Hangkuan
Ji
,
Zichen
Song
,
An
Wu
,
Yi-Chao
Zou
* and
Guowei
Yang
School of Materials Science and Engineering, State Key Laboratory of Optoelectronic Materials and Technologies, Nanotechnology Research Center, Sun Yat-sen University, Guangzhou, 510275, P. R. China. E-mail: zouych5@mail.sysu.edu.cn
First published on 20th February 2025
The growing need for integrating two-dimensional materials in electronic and functional devices requires the flexibility of the material. This necessitates the in situ characterization of their mechanical properties to understand their structure under stress loading in working devices. However, it is still challenging to directly characterize the mechanical behaviours of two-dimensional materials due to difficulties in handling these naturally fragile materials. In this review, we summarize the recent studies of mechanical properties in two-dimensional materials and their characterization using various microscopy techniques. This involves advances in fundamentals including the measurements of elastic properties, and the basic understanding of how structural parameters like defects and interfaces influence the deformation and failure process of two-dimensional materials. We also discuss the developed handling techniques for transferring two-dimensional materials to the characterization platforms, with the recent advances in in situ characterization studies based on atomic force microscopy and scanning/transmission electron microscopy. The above developments allowed the direct observation of unconventional mechanisms behind the deformation behaviour of two-dimensional materials, including plastic deformation, interlayer slip, phase transition and nanosized cracking. We then discuss the applications related to the mechanics of two-dimensional materials, including structural materials, electronic and optoelectronic properties, and further conclude with the opportunities and challenges in this field.
During the applications, the intrinsic properties of 2D materials determine their device stability. Therefore, it is important to understand their intrinsic elastic properties, deformation and fracture mechanisms. To date, the number of reports on the mechanics of 2D materials has been increasing annually.27–30 The corresponding characterization methods and instruments have made significant progress in determining mechanical properties that were not able to be measured experimentally before, such as the instability under stress loading,31 the fatigue life32 and the bending stiffness of heterostructures with complex 2D components.33 Meanwhile, recent developments in the fundamentals of mechanics and the fabrication techniques of materials have allowed more controllability over the transfer, stacking and assembling of 2D materials.34–38 This enables accessibility to a quantitative understanding of the mechanics of those thin samples that were too challenging to be handled before,34 the construction of regularly aligned 2D composites with high flexibility and high stiffness,38,39 as well as the unravelling of unusual deformation mechanisms30 and new functionalities in 2D materials.34,36 Therefore, a state-of-the-art review of the characterization and applications of 2D materials’ mechanics will be greatly helpful for readers to understand the research status, outline the new challenges and establish potential research directions for future.
In this review, we summarize the mechanical properties and related characterization approaches for 2D materials, and discuss how these properties influence their real device applications. In section 2, we discuss the intrinsic mechanical properties of 2D materials, including elastic moduli, fracture strength and bending stiffness. In section 3, we emphasize the recent development in instruments for characterization studies of deformation behaviours, where the understanding of the material's response to dynamic stress is new for 2D materials. The inaccuracies that may occur during measurements shall also be proposed, which may lie in the sample preparations themselves (defects and contamination from sample transfers etc.) or in the mechanical loading and measuring approaches (model simplification, local stress or global stress, imaging resolution etc.), In section 4, we discuss materials parameters that influence the fracture mechanisms of 2D materials, including layer thickness, grain boundaries, defects and interfaces. Section 5 reviews the applications based on the mechanics of 2D materials. Starting with the rational design of structural materials consisting of 2D materials, we introduce the electronic properties that are related to the strain engineering of 2D materials, including the electronic, optoelectronics, and ferroelectric properties.40,41 We then conclude with our prospects on the remaining challenges and opportunities that are potentially exciting for researchers.
![]() | ||
Fig. 1 (a) A material design plot comparing the failure strength with Young's moduli. Materials that maximize σf/E indicate that they can sustain a large elastic strain before fracture. Materials with σf/E smaller than graphene suggest that they are suitable for being integrated into graphene-based devices, as graphene cannot fracture during deformation.43 (b) Comparison of the elastic modulus and failure strength of various 2D crystals.28,29,34,42,44–47,49–83 |
Materials (fabrication methods) | Thickness | E 2D (N m−1) | Fracture strength σf (GPa) | E 3D (GPa) | Measurement methods | Ref. |
---|---|---|---|---|---|---|
Graphene (mechanical exfoliation) | 1 L | 340 | 130–110 | 1000 | AFM nanoindentation | 42 |
Graphene (mechanical exfoliation) | 1 L | 342 ± 8 | 125.0 ± 0 | 1026 ± 22 | AFM nanoindentation | 50 |
