ABSTRACT

The van der Waals (vdW) confined space provides a distinct environment from free space, enabling the production of two-dimensional Janus materials, like highly asymmetric hydrogenated graphene (AH-Gr). Here, we develop a vdW confined space assisted hydrogenation method to produce AH-Gr. The confined space between graphene and the substrate aggregates hydrogen radicals, making the bottom-side of graphene more prone to hydrogenation. The dense and homogeneous confined spaces between adjacent vdW crystals promote rapid and uniform distribution of carbon-hydrogen (C−H) bonds. The hydrogen-to-carbon atomic (H/C) ratios can be quantitatively controlled by adjusting the permeated proton dose. All AH-Gr, regardless of H/C ratios, remain vacancy-free. The spatial distributions of C−H bonds significantly influence the electrical and magnetic properties of AH-Gr. Asymmetric hydrogenation transforms graphene from a semi-metal to a semiconductor, suppresses the quantum Hall effect, and reduces the phase coherence length. This study provides new insights into the preparation and characteristics of hydrogenated graphene, broadening the applications of vdW confined space.

INTRODUCTION

Van der Waals (vdW) confined space, the empty space between adjacent vdW crystals spanning angstroms to nanometers [1], provides a fantastic nano-environment of causing various physical and chemical performances, including the vdW pressure [2,3], frictionless flow [4], untraditional phase transition [5,6], selective permeation [7,8], confined catalysis [9], etc. The narrowest interlayer space restricts the transport of most molecules and ions [10], but they can be expanded by intercalating alkalis and acid molecules under specific conditions [11,12]. The spaces between vdW crystals and the non-vdW substates are also regarded as vdW confined spaces [5], albeit less homogeneous but still prevalent and significantly influential. For example, water or oxygen molecules confined between graphene and SiO2/Si can induce p-type doping of graphene [13].

Graphene, the first exfoliated vdW crystal, holds promise in electronic, photonic and spintronic applications [14]. However, its lack of an intrinsic bandgap restricts its use in transistors and sensors [15]. Bandgaps can be induced by breaking the graphene's structural symmetry through functionalization with various molecules [16,17], such as hydrogen [18–22], fluorine [23], organic radicals [16], et al. Among these, hydrogenation is widely employed due to its accessibility, reversibility and harmlessness [18,21]. Depending on the distribution of carbon-hydrogen (C−H) bonds, hydrogenated graphene can be classified into lowly asymmetric forms (LH-Gr), such as ideal double-side hydrogenated graphene, and highly asymmetric forms (AH-Gr), such as ideal single-side hydrogenated graphene [18,24]. AH-Gr exhibits distinct properties compared to intrinsic graphene [25], including an indirect bandgap [26,27], strong spin-orbit coupling [28,29], spin-triplet exciton [30] and ferromagnetism [31]. Particularly, graphone, a fully hydrogenated form of single-side hydrogenated graphene, is theoretically predicted to manifest many distinctive properties [29,31].

Currently, hydrogenation of graphene typically utilizes hydrogen radicals or protons as reactants, including H2 plasma [18,21], hydrogen atom exposure [19,20,22,32] and hydrogen silsesquioxane dissociation [28]. AH-Gr films have been reported to be produced by treating graphene on SiO2/Si (Gr/SiO2) substrate with hydrogen plasma [18], generating active H radicals on epitaxial graphene under ultra-high vacuum (UHV) [32], or applying hydrogen silsesquioxane followed by electron beam irradiation [28]. These methods introduce magnetic moment asymmetry and significant spin-orbit coupling in hydrogenated graphene [28,32]. However, recent findings indicate that the graphene's hexagonal lattices are permeable to both protons and hydrogen molecules [33–36], complicating the construction of AH-Gr since both sides are exposed to active H radicals or protons. Therefore, achieving AH-Gr with a precise hydrogen-to-carbon atomic ratio (H/C) remains a significant challenge.

The crucial strategy for producing AH-Gr lies in designing differentiated chemical environments on each side of the graphene. Actually, graphene on substrate fulfills this criterion, i.e. free space on the upper side and confined space on the bottom. The vdW confined space between graphene and substrate serves as a nano-sized reactive chamber that influences internal chemical reactions [5,9], enhancing the stability of active sites and improving reaction activities [37]. In this study, we propose a vdW confined space assisted hydrogenation method for precise AH-Gr production. The permeated protons accumulate and recombine into H radicals between graphene and substate, resulting in a higher hydrogen concentration that facilitates hydrogenation at the bottom side. By controlling the stabilization conditions of C−H bonds, the bottom side of graphene can be controllably hydrogenated, while the upper side remains almost non-hydrogenated. Comprehensive characterizations confirm the highly asymmetreic feature and the recoverability of graphene lattices. The degree of hydrogenation in AH-Gr can be precisely controlled by adjusting the dose of permeated protons. The spatial distributions of hydrogen atoms significantly influence electrical and magnetic properties, allowing AH-Gr transformation from semi-metal to semiconductor even at low ppm hydrogenation levels.

RESULTS AND DISCUSSION

AH-Gr in the vdW confined space

Figure 1a illustrates the hydrogenation schematics of Gr/SiO2 via weak hydrogen plasma. At low temperature (LT), active H radicals bond to the upper side of graphene, while generated protons will permeate the graphene and recombine into H radicals, forming C−H bonds on the bottom side, leading to LH-Gr (scenario I). Noting that, the hydrogenation processes are hugely affected by temperature and concentration of H radicals [18,35,36]. As temperature increases, the formed C−H bonds are easily broken, particularly those on the upper side exposed to the free space. Once hydrogenation gets underway at moderate temperature (MT), C−H bonds on the upper side tend to break while those on the bottom side remain relatively stable because of the sub-nanometer confined space and the accumulated high-concentration of H radicals, leading to formation of AH-Gr (scenario II). If graphene is free-standing, both sides remain non-hydrogenated under MT (scenario III), serving as criteria for the hydrogenation temperature range. At high temperature (HT), C−H bonds on the bottom side of graphene also become unstable, preventing hydrogenation (scenario IV). Therefore, the crucial factors for producing AH-Gr in the confined space should be the hydrogenation temperature range under nondestructive plasma conditions.

