chem. C, 2024, 12, 9427–9454 RSC .
  181. S. Song, A. Yoon, S. Jang, J. Lynch, J. Yang, J. Han, M. Choe, Y. H. Jin, C. Y. Chen, Y. Cheon, J. Kwak, C. Jeong, H. Cheong, D. Jariwala, Z. Lee and S.-Y. Kwon, Nat. Commun., 2023, 14, 4747 CrossRef CAS PubMed .
  182. Z. Wang, J. Zhu, D. Liu, Q. Ren, R. Qiu, Q. Liu, W. Wang, Y. Jin and M. Zhang, Mater. Horiz., 2025, 12, 6842–6852 RSC .
  183. J. Miao, L. Wu, Z. Bian, Q. Zhu, T. Zhang, X. Pan, J. Hu, W. Xu, Y. Wang, Y. Xu, B. Yu, W. Ji, X. Zhang, J. Qiao, P. Samorì and Y. Zhao, ACS Nano, 2022, 16, 20647–20655 CrossRef CAS PubMed .
  184. J. Yan, D. Cao, M. Li, Q. Luo, X. Chen, L. Su and H. Shu, Small, 2023, 19, e2303675 CrossRef PubMed .
  185. J. Huang, K. Shu, N. Bu, Y. Yan, T. Zheng, M. Yang, Z. Zheng, N. Huo, J. Li and W. Gao, Sci. China Mater., 2023, 66, 4711–4722 CrossRef CAS .
  186. Z. Yang, X. Peng, J. Wang, J. Lin, C. Zhang, B. Tang, J. Zhang and W. Yang, ACS Appl. Mater. Interfaces, 2024, 16, 23752–23760 CAS .
  187. R. Kato, H. Uchiyama, T. Nishimura, K. Ueno, T. Taniguchi, K. Watanabe, E. Chen and K. Nagashio, ACS Appl. Mater. Interfaces, 2023, 15, 26977–26984 CrossRef CAS .
  188. B. Liu, X. Yue, C. Sheng, J. Chen, C. Tang, Y. Shan, J. Han, S. Shen, W. Wu, L. Li, Y. Lu, L. Hu, R. Liu, Z.-J. Qiu and C. Cong, ACS Appl. Mater. Interfaces, 2024, 16, 19247–19253 CrossRef CAS .
  189. Y. Yan, H. Ge, X. Zhang, C. Cui, J. Cheng, X. Li, X. Liu, L. Zhang and Q. Sun, Nano Lett., 2025, 4, 1358–1366 CrossRef PubMed .
  190. Y. Hou, H. Chen, W. Lian, H. Li, X. Hu and X. Liu, Adv. Funct. Mater., 2023, 33, 2306056 CrossRef CAS .
  191. M. Waldrip, Y. Yu, D. Dremann, T. Losi, B. Willner, M. Caironi, I. McCulloch and O. D. Jurchescu, Adv. Mater., 2024, 36, 2410442 CrossRef CAS PubMed .
  192. T. Guo, X. Xu, C. Liu, Y. Wang, Y. Lei, B. Fang, L. Shi, H. Liu, M. K. Hota, H. A. Al-Jawhari, X. Zhang and H. N. Alshareef, ACS Nano, 2023, 17, 8324–8332 CrossRef CAS .
  193. Z. Wang, X. Wu, S. Yang, J. Yao, X. Shen, P. Gao, X. Yao, D. Zeng, R. Li and W. Hu, Sci. China Mater., 2024, 67, 665–671 CAS .
  194. S. W. Koh, L. Rekhi, Arramel, M. D. Birowosuto, Q. T. Trinh, J. Ge, W. Yu, A. T. S. Wee, T. S. Choksi and H. Li, ACS Appl. Mater. Interfaces, 2024, 16, 66826–66836 CrossRef CAS PubMed .
  195. Z. Shao, Z. Zhao, P. Chen, J. Chen, W. Liu, X. Shen and X. Liu, Inorg. Nano-Met. Chem., 2024, 54, 898–906 CrossRef CAS .
  196. Q. Sun, H. Ge, S. Wang, X. Zhang, J. Zhang, S. Li, Z. Yao, L. Zhang and X. Liu, Mater. Horiz., 2024, 11, 5650–5661 RSC .
  197. A. Pareek, M. Y. Mehboob, M. Cieplak, M. Majdecki, H. Szabat, K. Noworyta, P. Połczyński, M. Morawiak, P. S. Sharma, C. Foroutan-Nejad and P. Gaweł, J. Am. Chem. Soc., 2025, 147, 5996–6005 CrossRef CAS PubMed .