Graphene (mechanical exfoliation) | 2 L | 645 ± 16 | 107.7 ± 4.3 | 962.7 ± 23.9 | AFM nanoindentation | 50 |
Graphene (mechanical exfoliation) | 3 L | 985 ± 10 | 105.6 ± 6.0 | 980.1 ± 9.9 | AFM nanoindentation | 50 |
Graphene (mechanical exfoliation) | 8 L | 2525 ± 8 | 85.3 ± 5.4 | 942 ± 3 | AFM nanoindentation | 50 |
Graphene (CVD) | 1 L | 309 | 50–60 | 920 | SEM MEMS tensile test | 51 |
Graphene (mechanical exfoliation) | 1 L | 390 | 1147 | Bulging test | 52 | |
hBN (mechanical exfoliation) | 1 L | 289 | 70 | 865 | AFM nanoindentation | 50 |
hBN (mechanical exfoliation) | 2 L | 590 | 68 | 881 | AFM nanoindentation | 50 |
hBN (mechanical exfoliation) | 3 L | 822 | 77 | 806 | AFM nanoindentation | 50 |
hBN (mechanical exfoliation) | 9 L | 2580 | 73.5 | 856 | AFM nanoindentation | 50 |
hBN (CVD) | 1 L | 200 | 601 | SEM + TEM MEMS tensile test | 53 | |
hBN (CVD) | 1 L | 144.87 | 7.9 | 439 | SEM MEMS tensile test | 44 |
hBN (mechanical exfoliation) | 10–70 L | 770 ± 13 | Bulging test | 54 | ||
MoS2 (mechanical exfoliation) | 1 L | 180 ± 60 | 22 ± 4 | 270 ± 100 | AFM nanoindentation | 55 |
MoS2 (mechanical exfoliation) | 2 L | 260 ± 70 | 21 ± 6 | 200 ± 60 | AFM nanoindentation | 55 |
MoS2 (CVD) | 1 L | 171.6 ± 12 | 264 ± 18 | AFM nanoindentation | 56 | |
MoS2 (CVD, mechanical exfoliation) | 1 L | 190 ± 35 | 283.6 ± 52.2 | Bulging test | 57 | |
MoS2 (mechanical exfoliation) | 10–70 L | 314.3 ± 8.4 | Bulging test | 54 | ||
MoS2 (mechanical exfoliation) | 3–11 L | 246 ± 35 | Surface wrinkling | 58 | ||
MoS2 (mechanical exfoliation) | 1 L | 172 | 265 ± 13 | AFM | 59 | |
2H-MoTe2 (mechanical exfoliation) | 3.6 nm | 316 | 5.6 ± 1.3 | 110 ± 16 | AFM nanoindentation | 49 |
2H-MoTe2 (mechanical exfoliation) | 6–6.7 nm | 670 ± 5 | 5.6 ± 1.3 | 110 ± 16 | AFM nanoindentation | 49 |
1T′-MoTe2 (mechanical exfoliation) | 9–11 nm | 1010 ± 60 | 2.6 ± 0.2 | 99 ± 15 | AFM nanoindentation | 49 |
Td-MoTe2 (mechanical exfoliation) | 10.5–14 nm | 1200 ± 100 | 2.5 ± 0.9 | 102 ± 16 | AFM nanoindentation | 49 |
MoSe2 (CVD) | 1 L | 124 ± 6.5 | 3 ± 1 | 177.2 ± 9.3 | SEM MEMS tensile test | 60 |
MoSe2 (CVD) | 2 L | 248 ± 13 | 6 ± 3 | 177.2 ± 9.3 | SEM MEMS tensile test | 60 |
2H-MoSe2 (mechanical exfoliation) | 2 L | 157.38 | 122 ± 3 | Brillouin light scattering | 61 | |
MoSe2 (physical vapor transport) | 5–10 L | 224 ± 41 | Surface wrinkling | 58 | ||
MoSe2 | 1 L | 12–23 | 175–215 | DFT | 62 | |
WSe2 (mechanical exfoliation) | 5 L | 596 | 12.4 | 170.3 ± 6.7 | AFM nanoindentation | 63 |
WSe2 (mechanical exfoliation) | 6 L | 690 | 166.3 ± 6.1 | AFM nanoindentation | 63 | |
WSe2 (mechanical exfoliation) | 12 L | 1411 | 167.9 ± 7.2 | AFM nanoindentation | 63 | |
WSe2 (mechanical exfoliation) | 14 L | 1615 | 164.8 ± 5.7 | AFM nanoindentation | 63 | |
WSe2 (mechanical exfoliation) | 1 L | 168 ± 25 | 38 ± 6 | 258.6 ± 38.3 | AFM nanoindentation | 64 |
WSe2 (mechanical exfoliation) | 3 L | 489 ± 63 | 36 ± 5 | 238.9 ± 29.4 | AFM nanoindentation | 64 |
WSe2 (mechanical exfoliation) | 4–9 L | 163 ± 39 | Surface wrinkling | 58 | ||
WS2 (mechanical exfoliation) | 1 L | 187.5± 14.9 | 47± 8.6 | 302.4± 24.1 | AFM nanoindentation | 64 |
WS2 (mechanical exfoliation) | 3 L | 489.4 ± 62.7 | 40.9 ± 6.0 | 263.1 ± 33.7 | AFM nanoindentation | 64 |
WS2 (CVD) | 1 L | 177 | 272 | AFM nanoindentation | 56 | |
WS2 (mechanical exfoliation) | 3–8 L | 236 ± 65 | Surface wrinkling | 58 | ||
WTe2 (mechanical exfoliation) | 1 L | 106 ± 6.6 | 6.4 ± 3.3 | 149.1 ± 9.4 | AFM nanoindentation | 64 |
WTe2 (mechanical exfoliation) | 3 L | 272 ± 38 | 127.7 ± 18.1 | AFM nanoindentation | 64 | |
WTe2 | 1 L | 14 ± 5 | 135–150 | DFT | 62 | |
WN (CVD) | 3 nm | 764 | 260 | AFM nanoindentation | 65 | |
WN (CVD) | 4.5 nm | 1804 | 400 | AFM nanoindentation | 65 | |
WN (CVD) | 12 nm | 4490 | 370 | AFM nanoindentation | 65 | |
HfS2 (mechanical exfoliation) | 12.2 nm | 553 | 5.8 ± 0.4 | 45.3 ± 3.7 | AFM nanoindentation | 66 |
HfS2 (mechanical exfoliation) | 14 nm | 1263 | 9.4 ± 0.3 | 90.2 ± 10.2 | AFM nanoindentation | 66 |
HfSe2 (mechanical exfoliation) | 13.4 nm | 526 | 4.6 ± 1.4 | 39.3 ± 8.9 | AFM nanoindentation | 66 |
HfSe2 (mechanical exfoliation) | 15.4 nm | 370 | 2.4 ± 0.1 | 24.1 ± 8 | AFM nanoindentation | 66 |
VS2 (CVD) | 9.2 nm | 409 | 4.6 ± 0.2 | 44.4 ± 3.5 | AFM nanoindentation | 67 |
InSe (mechanical exfoliation) | 6 L | 528 ± 64 | 110 ± 13.4 | AFM nanoindentation | 68 | |
InSe (mechanical exfoliation) | 7 L | 582 ± 125 | 104 ± 22.3 | AFM nanoindentation | 68 | |
InSe (mechanical exfoliation) | 8 L | 646 ± 129 | 8.68 | 101 ± 20.0 | AFM nanoindentation | 68 |
InSe (mechanical exfoliation) | 9 L | 710 ± 118 | 99 ± 16.4 | AFM nanoindentation | 68 | |
InSe (mechanical exfoliation) | 14 L | 1068 ± 202 | 95 ± 18.0 | AFM nanoindentation | 68 | |
GaS (liquid exfoliation) | 10 nm | 1732 | 8 | 173 | AFM nanoindentation | 69 |
GaSe (liquid exfoliation) | 10 nm | 819 | 4 | 82 | AFM nanoindentation | 69 |
GaTe (mechanical exfoliation) | 10 nm | 246 | 2 | 25 | AFM nanoindentation | 69 |
Bi2Te3 (CVD) | 10 L | 187 ± 70 | 0.40 ± 0.08 | 18.7 ± 7 | AFM nanoindentation | 70 |
Bi2Se3 (CVD) | 7–12 L | 21.66 ± 3.8 | AFM nanoindentation | 71 | ||
Black phosphorus (mechanical exfoliation) | 14.3 nm | 25 | 276 ± 32.4 | AFM nanoindentation | 72 | |
Black phosphorous (mechanical exfoliation) | 11 L | 110 | 106.6 | AFM nanoindentation | 73 | |
Black phosphorus (mechanical exfoliation) | 30–34 nm | 89.7 ± 26.4 | AFM nanoindentation | 72 | ||
Ti3C2Tx (liquid exfoliation) | 1 L | 326 ± 29 | 17.3 ± 1.6 | 333 ± 30 | AFM nanoindentation | 74 |
Ti3C2Tx (liquid exfoliation) | 1 L | 473.9 | 15.4 | 484 ± 13 | SEM MEMS tensile test | 34 |
Nb4C3Tx (liquid exfoliation) | 1.26 nm | 486 ± 18 | 26 ± 1.6 | 386 ± 13 | AFM nanoindentation | 75 |
Mica (mechanical exfoliation) | 2–14 L | 4–9 | 202 ± 22 | AFM nanoindentation | 76 | |