Production of AH-Gr in the vdW confined space via proton permeation. (a) Schematics of hydrogenating graphene through different treatment processes. I, Gr/SiO2 is exposed to H2 plasma at LT, where protons can permeate graphene and form LH-Gr with the C−H bonds on both sides. II, Gr/SiO2 is exposed to H2 plasma at MT, where C−H bonds are formed mainly on bottom side due to the high-density H radicals within the vdW confined space lead to AH-Gr. III and IV, no C−H bonds are formed for free-standing graphene at MT (III) and Gr/SiO2 at HT (IV). (b, c) Raman spectra of free-standing graphene (b) and Gr/SiO2 (c) hydrogenated at different temperatures for 300 s. The hydrogenation temperature of 350°C is highlighted by light yellow, where Gr/SiO2 is hydrogenated but free-standing graphene is not. Insets show the corresponding optical image of free-standing graphene on Cu grid (b) and AFM image of Gr/SiO2 (c). (d) Raman spectra of hydrogenating Gr/SiO2 at 350°C with different durations. The intensity of D peak increases with prolonging the proton permeation time. (e) Normalized Raman ID/IG ratios of free-standing graphene and Gr/SiO2 hydrogenated at different temperatures. The ratios of ID/IG correspond to the relative concentration of C−H bonds, and LT is thereby defined as the temperatures below 350°C, MT is 350–425°C labelled by light yellow, and HT is above 425°C. (f, h) Typical STM topographical images of AH-Gr/SiO2 with ID/IG of 0.2 (f) and 0.8 (h). There are no obvious atomic defects or hydrogen atoms chemisorbed on the upper side. (g, i) Corresponding STS spectra of (f) and (h) collected at the colored dots. The STS spectrum of pristine graphene is shown as the blue line in (g) for comparison. (j) Distribution histogram of STS bandgaps from (g) and (i). Their bandgaps range from 0 to 30 meV for AH-Gr with ID/IG of ∼0.2 (g) and 120–430 meV for AH-Gr with ID/IG of ∼0.8 (i).
Figure 1.

Production of AH-Gr in the vdW confined space via proton permeation. (a) Schematics of hydrogenating graphene through different treatment processes. I, Gr/SiO2 is exposed to H2 plasma at LT, where protons can permeate graphene and form LH-Gr with the C−H bonds on both sides. II, Gr/SiO2 is exposed to H2 plasma at MT, where C−H bonds are formed mainly on bottom side due to the high-density H radicals within the vdW confined space lead to AH-Gr. III and IV, no C−H bonds are formed for free-standing graphene at MT (III) and Gr/SiO2 at HT (IV). (b, c) Raman spectra of free-standing graphene (b) and Gr/SiO2 (c) hydrogenated at different temperatures for 300 s. The hydrogenation temperature of 350°C is highlighted by light yellow, where Gr/SiO2 is hydrogenated but free-standing graphene is not. Insets show the corresponding optical image of free-standing graphene on Cu grid (b) and AFM image of Gr/SiO2 (c). (d) Raman spectra of hydrogenating Gr/SiO2 at 350°C with different durations. The intensity of D peak increases with prolonging the proton permeation time. (e) Normalized Raman ID/IG ratios of free-standing graphene and Gr/SiO2 hydrogenated at different temperatures. The ratios of ID/IG correspond to the relative concentration of C−H bonds, and LT is thereby defined as the temperatures below 350°C, MT is 350–425°C labelled by light yellow, and HT is above 425°C. (f, h) Typical STM topographical images of AH-Gr/SiO2 with ID/IG of 0.2 (f) and 0.8 (h). There are no obvious atomic defects or hydrogen atoms chemisorbed on the upper side. (g, i) Corresponding STS spectra of (f) and (h) collected at the colored dots. The STS spectrum of pristine graphene is shown as the blue line in (g) for comparison. (j) Distribution histogram of STS bandgaps from (g) and (i). Their bandgaps range from 0 to 30 meV for AH-Gr with ID/IG of ∼0.2 (g) and 120–430 meV for AH-Gr with ID/IG of ∼0.8 (i).

Raman spectroscopy is used to characterize C−H bonds in graphene, and the D to G peak intensity ratios (ID/IG) serve as indicator of non-sp2 bonds [38,39]. Figure 1b and c show Raman spectra of free-standing graphene on Cu grid and Gr/SiO2, respectively, after undergoing hydrogenation processes at different temperatures. At LT, both exhibit high ID/IG ratios. The reversibility of ID/IG indicates that the D peaks originate from the formed C−H bonds (Fig. S1a and b). As temperature rises, the ID/IG ratios are both reduced, indicating that hydrogenation is suppressed by the elevated temperature. Notably, the D peak in free-standing graphene completely disappears at 350°C, but it still exists in Gr/SiO2. Additionally, with prolonged treatment, the D peak in Gr/SiO2 becomes more pronounced, while the free-standing graphene remains at a state of basically no-D peak (Fig. 1d and Fig. S1c). The absence of a D peak indicates that free-standing graphene is not hydrogenated. Both sides of the free-standing graphene are exposed to free space, there are no C−H bonds existing on either side. In contrast, the upper side of Gr/SiO2 also has free space and should be free of C−H bonds, hence the D peaks should mainly originate from the C−H bonds on the bottom side, i.e. AH-Gr. As temperature exceeds 425°C, D peaks are negligible, indicating no C−H bonds formed. Figure 1e summarizes the normalized ID/IG ratios of free-standing graphene and Gr/SiO2 hydrogenated at different temperatures (details in Fig. S1d and e). After undergoing the hydrogenation process at 350–425°C, only Gr/SiO2 can be hydrogenated indicated by the distinguishable ID, while free-standing graphene shows almost no C−H bonds. Consequently, we define the temperature ranges of hydrogenation as LT of 25–350°C for LH-Gr, where the asymmetry of hydrogenated graphene gradually increases with the increase in temperature, MT of 350–425°C and HT of >425°C for AH-Gr and non-hydrogenation, respectively.