  198. T. Sarkar, E. Stein, J. Vinokur and G. L. Frey, Mater. Horiz., 2022, 9, 2138–2146 RSC .
  199. S. Mori, S. Moles Quintero, N. Tabaka, R. Kishi, R. González Núñez, A. Harbuzaru, R. Ponce Ortiz, J. Marín-Beloqui, S. Suzuki, C. Kitamura, C. J. Gómez-García, Y. Dai, F. Negri, M. Nakano, S.-I. Kato and J. Casado, Angew. Chem., Int. Ed., 2022, 61, e202206680 CrossRef CAS .
  200. C. Huang, C. Li, B. Geng, X. Ding, J. Zhang, W. Tang, S. Duan, X. Ren and W. Hu, ACS Appl. Mater. Interfaces, 2024, 16, 30228–30238 CrossRef CAS PubMed .
  201. M. Kim, D. Yeom, Y. Seok, J. Song, H. Jang, Y. Choi, Y. Ko, K. Watanabe, T. Taniguchi and K. Lee, Nano Lett., 2024, 24, 16090–16098 CrossRef CAS PubMed .
  202. T. Chen, R. Yu, C. Gao, Z. Chen, H. Chen, T. Guo and W. Chen, Sci. China Mater., 2023, 66, 4453–4463 CrossRef CAS .
  203. Y. Yang, L. Meng, J. Zhang, Y. Gao, Z. Hao, Y. Liu, M. Niu, X. Zhang, X. Liu and S. Liu, Adv. Sci., 2024, 11, 2307862 CrossRef CAS PubMed .
  204. B. Zhuang, X. Wang, C. An, C. Wang, L. Liu, H. Chen, T. Guo and W. Hu, Sci. China Mater., 2023, 66, 2812–2821 CrossRef CAS .
  205. S. Gong, F. Zhao, Y. Zhang, H. Xu, M. Li, J. Qi, H. Wang, Z. Wang, Y. Hu, X. Fan, W. Peng, C. Li and J. Liu, J. Colloid Interface Sci., 2023, 632, 216–222 CrossRef CAS PubMed .
  206. M. Hilal, W. Yang, Y. Hwang and W. Xie, Nano-Micro Lett., 2024, 16, 84 CrossRef CAS PubMed .
  207. D. Majchrzak, K. Kulinowski, W. Olszewski, R. Kuna, D. Hlushchenko, A. Piejko, M. Grodzicki, D. Hommel and R. Kudrawiec, ACS Appl. Mater. Interfaces, 2024, 16, 59567–59575 CrossRef CAS .
  208. F. Wu, Y. Liu, J. Zhang, X. Li, H. Yang and W. Hu, Sci. China Mater., 2023, 66, 1891–1898 CrossRef CAS .
  209. R. V. Barrulas and M. C. Corvo, Gels, 2023, 9, 986 CrossRef CAS PubMed .
  210. H. Zhou, S. J. Han, H.-D. Lee, D. Zhang, M. Anayee, S. H. Jo, Y. Gogotsi and T.-W. Lee, Adv. Mater., 2022, 34, 2206377 CrossRef CAS .
  211. J. Yang, T. Wang, X. Guo, X. Sheng, J. Li, C. Wang and G. Wang, Nano Res., 2023, 16, 5592–5600 CrossRef CAS .
  212. L. Wei, H. Wu, S. Liu, Y. Zhou and X. Guo, Small, 2024, 20, 2312059 CrossRef CAS .
  213. G. S. Lee, T. Yun, H. Kim, I. H. Kim, J. Choi, S. H. Lee, H. J. Lee, H. S. Hwang, J. G. Kim, D.-W. Kim, H. M. Lee, C. M. Koo and S. O. Kim, ACS Nano, 2020, 14, 11722–11732 CrossRef CAS PubMed .
  214. A. Alghamdi, Eng. Technol. Appl. Sci. Res., 2025, 15, 21300–21305 CrossRef .
  215. S. Liu, Y. Zhao, W. Hao, X.-D. Zhang and D. Ming, Biosens. Bioelectron., 2020, 170, 112645 CrossRef CAS PubMed .
  216. T. Zhao, D. Yang, S. Hao, T. Xu, M. Zhang, W. Zhou and Z.-Z. Yu, J. Mater. Chem. A, 2023, 11, 1742–1755 RSC .
  217. U. Amara, I. Hussain, M. Ahmad, K. Mahmood and K. Zhang, Small, 2023, 19, 2205249 CrossRef CAS PubMed .
  218. L. Bi, R. Garg, N. Noriega, R. J. Wang, H. Kim, K. Vorotilo, J. C. Burrell, C. E. Shuck, F. Vitale, B. A. Patel and Y. Gogotsi, ACS Nano, 2024, 18, 23217–23231 CrossRef CAS .