COFTTA-DHTA (confined synthesis) | 4.7 nm | 119 ± 3 | 25.9 ± 0.6 | AFM nanoindentation | 77 | |
COF (confined synthesis) | 44 ± 7 nm polycrystalline film | 2494 ± 325 | 1.86 ± 0.2 | 56.7 ± 7.4 | AFM nanoindentation | 29 |
MOF(CuBDC) (liquid exfoliation) | 10 nm | 230 | 23 | AFM nanoindentation | 78 | |
Perovskite(C4n3) (mechanical exfoliation) | 1 L | 29.4 ± 3.6 | 0.7 ± 0.08 | 11.2 ± 1.4 | AFM nanoindentation | 79 |
Perovskite(C4n3) (mechanical exfoliation) | 2 L | 37.1 ± 4.9 | 0.44 | 7.1 ± 0.9 | AFM nanoindentation | 79 |
Perovskite(C4n3) (mechanical exfoliation) | 3 L | 43 | 0.36 | 5.7 | AFM nanoindentation | 79 |
Perovskite(C4n3) (mechanical exfoliation) | 4 L | 60 | 0.36 | 5.7 | AFM nanoindentation | 79 |
Bi2O2Se (CVD) | 5 L | 307 | 18–23 | 88.7 ± 14.4 | AFM nanoindentation | 80 |
Bi2O2Se (CVD) | 10 L | 507 | 18–23 | 88.7 ± 14.4 | AFM nanoindentation | 80 |
Bi2O2Se (CVD) | 15 L | 768 | 18–23 | 88.7 ± 14.4 | AFM nanoindentation | 80 |
Cr2Ge2Te6 (mechanical exfoliation) | 14 L | 515 | 56.2 ± 8.2 | AFM nanoindentation | 81 | |
Cr2Ge2Te6 (mechanical exfoliation) | 20 L | 770 | 55 | AFM nanoindentation | 81 | |
Cr2Ge2Te6 (mechanical exfoliation) | 34 L | 1240 | 53 | AFM nanoindentation | 81 | |
Cr2Ge2Te6 (mechanical exfoliation) | 42 L | 2000 | 70 | AFM nanoindentation | 81 | |
CrCl3 (mechanical exfoliation) | 2 L | 106 | 3.6 ± 0.4 | 62.1 ± 4.8 | AFM nanoindentation | 82 |
CrCl3 (mechanical exfoliation) | 10 L | 230 | 2.2 ± 0.2 | 27.1 ± 2.5 | AFM nanoindentation | 82 |
CrI3 (mechanical exfoliation) | 2 L | 61 | 2.2 ± 0.5 | 43.4 ± 4.4 | AFM nanoindentation | 82 |
CrI3 (mechanical exfoliation) | 9 L | 100 | 1.6 ± 0.04 | 15.8 ± 1.2 | AFM nanoindentation | 82 |
TaNi2Se5 (mechanical exfoliation) | 28 nm | 2.4 ± 0.8 for a axis, 1.2 ± 0.2for c axis | 56.9 ± 9.2 for a axis, 45.0 ± 4.5 for c axis | SEM MEMS tensile test | 28 | |
As2S3 (mechanical exfoliation) | 9–25 nm | 79.1 ± 10.1 for a axis, 47.2 ± 7.9 for c axis | AFM nanoindentation | 83 |
Beginning with graphene, the fundamental elastic properties of many other 2D materials have been subsequently revealed. Specifically, the deformability of hBN and TMDs has been extensively studied.44–47 This is because 2D electronic devices have aroused much research interest in recent decades, where hBN often acts as a dielectric layer and TMDs work as an active layer. It was found that, among the TMDs, MoS2 exhibits a comparatively high value of the elastic modulus, which remains one of the most important components for flexible electronics.33 For other types of metal chalcogenide, it was interesting to see that there are a few reported as highly deformable, where unexpected plasticity behaviour was observed,48 including InSe, GaTe and MoTe2.30 Also, it should be noted that the phase structure of the crystals can lead to variable mechanical performance. For example, in MoTe2, the breaking strengths of distorted 1T′ and Td phases are only half the value of 2H-MoTe2 phase due to the uneven distribution of bonding strengths.49
The high fracture strength measured from 2D materials requires high-quality samples, e.g. mechanical exfoliated single crystals. Therefore, there is a discrepancy between the value from single crystals for fundamental studies and their nanobulk counterparts for scalable devices, where the large elasticity is no longer preserved. Demonstrated approaches to increase the deformability and strength of 2D materials shall be discussed further in the following sections.
![]() | (1) |
![]() | ||
Fig. 2 Measuring bending stiffness through imaging the geometry of curved 2D materials. (a) Schematic showing the heterostructure used for measuring the bending stiffness of graphene. Bilayer graphene was transferred onto the top of an hBN edge step.87 (b–d) Annular dark-field scanning transmission electron microscopy (ADF-STEM) images showing the cross-sectional samples prepared from N-layer graphene samples over H-layer hBN steps. For each designed N and H, a bending profile was measured containing a radius of curvature R, step height H and bending angle θ, indicated in d.87 (e) Plots showing the bending stiffness measured as a function of the thickness of graphene from experiments. Power law relationships are denoted by the red and blue lines just for comparison.87 (f) The experimentally measured (filled symbols) and calculated (open symbols) bending stiffnesses as a function of thickness for few-layer graphene. The blue and red colouring is to show the various bending angles where the data are measured from.87 (g) Calculated contribution to bending stiffness from interfacial interaction, as a function of bending angle, based on simplified Frenkel–Kontorova model. The inset shows that the curvature is accommodated entirely by slip between the layers. Reproduced from ref. 87 with permissions from Springer Nature Limited Copyright 2019.87 |
Through varying the step height of the hBN edges (H), the influence of the deformation condition was investigated (Fig. 2c and d). It was found that the experimentally measured B–N curves can be described by a power law dependence (Fig. 2e). Beyond a certain bending angle (>40°), the B–N curves exhibit a nearly linear relationship, characteristic of a stack of frictionless plates, indicating the onset of superlubricity between the graphene atomic layers (Fig. 2f). The reduced bending stiffness at high bending angles can be explained as a result of an increased contribution from an interlayer slip, which can accommodate the large strain at high angles (Fig. 2g).87 It was also found that the bending stiffness of the graphene reaches a minimum when the thickness of the sample is reduced to a monolayer (Fig. 2e). Thus, the thickness of the 2D materials has an important impact on the flexibility of the sample that shall be discussed further in section 4.