Further studies using a scanning tunnelling microscope (STM) show the morphology of AH-Gr with ID/IG ratios of ∼0.2 and ∼0.8, revealing perfect hexagonal lattices, with no upper-side C−H bonds, as shown in Fig. 1f and h. The undamaged lattices without C−H domains further confirm its highly asymmetric feature, and the D peaks should mainly derive from bottom-side C−H bonds. Figure S2 shows typical STM images of graphene hydrogenated at LT, showing numerous upper-side C−H domains, which is the same as the reported surface topography of hydrogenated graphene hydrogenated at LT [19,20,22]. The C−H bonds can be also completely removed by UHV annealing. Notably, the plasma utilized for hydrogenation is too weak to introduce atomic vacancies, as confirmed by the recovered lattices and our reported results [13,35,36]. Moreover, scanning tunnelling spectroscopy (STS) (Fig. 1g and i) show the changed band structure of AH-Gr, with opened bandgaps of 0–30 meV and 120–430 meV for ID/IG of ∼0.2 and ∼0.8, respectively. The STS spectrum of pristine graphene is shown as the blue line in Fig. 1g for comparison. Figure 1j summarizes the bandgap distributions from dozens of spectra (Fig. S3), indicating the changed band structures and probable derivative properties.

Furthermore, we compared the wettability of pristine graphene, LH-Gr and AH-Gr through wetting angle measurements, as shown in Fig. S4. For pristine graphene, the wetting angle is ∼86°. After hydrogenation into LH-Gr (ID/IG = 3.7), the angle decreases to ∼60°, due to increased surface energy [40]. In contrast, AH-Gr consistently exhibits lower wetting angles and higher surface energy compared to LH-Gr with similar ID/IG ratios. Additionally, we further investigated the transformation between LH-Gr and AH-Gr using ex situ wetting angle measurements. Starting with AH-Gr (ID/IG = 2.4), LT hydrogenation increased the wetting angle from 42° to 57° (ID/IG = 2.6), and then to 73° (ID/IG = 3.5), indicating higher hydrogenation and changes in C−H bond distribution (LH-Gr). The annealing process further transforms LH-Gr back towards AH-Gr, decreasing the wetting angle from 73° to 53° (ID/IG = 2.7). These results demonstrate that hydrogenation temperature can tune the spatial distribution of C−H bonds.

Utilization factor of permeated protons for AH-Gr

Next, we estimate the utilization factor of aggregated H radicals for hydrogenating AH-Gr. First, we fabricate a double-layer graphene film consisting of top 13C-graphene and bottom 12C-graphene, with distinguishable Raman G and D peaks. Subsequently, this double-layer graphene is exposed to H2 plasma at MT. The protons permeate into the vdW confined space between 13C-graphene and 12C-graphene, re-combined H radicals will mainly bond to the bottom side of the top 13C-graphene and the upper side of bottom 12C-graphene (Fig. 2a). Meanwhile, a few protons permeate 12C-graphene and form 12C−H bonds at the bottom side. Their Raman spectra with different hydrogenation durations are plotted in Fig. 2b, and the isotopic 13C-graphene presents the redshifted 13G, 13D and 132D peaks. Although the 13G intensity is 1/3 weaker than 12G, the 13D is still twofold stronger than 12D. Figure2c plots the relative ratio for 13C−H and 12C−H bonds, which is normalized by the ratio of I13D/I13G to I12D/I12G. We find that the concentration of 13C−H is ∼80%, representing the utilization factor for the bottom side hydrogenation of top AH-Gr.

Depth distribution of C−H bonds confirmed through isotopic effects. (a) Schematic of hydrogenating double-layer Gr/SiO2, where top and bottom graphene films are constituted by 13C and 12C, respectively. (b) Raman spectra of double-layer graphene hydrogenated at MT for different durations. The 13C−H and 12C−H bonds cause the different dividing D peaks, where 13D from 13C−H is located at 1308 cm−1 and 12D from 12C−H is at 1355 cm−1. (c) Distribution of C−H bonds on different sides of graphene, estimated from the relative Raman intensity of the D peak. The proportion of C−H bonds at the bottom side of top graphene reaches ∼80% (A2), where the upper side of top graphene (A1) is negligible. (d) Schematic of hydrogenating triple-layer Gr/SiO2, where top, middle and bottom graphene films are constituted by 13C, 13C and 12C, respectively. (e) Raman spectra of triple-layer graphene hydrogenated at MT for different durations. The 12D peak from the bottom graphene is very weak. (f) Depth distribution of C−H bonds at different sides. The proportion of C−H bonds at bottom graphene is ∼6.7%. The ratios of C−H bonds at top, middle and bottom graphene are ∼12:2:1.
Figure 2.

Depth distribution of C−H bonds confirmed through isotopic effects. (a) Schematic of hydrogenating double-layer Gr/SiO2, where top and bottom graphene films are constituted by 13C and 12C, respectively. (b) Raman spectra of double-layer graphene hydrogenated at MT for different durations. The 13C−H and 12C−H bonds cause the different dividing D peaks, where 13D from 13C−H is located at 1308 cm−1 and 12D from 12C−H is at 1355 cm−1. (c) Distribution of C−H bonds on different sides of graphene, estimated from the relative Raman intensity of the D peak. The proportion of C−H bonds at the bottom side of top graphene reaches ∼80% (A2), where the upper side of top graphene (A1) is negligible. (d) Schematic of hydrogenating triple-layer Gr/SiO2, where top, middle and bottom graphene films are constituted by 13C, 13C and 12C, respectively. (e) Raman spectra of triple-layer graphene hydrogenated at MT for different durations. The 12D peak from the bottom graphene is very weak. (f) Depth distribution of C−H bonds at different sides. The proportion of C−H bonds at bottom graphene is ∼6.7%. The ratios of C−H bonds at top, middle and bottom graphene are ∼12:2:1.