  219. S. Lee, D. H. Ho, J. Jekal, S. Y. Cho, Y. J. Choi, S. Oh, Y. Y. Choi, T. Lee, K.-I. Jang and J. H. Cho, Nat. Commun., 2024, 15, 5974 CrossRef PubMed .
  220. W. Zhang, K. Jin, Z. Ren, L. Li, L. Chang, C. Zhang, R. Wang, B. Li, G. Wu and Y. Hu, Adv. Funct. Mater., 2024, 34, 2408496 CrossRef CAS .
  221. Y. Gao, J. Wei, X. Li, Q. Hu, J. Qian, N. Hao and K. Wang, Sens. Actuators, B, 2022, 364, 131897 CrossRef CAS .
  222. L. Huang, L. Ding and H. Wang, Small Sci., 2021, 1, 2100013 CrossRef CAS PubMed .
  223. J. S. Choi, J. S. Meena, S. B. Choi, S.-B. Jung and J.-W. Kim, Small, 2024, 20, 2306434 CrossRef CAS PubMed .
  224. S. Lee, E. H. Kim, S. Yu, H. Kim, C. Park, S. W. Lee, H. Han, W. Jin, K. Lee, C. E. Lee, J. Jang, C. M. Koo and C. Park, ACS Nano, 2021, 15, 8940–8952 CrossRef CAS PubMed .
  225. Y. Hu, S. Pang, J. Li, J. Jiang and D. G. Papageorgiou, Composites, Part B, 2022, 237, 109871 CrossRef CAS .
  226. S. Liu, Q. Meng, Y. Gao, J. Zhang, J. Li, Y. Yang, X. Zhang, H. Li and X. Liu, J. Mater. Chem. A, 2023, 11, 13238–13248 RSC .
  227. S. Wang, W. Deng, T. Yang, Y. Ao, H. Zhang, G. Tian, L. Deng, H. Huang, J. Huang, B. Lan and W. Yang, Adv. Funct. Mater., 2023, 33, 2214503 CrossRef CAS .
  228. X. Dong, X. Li, S. Yin, Z. Li, L. Li and J. Li, Sci. China Mater., 2024, 67, 2737–2748 CrossRef CAS .
  229. Y. Zheng, Y. Yu, W. Chen, H. Hu, T. Guo and F. Li, Sci. China Mater., 2023, 66, 2128–2145 CrossRef .
  230. Y. Zhu, Z. Xu, F. Wu, M. Wang and L. Chen, Sci. China Mater., 2023, 66, 3004–3021 CrossRef CAS .
  231. Y. Zhang, C. Song, J. Bao, Z. Li, Z. Wang, J. Xiao, M. Yu, Y. Gao, L. Zhang, R. Lan, C. Zou and H. Yang, Sci. China Mater., 2023, 66, 4803–4813 CrossRef CAS .
  232. J. Gao, X. Cong, Y. Tang and J. Guo, Sci. China Mater., 2024, 67, 355–362 CrossRef CAS .
  233. D. Wu, Y. Zhang, H. Yang, A. Wei, Y. Zhang, A. Mensah, R. Yin, P. Lv, Q. Feng and Q. Wei, Mater. Horiz., 2023, 10, 2587–2598 RSC .
  234. F. Cheng, Y. Wang, C. Cai and Y. Fu, Nano Lett., 2024, 24, 9477–9486 CrossRef CAS PubMed .
  235. S. Sun, X. Sha, J. Liang, G. Yang, X. Hu, Z. He, M. Liu, N. Zhou, X. Zhang and Y. Wei, J. Hazard. Mater., 2021, 420, 126580 CrossRef CAS PubMed .
  236. K. A. S. Usman, S. Qin, L. C. Henderson, J. Zhang, D. Y. Hegh and J. M. Razal, Mater. Horiz., 2021, 8, 2886–2912 RSC .
  237. Q. Zhou, P. Hong, X. Shi, Y. Li, K. Yao, W. Zhang, C. Wang, J. He, K. Zhang and L. Kong, J. Hazard. Mater., 2023, 448, 130995 CrossRef CAS .

This journal is © The Royal Society of Chemistry 2026
Click here to see how this site uses Cookies. View our privacy policy here.