![]() | (2) |
The failure mechanism of 2D materials in response to cyclic and impact loading has been a recent research interest. Fatigue behaviour and damage mechanisms are critical to evaluating the long-term reliability of devices made from 2D materials, because fatigue can cause material failure at stress levels significantly lower than that under static loading. A modified AFM instrument was developed recently to enable applications of both static and cyclic loading to suspended 2D materials. Through adding alternating current inputs using a ‘shake’ piezo (Fig. 3a–c), Cui et al. conducted a fatigue test on graphene and found that it exhibits a fatigue life of more than 109 cycles under a mean stress of 71 GPa, higher than any material reported so far.32 Unlike metals, there is no progressive damage during the fatigue loading of graphene, and its failure is global and catastrophic (see bottom inserts, Fig. 3b). Furthermore, this study illustrates the difference between the morphological changes of bilayer and monolayer graphene during fatigue. Bilayer samples exhibit obvious interlayer shearing and wrinkling after failure, whereas no such behaviour was observed for monolayers after one billion cyclic loadings (Fig. 3c and d). The fatigue behaviour of graphene and graphene oxide (GO) was also compared. In contrast to the catastrophic failure behaviour in graphene, monolayer GO films exhibited localized failure. This could be attributed to the enriched oxygen functional groups in graphene oxide, where epoxide-to-ether transformation can happen during stress loading, which provides trapping sites for cracking arresting, leading to the occurrence of unusual plasticity in graphene oxide.
![]() | ||
Fig. 3 Fatigue test conducted on free-standing graphene using AFM.32 (a) A schematic illustrating the set-up for fatigue testing.32 (b) A typical result from a fatigue test conducted on a bilayer graphene under a static loading with a value at half of its fracture force. The abrupt jump in the amplitude and deflection signals at ∼100 million cycles indicate the occurrence of fatigue failure. The insets are AFM images taken before and after the fatigue failure. The diameter of the sample is 2.5 μm. In such a case, the maximum in-plane stress varies from 69.5 to 75.1 GPa calculated using DFT-based nonlinear FEM.32 (c) Normalized elastic modulus E2D of monolayer and bilayer graphene. E2D values are normalized by that measured from the pristine samples without fatigue loading. The red and blue dashed lines are guides to the eye for showing the larger scatter in the E2D for the bilayer compared with monolayer.32 (d) Corresponding cyclic loading-induced wrinkling and local delamination, while no morphological change was observed for the monolayer after cyclic loading. Reproduced from ref. 32 with permissions from Springer Nature Limited Copyright 2020.32 |
However, nanoindentation tests have their own limitations. For example, they can only measure a small sample area underneath the indenter tip (tip radii < 50 nm), which does not necessarily represent the properties for the whole membrane.97–99 Also, nanoindentation cannot directly measure fracture strength σ2D. It gives an estimated value of inferred from E2D and δ, using theoretical stress–strain relationships, e.g. with R being the indenter tip radius, assuming a clamped, linear elastic, circular membrane, or σ ∼ δ curves that requires combination with theoretical predictions from density functional theory (DFT) and finite elemental analysis (FEM).32
![]() | ||
Fig. 4 (a) Digital photos of the bent metal chalcogenide samples. The smallest grid is 1 mm.39 (b) A combined index ξ, (Ec,inter × Ec,cross)/Eb,inter to predict 2D vdW materials with plasticity. A plastically deformable 2D vdW crystal should possess a large ξ.39 (c) In situ TEM compression test results, with the top showing stress–strain curves taken from a small micro-machined MoS2 pillar and the bottom part showing TEM images taken before and after the test.39 (d–g) SEM images taken from an in situ SEM compression test on InSe micropillars along (d and e) and perpendicular (f and g) to the (001) axis. Reproduced from ref. 26 with permissions from The American Association for the Advancement of Science Copyright 2020.26 |
The in situ probe test can also be used to illustrate the detailed structural evolution process of the 2D metal chalcogenides during deformation, through utilizing the resolution advantages of TEM imaging, i.e. high temporary and spatial resolution. Recently, Zhao et al. directly observed the atomic-scale plasticity mechanism in 2D InSe flakes using an in situ TEM bending test (Fig. 5), with complementary high-resolution STEM imaging.30 It was interesting to see that a phase transition from 2H to 3R occurred in InSe crystals during the deformation (Fig. 5a–c). The in situ characterization studies found that InSe exhibits an unusual plastic deformability, where not only do the interlayer gliding and formation of high-density dislocation networks play a role, but the appearance of numerous discontinuous nanoscale cracks also has an impact, which helps release the increased local elastic energy due to deformation (Fig. 5d–i). Such behaviour distinguishes InSe from other materials such as MoS2, and MoTe2, where large cracks across the whole crystal were observed.30 To illustrate the different deformation behaviour between the materials, further DFT calculations were conducted and indicated that the bonding strength of In–Se (3.855 eV per bond) is weaker than Mo–S (4.368 eV per bond). Therefore, it is relatively easier to break the intralayer In–Se bond, so that the occurrence of cracks is energetically permitted.30 DFT results also explained how phase transition initiates during the deformation of InSe, which suggests that the transition from 2H to 3R in InSe is more energetically favourable compared with those phase transitions in other materials systems such as MoS2 and MoTe2. The above work demonstrates the important role of in situ TEM tests in mechanistic studies for deformation.30
![]() | ||
Fig. 5 (a) Digital photos taken from a InSe crystal, before (left panel) and after (right panel) the ex situ compression test.30 (b) High-angle angular dark field-scanning transmission electron microscopy (HAADF-STEM) images taken from InSe crystals, before (left panel) and after (right panel) the ex situ compression test, showing the pristine sample has a 2H phase, while the compressed sample has a 3R phase.30 (c) Schematic showing the experimental set up for the in situ TEM deformation tests. Reproduced from ref. 100 with permissions from The American Association for the Advancement of Science Copyright 2022.100 (d–g) Time-series TEM snapshots, selected from the recorded video taken during the in situ bending experiment, which included TEM images taken before the fracture happens (d), with a corresponding magnified TEM image (e and f) and a rotated and high-resolution TEM image shown in g.30 (h) Corresponding HAADF-STEM images taken before (left panel) and after the in situ experiment.30 (i) Corresponding high-resolution TEM image. The fracture and nanoscale cracks formed at the edge of the sample were marked by a green line and a green box respectively. Reproduced from ref. 30 with permissions from Springer Nature Limited Copyright 2024.30 |
However, the application of in situ probe tests for 2D samples is still limited, especially for those atomically-thin ones. The mechanical loading onto a sample is based on a tip with nanometre-scale contact size (see Fig. 5), which is much smaller than the lateral size of most 2D samples while much larger than their thickness, leading to a deformation geometry limited to quasi-1D nanoscrolls. It is therefore remotely possible to use this approach to achieve a controllable and even stress loading, or a quantitative understanding of mechanics in 2D materials. This necessitates the development in micro-electro-mechanical system (MEMS) devices for quantitative experiments, which shall be discussed in the following section.