Additionally, we fabricate a triple-layer film consisting of top 13C-graphene, middle 13C-graphene and bottom 12C-graphene. After MT hydrogenation for different durations, we use the 13D and 12D peaks to estimate the hydrogenation depth through permeating protons, as illustrated in Fig. 2d. Figure 2e displays the Raman spectra of the triple-layer film following varying hydrogenation durations. The weak intensity of 12D peak indicates minimal hydrogenation in the bottom 12C-graphene, with the concentration of 12C−H estimated to be <6.7% (Fig. 2f). The proton permeation depth here is a single or double-layer graphene lattice, consistent with the H2 bubble encapsulated by single or double-layer graphene [36].

Quantification of H/C ratios in AH-Gr

We utilize exfoliated hBN as a denser substrate to create homogeneous vdW confined spaces, where recombined H radicals are assumed to efficiently form C−H bonds [35,36,41,42]. Figure 3a illustrates the schematic of AH-Gr on hBN, the distance of vdW confined space is typically 0.34–0.65 nm for various 2D materials [10,33,43,44]. After transferring monolayer graphene onto few-layer hBN [45], we perform hydrogenation, Gr/hBN exhibits a significantly enhanced hydrogenation effect compared to Gr/SiO2, and with an extended temperature threshold for AH-Gr up to 460°C (Fig. S5).

Quantification of H/C ratios in AH-Gr. (a) Schematic of hydrogenating graphene in the vdW confined space between graphene and hBN. The confined space is dense so wastes few active H radicals for hydrogenation. (b) Raman spectra of Gr/hBN with different hydrogenation time periods at MT, showing a rapid intensity increase of D peak. (c) Correlation between Raman ID/IG and H/C ratios. Insets are the optical images of graphite flakes before (top) and after (bottom) proton permeation at HT under the same proton conditions. The circular dots in graphite are formed by H2 bubbles in the permeated protons. H2 bubbles formed in graphene lattices prolong the permeation duration at HT, but there is no bubble formed at MT. (d) Correlation between H/C ratios and hydrogenation time for graphene on different substrates. The H/C ratio of Gr/hBN increases more rapidly than that of Gr/SiO2. (e) Statistic histograms of Raman ID/IG for Gr/SiO2 (top) and Gr/hBN (bottom) hydrogenated at MT. AH-Gr on hBN shows more homogeneous distribution for H/C ratio. Insets are their corresponding optical images before undergoing hydrogenation. (f) XPS of AH-Gr film on SiO2/Si with H/C ratio of 160 ppm. The sub-peak indicates the existence of sp3 C−H component.
Figure 3.

Quantification of H/C ratios in AH-Gr. (a) Schematic of hydrogenating graphene in the vdW confined space between graphene and hBN. The confined space is dense so wastes few active H radicals for hydrogenation. (b) Raman spectra of Gr/hBN with different hydrogenation time periods at MT, showing a rapid intensity increase of D peak. (c) Correlation between Raman ID/IG and H/C ratios. Insets are the optical images of graphite flakes before (top) and after (bottom) proton permeation at HT under the same proton conditions. The circular dots in graphite are formed by H2 bubbles in the permeated protons. H2 bubbles formed in graphene lattices prolong the permeation duration at HT, but there is no bubble formed at MT. (d) Correlation between H/C ratios and hydrogenation time for graphene on different substrates. The H/C ratio of Gr/hBN increases more rapidly than that of Gr/SiO2. (e) Statistic histograms of Raman ID/IG for Gr/SiO2 (top) and Gr/hBN (bottom) hydrogenated at MT. AH-Gr on hBN shows more homogeneous distribution for H/C ratio. Insets are their corresponding optical images before undergoing hydrogenation. (f) XPS of AH-Gr film on SiO2/Si with H/C ratio of 160 ppm. The sub-peak indicates the existence of sp3 C−H component.

To quantitatively estimate the H/C ratios of AH-Gr, we establish an analogy approach. Given that hBN and graphene exhibit lower permeability to protons compared to amorphous SiO2 [33], ∼80% of the permeated protons are utilized for generating AH-Gr. This allows us to calculate the C−H ratios via the dose of the permeated protons following the equation H/C = ΓP  × τ/nC, where ΓP is the permeating rate of proton through graphene and is estimated to be ∼940 s−1μm−2 under the same plasma condition [36], τ is the proton permeation time, and nC is the carbon atom density in graphene of 3.82 × 1019 m−2.

Figure 3b and c plot the typical Raman spectra of Gr/hBN after different permeation times and the extracted ID/IG versus calculated H/C ratios, respectively. After 5 s of hydrogenation, the ID/IG ratio reaches 0.6, corresponding to an H/C ratio of 98 ppm, with the average distance between neighboring C−H bonds (LH) being ∼16.3 nm. This distance aligns closely with Raman analysis [38], but is slightly larger than that established by e-beam irradiated HSQ [28]. We then compare the correlation between H/C ratios and hydrogenation duration for Gr/hBN and Gr/SiO2. Notably, AH-Gr on both substrates with the same H/C ratios exhibit identical ID/IG, confirmed by transferring the same AH-Gr films to SiO2/Si and hBN (Fig. S6). This indicates that Raman ID/IG for AH-Gr only derives from H/C ratios, and are irrelevant to the substrate. Figure 3d shows that the average hydrogenation rate is ∼20 ppm/s for Gr/hBN, significantly higher than the 0.2 ppm/s for Gr/SiO2. The two orders magnitude difference indicating a narrower space of Gr/hBN with denser H radicals. The higher ID/IG ratio for AH-Gr/SiO2 requires a longer hydrogenation duration (Fig. 1e). Additionally, the vdW confined space enhances the homogeneity of AH-Gr, as confirmed in Fig. S7. Figure 3e plots the ID/IG spatial distributions of AH-Gr/SiO2 and AH-Gr/hBN. Although both exhibit similar H/C ratios with an average ID/IG ratio of 0.67, the distribution for Gr/SiO2 is notably broader, with ∼90% of counts between 0.52 and 0.79, whereas Gr/hBN is more homogeneous with ∼90% of counts between 0.55 and 0.76.

X-ray photoelectron spectroscopy (XPS) is employed to analyze the elemental composition and chemical states. Figure 3f shows the XPS spectra of AH-Gr with H/C ratio of 160 ppm, with distinct subpeaks for different components. The C−H bonds (sp3-1) peak is located at 285.0 eV, C−C bonds (sp2) of graphene at 284.5 eV. Through peak differentiation, the C−H (sp3-1) component is extracted to estimate the H/C ratio. The XPS spectra of as-transfer graphene are shown in Fig. S8, where no C−H (sp3-1) peak is visible. Additional component peaks, like C−C (sp3-2) at 285.9 eV, C−O at 286.7 eV and O=C−O at 288.9 eV correspond to the quaternary carbon, methoxy group and carboxyl group, respectively, which may originate from the residual polymethyl methacrylate (PMMA) [46].

Electrical and magnetic properties of AH-Gr with different H/C ratios

To investigate the impact of hydrogenated graphene with varying C−H bond distributions, we perform electrical transport measurements on films hydrogenated at different temperatures, all with an identical Raman ID/IG ratio of ∼0.5. Figure 4a and Fig. S9a plot gate voltage (Vg) dependent four-probe longitudinal resistances (Rxx). All the electrical transport curves remain bipolar, the maximum Rxx at charge neutrality point (CNP) increasing with hydrogenation temperature, with values of 7.5, 14, 31,104 and 107 kΩ for 25,150, 200,300 and 370°C, respectively. This indicates a transformation of samples from LH-Gr to AH-Gr. Notably, graphene films hydrogenated at 300°C have similar resistance to AH-Gr prepared at 370°C, indicating that hydrogenation at 300°C approximates AH-Gr.

Electrical performances of AH-Gr films. (a) Rxx of the hydrogenated Gr/SiO2 as a function of Vg − VCNP. The films are hydrogenated at different temperatures but with the same Raman ID/IG of 0.5. Inset is the typical optical image of a hydrogenated graphene Hall device. (b) Rxx of AH-Gr with different H/C ratios as a function of Vg—VCNP. (c) Statistical distribution of Rxx of AH-Gr with various H/C ratios. Rxx values are collected from the CNP (green), light doped state (yellow, VCNP + 30 V) and heavy doped state (red, VCNP + 50 V). The resistances increase sharply among the weak hydrogenation range, highlighted by light red. (d) Statistics of carrier mobility of AH-Gr with different H/C ratios. AH-Gr films with low H/C ratios behave like semi-metals while maintaining high mobility, highlighted by light red. Inset is the schematic of the LH, which is ∼21 nm for AH-Gr with an H/C ratio of 40 ppm. (e) Temperature-dependent Rxx of AH-Gr with different H/C ratios. Rxx is collected at CNP. Most AH-Gr films behave like a typical semiconductor. (f) Extracted Eg of AH-Gr and LH-Gr with different H/C ratios. LH-Gr is prepared at 200°C [47]; the dashed line is a guideline to the trend. Inset is the fitted Eg from the temperature-dependent Rxx.
Figure 4.

Electrical performances of AH-Gr films. (a) Rxx of the hydrogenated Gr/SiO2 as a function of VgVCNP. The films are hydrogenated at different temperatures but with the same Raman ID/IG of 0.5. Inset is the typical optical image of a hydrogenated graphene Hall device. (b) Rxx of AH-Gr with different H/C ratios as a function of VgVCNP. (c) Statistical distribution of Rxx of AH-Gr with various H/C ratios. Rxx values are collected from the CNP (green), light doped state (yellow, VCNP + 30 V) and heavy doped state (red, VCNP + 50 V). The resistances increase sharply among the weak hydrogenation range, highlighted by light red. (d) Statistics of carrier mobility of AH-Gr with different H/C ratios. AH-Gr films with low H/C ratios behave like semi-metals while maintaining high mobility, highlighted by light red. Inset is the schematic of the LH, which is ∼21 nm for AH-Gr with an H/C ratio of 40 ppm. (e) Temperature-dependent Rxx of AH-Gr with different H/C ratios. Rxx is collected at CNP. Most AH-Gr films behave like a typical semiconductor. (f) Extracted Eg of AH-Gr and LH-Gr with different H/C ratios. LH-Gr is prepared at 200°C [47]; the dashed line is a guideline to the trend. Inset is the fitted Eg from the temperature-dependent Rxx.

Figure 4b shows that Rxx of AH-Gr increases with H/C ratios. At H/C ratio of 175 ppm, Rxx is 216 kΩ at CNP, about 20-fold higher than pristine graphene. This substantial increase should be attributed to the C−H scattering sites and the enhanced inversion asymmetry. Comparatively, for AH-Gr with 34 ppm H/C ratios, Rxx demonstrates a slight increase of 15%, indicating weak hydrogenation, where partial graphene exhibits the properties of a semiconductor. Figure 4c presents Rxx statistics of AH-Gr with various H/C ratios at for CNP (green circles), moderate doping (yellow circles, VCNP+30 V) and heavy doping (red circles, VCNP+50 V). Generally, Rxx at CNP is ∼2 orders of magnitude higher than under both moderate doping and heavy doping. Moreover, for an H/C ratio of 474 ppm, Rxx at CNP reaches 3.4 MΩ, ∼300 times higher than pristine graphene. A significant trend shows that Rxx rapidly increases from 0 to 40 ppm H/C ratio, indicating a weak C/H region with coexisting semi-metal and semiconductor regions. Furthermore, higher ratios indicate bandgap opening in all carbon lattices.

Next, we calculate the carrier mobility of AH-Gr with different H/G ratios, summarized in Fig. 4d. Carrier mobilities decrease with increasing H/C ratio, consistent with the reduced conductivity. It drops sharply from ∼12 000 to ∼2000 cm2V−1s−1 within the weak hydrogenation degree, then gradually declines. The AH-Gr with an H/C ratio of ∼174 ppm (STS-measured bandgap of 240 meV) maintains high mobility ∼1000 cm2V−1s−1, while with ∼480 ppm, mobility decreases to ∼40 cm2V−1s−1. Notably, the average distance between the neighboring C−H bonds (LH) is calculated to be ∼21 nm for AH-Gr with an H/C ratio of 40 ppm, which may be the sparsest distance of hydrogen distribution that can open a bandgap for surrounding carbon lattices.