![]() | ||
Fig. 6 A schematic showing the typical transfer process for 2D materials samples onto MEMS chips, composed of three steps. (a) The polymer-assisted transfer process onto MEMS chips.28 (b) Removal of the polymer after transfer, confirmed by Raman spectroscopy.28 (c) The clamping and shaping of sample under FIB.28 |
The sample quality control is highly important for realizing a reliable and quantitative tensile test on 2D materials.51Fig. 7a shows a schematic of a shaped graphene flake on a tensile testing MEMS chip. By improving sample quality using a modified transfer method, Lu et al. measured a Young's modulus E3D ∼ 900–1000 GPa from a CVD-grown graphene, fairly close to the theoretical value of a pristine monolayer graphene.51 This work approached an elastic strain limit of ∼6% (Fig. 7b), higher than previously reported experimental values. It should be noted that this value ∼6% is still lower compared with the theoretical strain limit of ∼20%, with the representative tensile fracture strength of ∼50–60 GPa measured as being smaller than the ideal strength of monolayer graphene (∼100–130 GPa). This could result from the presence of defects when preparing shaped samples for tensile tests.51 It is well-documented that point defects, line defects, pre-cracks at the sample edge or the sample clamping sites, and oxidized impurities, can be introduced simply due to sample preparation, which leads to considerably reduced strength in 2D materials.32,51,92,94,108 The current shaping methods are generally based on FIB ion milling,109 where ion implantation can lead to surface damage and edge defects in the specimen.51 On the other hand, the transfer of 2D materials with an atomically-thin thickness, especially monolayers, is still challenging due to the low accessibility of the source materials, and the high adhesion of these samples to their original substrates. Very recently, Rong et al. developed a copper mesh-assisted transfer technique, which uses the thin flakes that attached locally to a mesh edge as the source materials.34 The limited attaching area leads to limited adhesion to the substrates, therefore allowing the subsequent transfer of the flakes by FIB nanomanipulation.34 This increases the accessibility to the intrinsic mechanics of monolayer samples. Therefore, to fully understand the fundamental factors that govern the intrinsic mechanical performance of 2D materials, there is much room for improving sample preparation and testing methods.
![]() | ||
Fig. 7 Experimental in situ tensile techniques for measuring elastic properties. (a) Schematic showing a single-crystalline graphene sample suspended over the push-to-pull (PTP) micromechanical device. Left top is an SEM image showing an overview of the PTP device actuated by an external pico-indenter. Right top, zoom-in SEM image taken from the region marked by a rectangle in the left-top inset. The yellow arrows indicate the indentation or stress loading direction. Left-bottom inset showing a Raman spectrum taken from the graphene sample. Right-bottom inset showing a TEM bright-field image taken from the sample edge, where an amorphous edge can be observed.51 (b) SEM images taken before and after an in situ tensile test conducted on the suspended graphene.51 (c) Stress–strain curves recorded during the loading and unloading process, in which the arrows denote the emergences and disappearances of the instabilities. The inset is a schematic showing the push-to-shear experimental setup for introducing shearing strain into the 2D materials.31 (d) Time-sequence SEM images taken from the in situ shear test conducted from the monolayer graphene, and the corresponding models describing the sample morphologies during deformation. The top inset is a schematic showing the geometrical parameters of the wrinkling structure.31 (e) Curves present the theoretical normalized wrinkling wavelengths as a function of the shear strains during different stages of instability, in which the solid balls are experimental data points.31 |
Meanwhile, there are tensile tests finding that flaw insensitivity exists in 2D materials.53,107 Insensitivity to the pre-existing flaw, such as voids53 and pre-cracks, was observed in hBN and 2D COF.53,107 For example, Han et al. found that the maximum tensile strain of hBN monolayers reaches ∼6%, even though some pre-existing voids were present in the testing samples.53 They found that the naturally occurring voids are not detrimental to the mechanical resistance of hBN. Instead, those defects introduced by FIB to the sample clamped region and the sample edge are responsible for the maximum strain loss of the monolayer. This indicates that the contribution from defects depends on the specific type of the defect and the materials.
For applications especially in resonator and acoustic devices, it is important to know how the morphology and structure of the 2D materials evolve in real working conditions. Recently, a mechanical push-to-shear approach was developed to illustrate the dynamic wrinkling–splitting–smoothing process of suspended 2D materials (Fig. 7c–e). Shear stress–strain curves of single-layer graphene are shown in Fig. 7c, from which the in-plane shear modulus of the monolayer graphene was determined as ∼70 GPa based on the initial linear stage, slightly larger than the previously measured result, possibly due to the initial corrugations in the sample.110 As illustrated in Fig. 7c and d, during stress loading, the first appearance of wrinkling is marked as the 1st instability. As the shear strain increases beyond a certain threshold, wrinklons are observed with a reducing wavelength of the wrinkles, marked as the 2nd instability, where the wrinkling–splitting happens at a halving of the wavelength, while for the unloading process, the smoothing happens mainly as a result of the reduced amplitude instead of wavelength changing or merging of the wrinkles. Such a difference in stability between the formation and recovery processes can be explained by the redistribution of local compressive stain. The function between D ∼ f(E,εPre,γ)λ4 was also summarized for the initial instability stage, where D denotes the bending stiffness, E denotes the Young's modulus, εpre denotes the pre-tension strain applied on the film, γ denotes the strains, and λ denotes the observed wrinkling wavelength (Fig. 7d and e). The wrinkling wavelength possesses positive correlations with bending stiffness and pre-tension and a negative correlation with Young's modulus and shear strain. Thus, the MEMS in situ tensile test provides a direct pathway to observe and understand the wrinkling behaviour of suspended 2D materials under dynamic stress loading.31
Compared with other in situ methods, the tensile test under TEM has such advantages that it combines the sub-angstrom spatial resolution imaging capability of TEM during the experiments. Fig. 8 shows such a case study conducted by Zhang et al., where the atomic structure at the crack edge was characterized using HRTEM, in combination with the in situ SEM tensile test. It was revealed that the Young's modulus and fracture strength of 2D Ta2NiSe5 (TNS) along the a-axis were measured to be 56.9 ± 9.2 GPa and 2.4 ± 0.8 GPa, respectively, both higher than those along the c-axis (45.0 ± 4.5 GPa and 1.2 ± 0.2 GPa). The correlated high-resolution TEM imaging confirms the crack paths of 2D TNS along different orientations. As shown in Fig. 8a and c, for a TNS sample stretched along the a-axis, the crack edge is sharply formed at an angle of 30° related to the c-axis, and a similar crack structure was observed from the other samples subjected to a 45° counterclockwise rotation (Fig. 8e). In contrast, when the loading direction is parallel to the c-axis (Fig. 8d), a straight crack with an angle of 90° is observed, confirmed by the atomic resolution imaging using TEM. The high-resolution imaging confirms the accuracy of the atomic models used by the subsequent DFT calculation, so that the energetic reasons behind the observed cracking behaviour can be well explained (Fig. 8f).