Figure 4e shows the temperature-dependent resistance (R-T) measurements of AH-Gr. For AH-Gr with weak H/C ratio of 6 ppm, Rxx at CNP drops from 110 to 1.5 K, then increases from 110 to 295 K, similar to pristine graphene (Fig. S9b–e). For higher H/C ratios (54–278 ppm), Rxx decreases with rising temperature, indicating typical semiconductor behavior. We further fit their bandgaps from the Arrhenius equation Rxx = exp[Eg/(2kBT)], where Eg is the bandgap, kB is the Boltzmann constant and T is the temperature [21], as plotted in Fig. 4f, marked by blue dots, with the dashed line being a guideline to the trend. The fitted Eg is ∼1 order magnitude smaller than the STS value, likely due to the thermally excited carriers and substrate-induced pseudo-bandgap [48]. By comparing LH-Gr hydrogenated at 200°C, as redrawn in Fig. 4f by green dots, AH-Gr opens a larger bandgap while they have the same hydrogenation concentration. The R-T measurements of LH-Gr are shown in Fig. S9f.

Magnetotransport measurements of AH-Gr films with varying C−H bonds monitor the evolution changes of the quantum Hall effect (QHE) [49]. Figure 5a and b plot Rxx and Hall conductivity (σxy) of AH-Gr films under out-of-plane magnetic field (B) of 7 T at 1.5 K. The Rxx in as-transferred graphene film fall to zero, and σxy reach ± 2 e2/h plateaus near Vg = ±7 V, corresponding to Landau filling factors ν = ±2. Meanwhile, only AH-Gr with weak H/C ratio (34 ppm) shows zero Rxx, and σxy remains plateau shaped at ν = ±6. For other AH-Gr films, Rxx cannot reach zero and σxy plateaus mismatch filling factors, because of the hydrogen-induced broadening of Landau levels [49]. Additionally, the σxy of AH-Gr with H/C ratios ranging from 84–175 ppm show slight slopes near CNP, corresponding to the Rxy peaks (Fig. S9g). These dramatically changed peaks are expected to originate from the disruption of topological edge states induced by C−H bond-induced disorders at CNP [50]. Moreover, weak H/C ratio AH-Gr (34 ppm) still shows topological edge states, indicating interconnections between semimetal and semiconductor regions.

Magnetotransport measurement of AH-Gr on SiO2/Si. (a, b) Rxx and σxy of the AH-Gr with different H/C ratios as a function of Vg—VCNP at 1.5 K under B┴ of 7 T. The horizontal dashed lines in (b) are the guidelines of the Hall plateaus. The Hall plateaus at half-integer are becoming deviated as the H/C ratios increase. Inset of (b) shows the enlarged region of σxy plateaus between −2 and 2. (c) The change rates of MR with B⊥. The MR of AH-Gr with different H/C ratios is collected at CNP and 1.5 K. The results are fitted with solid lines by the WL theory. (d) Relationship between Lφ and LH, which are approximately fitted by the power function of Lφ = LH1.5.
Figure 5.

Magnetotransport measurement of AH-Gr on SiO2/Si. (a, b) Rxx and σxy of the AH-Gr with different H/C ratios as a function of VgVCNP at 1.5 K under B of 7 T. The horizontal dashed lines in (b) are the guidelines of the Hall plateaus. The Hall plateaus at half-integer are becoming deviated as the H/C ratios increase. Inset of (b) shows the enlarged region of σxy plateaus between −2 and 2. (c) The change rates of MR with B. The MR of AH-Gr with different H/C ratios is collected at CNP and 1.5 K. The results are fitted with solid lines by the WL theory. (d) Relationship between Lφ and LH, which are approximately fitted by the power function of Lφ = LH1.5.

Magnetoresistance (MR) of AH-Gr at CNP with different H/C ratios is performed under weak B at 1.5 K. Figure 5c shows that MR decreases with enlarged B, and the change rates of MR with B, that is (ΔR)/R = [R(B)—R(B = 0)/R(B = 0)]. Using weak localization (WL) theory [51], we fitted MR to obtain a phase coherence length (Lφ) of AH-Gr, with relationship to LH plotted in Fig. 5d. Lφ of AH-Gr increases from 34 to 233 nm as LH increases from 13 to 41 nm, and they have an approximate power function dependence of LφLH1.5. The reason is that C−H bonds can act as scattering sites causing charge carrier de-coherence [52].

CONCLUSIONS

In summary, we develop a confined space assisted production of AH-Gr via proton permeation at an MT of 350–425°C. LT hydrogenation leads to LH-Gr, while HT terminates hydrogenation. AH-Gr films are defect-free with opened bandgaps. HT annealing can remove all the formed C−H bonds, recovering their hexagonal lattices. The H/C ratios can be quantitatively controlled by calculating the permeated protons in the vdW confined space between graphene and hBN. AH-Gr on hBN shows a higher hydrogenation rate of 20 ppm/s, which is 2 orders of magnitude greater than on amorphous SiO2. Due to larger structural asymmetry, AH-Gr exhibits higher resistance and larger bandgap than LH-Gr at similar hydrogenation degree. When H/C ratios exceed 40 ppm, AH-Gr changes from semi-metal to semiconductor. Magnetotransport measurements indicate that hydrogenation-induced disruption affects the QHE and reduces the Lφ. This study re-examines the production and features of hydrogenated graphene, expands the applications of vdW confined space, and presents a quantitative hydrogenation approach for AH-Gr with controlled spatial distribution.

FUNDING

This work was supported by the National Natural Science Foundation of China (52425203, 12104218, 11904163, 12374183 and 92165205), the Natural Science Foundation of Jiangsu Province (BK20240008), the National Key Research and Development Program of China (2021YFA1400403), the Innovation Program for Quantum Science and Technology (2021ZD0302800), the China National Postdoctoral Program for Innovative Talents (BX2021120), the Xiaomi Foundation, the Postdoctoral Fellowship Program of China Postdoctoral Science Foundation (CPSF) (GZC20231093), the Jiangsu Funding Program for Excellent Postdoctoral Talent (2023ZB553) and the Program A for outstanding PhD Candidates of Nanjing University (202401A04).

AUTHOR CONTRIBUTIONS

L.G. and G.Y. designed and supervised the experiments. X.H. performed hydrogenation, AFM and Raman. W.L. and X.H. performed the device fabrication and transport measurements. H.Z. and J.X. assisted in graphene growth and transfer, hydrogenation, AFM and Raman. K.W. and L.W. assisted in fabrication of exfoliated Gr/hBN heterostructures. L.Z. and S.L. performed the STM and STS. L.G., G.Y. and X.H. analyzed the data and wrote the manuscript, J.X., L.W. and S.L. revised it, and all authors commented on it.

Conflict of interest statement. None declared.

REFERENCES

1.

Geim
 
AK
.
Exploring two-dimensional empty space
.
Nano Lett
 
2021
;
21
:
6356
8
.

2.

Vasu
 
KS
,
Prestat
 
E
,
Abraham
 
J
 et al.  
Van der Waals pressure and its effect on trapped interlayer molecules
.
Nat Commun
 
2016
;
7
:
12168
.

3.

Hu
 
C
,
Chen
 
JJ
,
Zhou
 
XL
 et al.  
Collapse of carbon nanotubes due to local high-pressure from van der Waals encapsulation
.
Nat Commun
 
2024
;
15
:
3486
.

4.

Keerthi
 
A
,
Geim
 
AK
,
Janardanan
 
A
 et al.  
Ballistic molecular transport through two-dimensional channels
.
Nature
 
2018
;
558
:
420
4
.

5.

Thi
 
QH
,
Man
 
P
,
Liu
 
HJ
 et al.  
Ultrahigh lubricity between two-dimensional ice and two-dimensional atomic layers
.
Nano Lett
 
2023
;
23
:
1379
85
.

6.

Yang
 
Q
,
Sun
 
PZ
,
Fumagalli
 
L
 et al.  
Capillary condensation under atomic-scale confinement
.
Nature
 
2020
;
588
:
250
3
.

7.

Esfandiar
 
A
,
Radha
 
B
,
Wang
 
FC
 et al.  
Size effect in ion transport through angstrom-scale slits
.
Science
 
2017
;
358
:
511
3
.

8.

Mouterde
 
T
,
Keerthi
 
A
,
Poggioli
 
AR
 et al.  
Molecular streaming and its voltage control in angstrom-scale channels
.
Nature
 
2019
;
567
:
87
90
.

9.

Li
 
HB
,
Xiao
 
JP
,
Fu
 
Q
 et al.  
Confined catalysis under two-dimensional materials
.
Proc Natl Acad Sci USA
 
2017
;
114
:
5930
4
.

10.

An
 
Y
,
Kuc
 
A
,
Petkov
 
P
 et al.  
On the chemistry and diffusion of hydrogen in the interstitial space of layered crystals h-BN, MoS2, and graphite
.
Small
 
2019
;
15
:
1901722
.

11.

Dresselhaus
 
MS
,
Dresselhaus
 
G
.
Intercalation compounds of graphite
.
Adv Phys
 
1981
;
30
:
139
326
.

12.

Kühne
 
M
,
Paolucci
 
F
,
Popovic
 
J
 et al.  
Ultrafast lithium diffusion in bilayer graphene
.
Nat Nanotechnol
 
2017
;
12
:
895
900
.

13.

Yuan
 
GW
,
Liu
 
WL
,
Huang
 
XL
 et al.  
Stacking transfer of wafer-scale graphene-based van der Waals superlattices
.
Nat Commun
 
2023
;
14
:
5457
.

14.

Geim
 
AK
,
Novoselov
 
KS
.
The rise of graphene
.
Nat Mater
 
2007
;
6
:
183
91
.

15.

Schwierz
 
F
.
Graphene transistors
.
Nat Nanotechnol
 
2010
;
5
:
487
6
.

16.

Georgakilas
 
V
,
Otyepka
 
M
,
Bourlinos
 
AB
 et al.  
Functionalization of graphene: covalent and non-covalent approaches, derivatives and applications
.
Chem Rev
 
2012
;
112
:
6156
214
.

17.

Zhang
 
LM
,
Yu
 
JW
,
Yang
 
MM
 et al.  
Janus graphene from asymmetric two-dimensional chemistry
.
Nat Commun
 
2013
;
4
:
1443
.

18.

Elias
 
DC
,
Nair
 
RR
,
Mohiuddin
 
TMG
 et al.  
Control of graphene's properties by reversible hydrogenation: evidence for graphane
.
Science
 
2009
;
323
:
610
3
.

19.

Balog
 
R
,
Jørgensen
 
B
,
Nilsson
 
L
 et al.  
Bandgap opening in graphene induced by patterned hydrogen adsorption
.
Nat Mater
 
2010
;
9
:
315
9
.

20.

Haberer
 
D
,
Giusca
 
CE
,
Wang
 
Y
 et al.  
Evidence for a new two-dimensional C4H-type polymer based on hydrogenated graphene
.
Adv Mater
 
2011
;
23
:
4497
503
.

21.

Son
 
J
,
Lee
 
S
,
Kim
 
SJ
 et al.  
Hydrogenated monolayer graphene with reversible and tunable wide band gap and its field-effect transistor
.
Nat Commun
 
2016
;
7
:
13261
.

22.

Chen
 
H
,
Bao
 
DL
,
Wang
 
DF
 et al.  
Fabrication of millimeter-scale, single-crystal one-third-hydrogenated graphene with anisotropic electronic properties
.
Adv Mater
 
2018
;
30
:
1801838
.

23.

Nair
 
RR
,
Ren
 
WC
,
Jalil
 
R
 et al.  
Fluorographene: a two-dimensional counterpart of Teflon
.
Small
 
2010
;
6
:
2877
84
.

24.

Xiang
 
HJ
,
Kan
 
EJ
,
Wei
 
SH
 et al.  
Thermodynamically stable single-side hydrogenated graphene
.
Phys Rev B
 
2010
;
82
:
165425
.