![]() | ||
Fig. 8 Tensile test conducted on an anisotropic 2D Ta2NiSe5 crystal. (a) An overview SEM image showing Ta2NiSe5 membrane on a “Push-to-Pull” MEMS device.28 (b) Measurements of anisotropic fracture toughness for different samples. Scale bar: 1 μm.28 (c) SEM image taken from a cracked a-axis sample, and a corresponding HRTEM image taken from the crack edge.28 (d) SEM image taken from a cracked sample with its long axis being the c-axis and the corresponding HRTEM image taken from the crack edge.28 (e) SEM image taken from a cracked sample with its axis being 45° to the c-axis.28 (f) DFT calculated fracture toughness using samples cut with various angles to the c-axis direction. Reproduced from ref. 28 with permissions from American Chemical Society Copyright 2024.28 |
Despite the as-mentioned developments, there remain few concerns regarding the accuracies of the above microscopy methods, which are often related to sample preparations and displacement measurements. Take the widely used AFM method as an example; since the cantilever is perpendicular to the basal plane of the 2D materials, the atomic layer that contacts the AFM probe may deviate and slip from the normally aligned atomic structure, resulting in a serious mis-arrangement of the atoms and inhomogeneity in the stress field.34 The other controversy still lies in the uncertainty of the sample quality, i.e. the crystal defects and the contamination introduced during sample preparation, e.g. the polymer adsorbent from polymer-assisted 2D sample transfer, the Pt contamination and Ga ion implantation during Pt deposition in the FIB, and the amorphization and artificial cracking due to ion beam damage on 2D samples in the FIB. On the other hand, the data processing for quantitative measurement needs extra care. For instance, the sample thickness is difficult to precisely measure, either on the SiNx membrane device used for AFM nanoindentation, or on the MEMS chips for electron microscopy. These inaccuracies can arise from the contamination added on the top of the sample surface, or simply due to the fact that the substrate has limited flatness, where angstrom-scale inaccuracies can lead to errors when determining the number of atomic layers in the sample. Also, the sample thickness may change before and after the test, and therefore nominal thickness, calculated from an assumed number of atomic layers based on the measurement before testing, is often used, leading to inaccuracies in the calculation of E3D. Besides, the imaging quality during tests is influencing the accuracy of the final results. Take the in situ MEMS SEM/TEM test as an example; the elastic strain of most 2D materials is within 10%, which means the deformation is on a nanometre scale. However, the recording of the in situ test requires a low magnification imaging for the whole sample area. This causes a limitation in pixel resolution and may not satisfy the accuracy required for the measuring the elastic deformation. Therefore, improvements in measurements are still required, especially on sample preparation and imaging techniques such as polymer-free sample transfer104 and correlated imaging.30,38,111
![]() | ||
Fig. 9 Fracture and deformation mechanisms under tensile stress. (a) Molecular dynamics (MD) modelling of the failure process of a monolayer graphene with a single-vacancy defect, under a strain rate of 109 s−1, showing bond-stretching failure mode.32 (b) MD modelling of the uniaxial tensile fracture processes of a monolayer MoS2 under a strain rate of 109 s−1, loading along zig zag directions, with the left panel showing the result from elastic strain ε = 0.2735, the middle panel showing ε = 0.3297, and the right panel showing ε = 0.3312. Reproduced from ref. 120 with permissions from IOP Publishing Ltd Copyright 2020.120 (c) Schematic illustration of atomic structure evolution during shear deformation of trilayer graphene, with left panel: initial ABA Bernal structure; middle panel: deformed structure with a rigid lattice with the shear process taking place at the top and bottom layers; right panel: relaxed deformation structure, with the formation of partial dislocations as a result of the low stacking fault energy.115 |
![]() | ||
Fig. 10 Deformation mechanisms under bending. (a) Bending in graphene that is subjected to various interlayer interactions. The left top panel shows a model where bending is accommodated by in-plane strain. The left bottom panel shows a model where bending is accommodated through interlayer shear and slip. Experimental STEM image taken from a 12-layer graphene bent to 12° (right panel).87 (b–d) Cross-sectional STEM images demonstrating the different ways for 2D materials to accommodate the strain induced by bending. Scale bars, 5 nm. (b) BF STEM image taken from bilayer graphene, with a 95° bend angle. (c) HAADF STEM image showing the formation of a twin structure in multilayer graphite subjected to bending. (d) HAADF STEM image taken from multilayer graphite subjected to a large bending deformation, constituting areas of discrete twin boundaries (orange) and areas of nanotube-like curvature (green).121 |
![]() | ||
Fig. 11 Theoretical understanding of a defect's impact on mechanical performance. (a) Schematic atomic model of the vacancies in graphene.123 (b) Schematic atomic model of Stone–Thrower–Wales defects in graphene.123 (c) Tensile strength of graphene at failure, as a function of point vacancy, bivacancy and Stone–Wales defects concentrations in graphene, calculated by MD. Reproduced from ref. 123 with permissions from Elsevier Ltd. Copyright 2013.123 (d) Atomic model showing a pair of edge dislocations observed in graphene by TEM. Reproduced from ref. 135 with permissions from The American Association for the Advancement of Science Copyright 2012.135 (e) Atomic model of screw dislocations with a Burgers vector parallel to a zigzag direction, acting as a 2H|2H domain wall in twisted bilayer TMDs, observed by STEM.127 (f) Atomic model of screw dislocation with a Burgers vector parallel to the out-of-plane direction,136 likely to exist in faulty and disordered graphite. Reproduced from ref. 136 with permissions from American Chemical Society Copyright Copyright 2016.136–138 (g) Atomic models of zigzag-oriented grain boundaries (GBs) in graphene with a tilt angle of 5.5° (left panel), 13.2° (middle panel), and 21.7° (right panel).131 (h) Corresponding MD calculated stress–strain curves of zigzag-oriented graphene sheets pulled perpendicular to the GBs. Reproduced from ref. 131 with permissions from The American Association for the Advancement of Science Copyright 2010.131 |
There are two pathways to enhance the strength of the materials: one is to design the arrangement of defects to inhibit dislocation motion or cancel defect effects;130 the other is to go from the opposite: eliminating the defect in crystals using high-quality single-crystal 2D materials. Since it is challenging to prepare large-size single crystal 2D materials for device applications, a usage of polycrystalline 2D crystals is often necessary,131–133 and therefore understanding the role of defects and GBs in mechanical performance is important. There have been theoretical works demonstrating that the detailed arrangement of defects, e.g. the pentagon–heptagon defects associated with graphene GBs, can increase the strength (Fig. 11h),131–133 although it is highly challenging to achieve such a precise defect engineering in materials for real applications.
In contrast to much research interest in theoretical works, the related experimental works on the effects of defect and grain boundaries are of limited quantity. Previous experiments show that presence of GBs provides weak points for oxygen attack111 and impurity segregation, leading to a reduced strength of 2D materials,111 even though careful sample preparation may achieve materials with a strength close to the pristine ones.92 From the perspective of defect types, there are in situ characterization results showing the insensitivity of 2D materials’ stretchability to specific types of defect, as demonstrated in previous sections.53 Regarding the effect of defect density, counter-intuitively, results reported by López-Polín et al. showed that the 2D Young's modulus of graphene increases with an increased density of vacancies created by ion implantation, up to almost twice the initial value when the vacancy content reaches ∼0.2%, as shown in Fig. 12.134 Therefore, controversies still exist regarding the impact of defects on the strength of 2D materials, and more detailed and quantitative studies into defect impacts are needed.
![]() | ||
Fig. 12 Experimental understanding of a defect's impact on elastic modulus. (a) Atomic-resolution scanning tunnelling microscopy (STM) characterization of a pristine graphene, before Ar+ irradiation (left panel), and after irradiation, which shows a single defect containing a vacancy cluster (right panel).134 (b) Raman spectra taken from the sample before and after the irradiation treatment, denoted by blue and red colours respectively. The intensity ratio between the D and G peaks can be used to evaluate to defect density.134 (c) E2D measured by AFM nanoindentation as a function of defect concentration. Reproduced from ref. 134 with permissions from Springer Nature Limited Copyright 2014.134 |
However, fracture mechanics of vdW heterostructures have been quite limited.45,56,139,140 The E2D of the bilayer heterostructure is lower than the sum of the E2D of each layer but comparable to the corresponding bilayers, when a strong interlayer interaction is achieved (Fig. 13). Nevertheless, the interlayer interaction varies with the material components (e.g. MoS2–WS2 interaction > MoS2-graphene), and can only be as strong as a homo-bilayer when the interface is clean or coherent (lattice matching). The MD calculation reported showed that local delamination/buckling can happen when the heterostructure is loaded with a large tensile strain (Fig. 13e).
![]() | ||
Fig. 13 Elastic properties and fracture failure mechanism in vdW heterostructures. (a–d) Experimentally measured elastic properties for CVD-grown MoS2 and WS2 monolayers, and their stacked bilayer heterostructure. (a) Histogram of E2D for CVD MoS2 nanoplates. (b) Histogram of E2D for CVD WS2 triangular nanoplates. (c) Histogram of E2D for a CVD MoS2/WS2 heterostructure. (d) Histogram of E2D for a CVD MoS2/Gr heterostructure. Reproduced from ref. 56 with permissions from American Chemical Society Copyright 2014.56 (e) MD calculated tensile deformation process for a graphene–MoS2–graphene heterostructure, which shows buckling at the ultimate strain ε = 0.26 at 1.0 K. From top to bottom, the tension in the x direction increases. Reproduced from ref. 139 with permissions from AIP Publishing Copyright 2014.139 |
In fact, the interlayer interaction between the stacked components varies with the stacking method and the twist angle.56 In a twisted heterostructure, for a twist angle close to 0° (identical to n × 60° with n being an integer, in hexagonal-symmetry 2D materials), lattice matching induces commensurate lattices at the interface, where the interlayer interaction is stronger141 compared with the incommensurate interface formed at twist angles deviating from n × 60°. Friction experiments prove that superlubricity exists at graphene–hBN heterostructure interfaces for specific twist angles (e.g. 30°, 90°, 150°) where the interface is incommensurate (Fig. 14). This used the other mechanical loading mode of AFM, which is nanofriction with the tip moving parallel to the material surface, and meanwhile the horizontal change caused by the frictional force is recorded by the piezoelectric ceramic transducer, suitable for studying the interface characteristics of 2D materials. In contrast, for those twist angles retaining high symmetry (0°, 60°, 120° etc.), the interface structure can be commensurate and the friction is found to be much higher (Fig. 14).142
![]() | ||
Fig. 14 Measuring the friction at the interface of a graphite/hBN heterostructure. (a) Schematic diagram of the experimental set-up to measure the friction in graphite/hBN junctions.142 (b) Schematic atomic model shows the stress loading onto the graphite/hBN heterostructures during the friction experiments.142 (c) Dependence of the frictional stress on the relative interfacial orientation between monocrystalline graphite and hBN measured under ambient conditions. Reproduced from ref. 142 with permissions from Springer Nature Limited Copyright 2018.142 |
Taking advantage of the superlubricity between heterostructures, Huang et al. investigated how the bending stiffness of 2D heterostructures evolves with the composition of the stack,33 following the bending stiffness measurement work shown in section 2. They fabricated four-layer graphene/MoS2 heterostructures with varied component sequences, including Gr/MoS2/Gr/MoS2 (denoted as GMGM here) and Gr/Gr/MoS2/MoS2 (GGMM) heterostructures (Fig. 15). MGGM, GMMG, and MMGG show a strong bending angle dependence in bending stiffness. In contrast, the bending stiffness of GMGM exhibits no dependence on the bending angle. At high bending angles, the bending stiffnesses of all four structures converge to approximately 20–25 eV. At low bending angles, the measured bending stiffness is much higher for structures with more aligned interfaces, i.e. those containing MM or GG (Fig. 15d). The interfacial friction can be further reduced by large-angle twisting, and the bending stiffness of the resulted heterostructures is largely lower by over several hundred percent compared with other heterostructures. This demonstrates the importance of interfacial engineering in achieving a flexible 2D multilayer, where a minimum bending stiffness can be achieved through misaligning heterointerfaces.
![]() | ||
Fig. 15 Bending of four-layer 2D heterostructures, composed of various orders of graphene (G) and MoS2 (M) layers. (a) Schematic of a heterostructure draped over an atomically sharp step of hBN.33 (b–e) Cross-sectional ADF-STEM images of four different 2D heterostructures (GMGM, MGGM, GMMG, and MMGG) with an identical composition but different stacking orders. Scale bars: 2 nm.33 (f) Plot of bending stiffness for each heterostructure, coloured by bending angle. Reproduced from ref. 33 with permissions from Wiley-VCH GmbH Copyright 2021.33 |
To achieve 2D material composites with enhanced flexibility, recent developments on aligning strategy have enabled the rational design of the spatial alignment of 2D materials. Taking graphene fibre as an instance, previous reports have proved the importance of the alignment of graphene sheets to enhance the mechanical properties. This has been achieved through modifying liquid processing parameters, such as shear flow and enlarged crystal concentration in liquid crystal precursors.153 However, the assembly of sheets remains loose in the transverse direction of the fibre. To improve the order in transverse directions, Gao et al. realized the concentric arrangement of graphene oxide nanosheets instead of a random structure through applying a multiple shear flow field. An increased assembly order is achieved through introducing a rotating angular velocity imposed by rotational shear flow (Fig. 16a).38 Please note that such an angular velocity–assembly order relationship is non-monotonic. Theoretical modelling indicates that when the angular velocity is too high, the excessive centrifugal force makes the radial pressure gradient and viscous force unable to suppress the disturbance in the flow, resulting in a secondary vortex velocity field and a spiral arrangement with defects.38 Indeed, experiments show that a higher or lower angular velocity results in the formation of helical disorder or random disorder, which can be characterized by the Hermans orientation function, accounting for the lower thermal conductivity and Young's modulus.38 Through optimizing the angular velocity, combined with tuning of polymer components in the composites, they achieved an enhanced assembly order in both the longitudinal and transverse directions, and thus synergistically improved and extraordinary mechanical and thermal properties.38 The above study indicates that it is necessary to correlate the theoretical mechanics with the assembly techniques of 2D flakes for rationally designed structural 2D materials with enhanced mechanical and functional properties.
![]() | ||
Fig. 16 (a) Rationally constructed high-strength and thermally conductive graphene oxide fibre. The top panel is a schematic showing the sheet-order in the spinning tube under the unidirectional flow field (denoted as Plane I) and the aligned graphene oxide sheets under the multiple flow fields (Planes II and III). The middle panel displays the cross-sectional views of the velocity distributions, calculated under unidirectional tubular shear-flow field (left), and under multiple shear flow with moderate (middle) and overhigh (right) rotating angular velocities. The bottom left panel displays the velocity distribution across the spinning tube calculated under multiple shear-flow with a rotating angular velocity of 100 rad s−1 (left) and 1000 rad s−1 (right). The bottom middle panel shows the experimentally measured density and orientation order of the graphene fibres. The bottom right panel shows the stress–strain curve measured from the graphene fibres.38 (b) The top left panel is a digital photo taken from the flexible substrate integrated with hundreds of WS2 devices. The top right panel is a schematic showing the strain distribution within the WS2 device under a biaxial strain. The bottom left panel shows the increased mobility as a function of the strain applied to WS2. Bottom right is a DFT calculation of conduction bands from monolayer WS2 structures built without strain (black line), with 1% uniaxial strain (dashed blue line), and 1% biaxial strain (dashed orange line) relative to the lowest band edge. The inset is a schematic of the unit cell structure used for the simulation with applied strain vectors. Reproduced from ref. 167 with permissions from American Chemical Society Copyright 2024.167 (c) Quantum photon emitters fabricated from strain-engineered monolayer and bilayer WSe2. The left panel is an optical image taken after transfer onto the nanopillars. The right panel is a spatial mapping showing the intensity integrated from the as-measured photoluminescence spectrum between 700–860 nm. Reproduced from ref. 157 with permissions from Springer Nature Limited Copyright 2022.157,179 (d) Ferroelectricity due to strain in CuInP2S6. The top left panel shows a schematic of the device structure. The top right panel presents an amplitude map showing the domain wall structure imaged by band excitation piezo-response force microscopy. The bottom panel displays a schematic for the bent nanoflake on the patterned substrate. Reproduced from ref. 172 with permissions from American Chemical Society Copyright 2023.172 |
Deformation has been demonstrated as an established approach to tune the electronic and optical properties of 2D materials, including mobility, photon emitting and ferroelectricity.154–156 Quite a few theories and experiments have found that strain engineering can be utilized to increase the electron mobility of 2D semiconductors, which allows their applications in flexible electronics and sensors.157–166 Recently, Yang et al. developed a force loading approach that enables a biaxial tensile strain in 2D MoS2 and WS2.167 Compared with the case of uniaxial tensile strain, the mobility of WS2 can be much higher in biaxial strain status (Fig. 16b).167 DFT calculations reveal that this resulted from a reduced bandgap as well as a reduced intervalley electron–phonon scattering.167 For optoelectronic applications, single-photon emitters (SPEs) created by strained 2D materials, including hBN, WSe2, WS2, and MoTe2, have attracted much interest in the recent decade.157,159,168–170 This was achieved by suspending 2D semiconductors over arrays of nano-pillars.159 For example, WSe2 SPEs were created through transferring a WSe2 flake on top of lithographically defined nanopillars, where point-like defect or strain perturbations locally change the bandgap and lead to quantum confinement of excitons (Fig. 16c). The performance of the 2D semiconductor SPEs can be modulated by changing the strain applied to the 2D semiconductors. For example, Ferrari et al. show that quantum-light emitters with deterministic positions surpass their randomly distributed counterparts, in which case the spectral wanderings were reduced by an order of magnitude.158 Recently, Chen et al. reported that a large local strain of up to 5% in WSe2 can increase the brightness of the resulting SPEs.168 They further improved the emitting stability through tightly attaching the 2D semiconductor to the surface of the pillar with enhanced fitting.168 On the other hand, curvature and strain effects are also important for ferroelectric 2D materials, such as In2Se3, CuInP2S6, and Bi2TeO5, and various twisted hetero/homostructures.100,118,171–177 Enhanced polarization is expected under a large curvature.171,175 Taking CuInP2S6 as an example, it was observed that ferroelectric domain boundaries tend to form near or move towards the high-curvature areas, and the polarization–voltage hysteresis loops in the bending regions differ from the non-bending regions (Fig. 16d).172 The above studies all indicate that achieving precise strain modulation in 2D materials is a crucial strategy for enhancing their performance in electronic and optical devices.157,167,172,178
![]() | ||
Fig. 17 Strategies and prospects for characterization studies of mechanics in 2D materials. Reproduced from ref. 118 with permissions from American Chemical Society Copyright 2023.30,32,33,38,51,87,104,118,123 |
This journal is © The Royal Society of Chemistry 2025 |