25.

Pujari
 
BS
,
Gusarov
 
S
,
Brett
 
M
 et al.  
Single-side-hydrogenated graphene: density functional theory predictions
.
Phys Rev B
 
2011
;
84
:
041402
.

26.

Haberer
 
D
,
Vyalikh
 
DV
,
Taioli
 
S
 et al.  
Tunable band gap in hydrogenated quasi-free-standing graphene
.
Nano Lett
 
2010
;
10
:
3360
6
.

27.

Kharche
 
N
,
Nayak
 
SK
.
Quasiparticle band gap engineering of graphene and graphone on hexagonal boron nitride substrate
.
Nano Lett
 
2011
;
11
:
5274
8
.

28.

Balakrishnan
 
J
,
Koon
 
GKW
,
Jaiswal
 
M
 et al.  
Colossal enhancement of spin-orbit coupling in weakly hydrogenated graphene
.
Nat Phys
 
2013
;
9
:
284
7
.

29.

Gmitra
 
M
,
Kochan
 
D
,
Fabian
 
J
.
Spin-orbit coupling in hydrogenated graphene
.
Phys Rev Lett
 
2013
;
110
:
246602
.

30.

Jiang
 
ZY
,
Lou
 
WK
,
Liu
 
Y
 et al.  
Spin-triplet excitonic insulator: the case of semihydrogenated graphene
.
Phys Rev Lett
 
2020
;
124
:
166401
.

31.

Zhou
 
J
,
Wang
 
Q
,
Sun
 
Q
 et al.  
Ferromagnetism in semihydrogenated graphene sheet
.
Nano Lett
 
2009
;
9
:
3867
70
.

32.

González-Herrero
 
H
,
Gómez-Rodríguez
 
JM
,
Mallet
 
P
 et al.  
Atomic-scale control of graphene magnetism by using hydrogen atoms
.
Science
 
2016
;
352
:
437
41
.

33.

Hu
 
S
,
Lozada-Hidalgo
 
M
,
Wang
 
FC
 et al.  
Proton transport through one-atom-thick crystals
.
Nature
 
2014
;
516
:
227
30
.

34.

Sun
 
PZ
,
Yang
 
Q
,
Kuang
 
WJ
 et al.  
Limits on gas impermeability of graphene
.
Nature
 
2020
;
579
:
229
32
.

35.

Yuan
 
GW
,
Lin
 
DJ
,
Wang
 
Y
 et al.  
Proton-assisted growth of ultra-flat graphene films
.
Nature
 
2020
;
577
:
204
8
.

36.

Xu
 
J
,
Liu
 
WL
,
Tang
 
WN
 et al.  
Trapping hydrogen molecules between perfect graphene
.
Nano Lett
 
2023
;
23
:
8203
10
.

37.

Deng
 
DH
,
Novoselov
 
KS
,
Fu
 
Q
 et al.  
Catalysis with two-dimensional materials and their heterostructures
.
Nat Nanotechnol
 
2016
;
11
:
218
30
.

38.

Cancado
 
LG
,
Jorio
 
A
,
Ferreira
 
EHM
 et al.  
Quantifying defects in graphene via Raman spectroscopy at different excitation energies
.
Nano Lett
 
2011
;
11
:
3190
6
.

39.

Ferrari
 
AC
,
Basko
 
DM
.
Raman spectroscopy as a versatile tool for studying the properties of graphene
.
Nat Nanotechnol
 
2013
;
8
:
235
46
.

40.

Son
 
J
,
Lee
 
JY
,
Han
 
N
 et al.  
Tunable wettability of graphene through nondestructive hydrogenation and wettability-based patterning for bioapplications
.
Nano Lett
 
2020
;
20
:
5625
31
.

41.

Tedeschi
 
D
,
Blundo
 
E
,
Felici
 
M
 et al.  
Controlled micro/nanodome formation in proton-irradiated bulk transition-metal dichalcogenides
.
Adv Mater
 
2019
;
31
:
1903795
.

42.

He
 
L
,
Wang
 
HS
,
Chen
 
LX
 et al.  
Isolating hydrogen in hexagonal boron nitride bubbles by a plasma treatment
.
Nat Commun
 
2019
;
10
:
2815
.

43.

Li
 
ZJ
,
Zhang
 
Z
,
Lu
 
J
.
van der Waals gap engineering of emergent two-dimensional materials
.
Accounts Mater Res
 
2024
;
6
:
52
63
.

44.

Lyu
 
P
,
Wang
 
ZY
,
Guo
 
N
 et al.  
Air-stable wafer-scale ferromagnetic metallo-carbon nitride monolayer
.
J Am Chem Soc
 
2024
;
146
:
20604
14
.

45.

Wang
 
L
,
Meric
 
I
,
Huang
 
PY
 et al.  
One-dimensional electrical contact to a two-dimensional material
.
Science
 
2013
;
342
:
614
7
.

46.

Xie
 
WJ
,
Weng
 
LT
,
Ng
 
KM
 et al.  
Clean graphene surface through high temperature annealing
.
Carbon
 
2015
;
94
:
740
8
.

47.

Huang
 
XL
,
Wan
 
ZH
,
Yuan
 
GW
 et al.  
Controllable defects in monolayer graphene induced by hydrogen and argon plasma
.
J Phys Condens Mater
 
2024
;
36
:
335304
.

48.

Zhang
 
YB
,
Brar
 
VW
,
Girit
 
C
 et al.  
Origin of spatial charge inhomogeneity in graphene
.
Nat Phys
 
2009
;
5
:
722
6
.

49.

Guillemette
 
J
,
Sabri
 
SS
,
Wu
 
BX
 et al.  
Quantum Hall effect in hydrogenated graphene
.
Phys Rev Lett
 
2013
;
110
:
176801
.

50.

Jia
 
X
,
Goswami
 
P
,
Chakravarty
 
S
.
Dissipation and criticality in the lowest Landau level of graphene
.
Phys Rev Lett
 
2008
;
101
:
036805
.

51.

McCann
 
E
,
Kechedzhi
 
K
,
Fal'ko
 
VI
 et al.  
Weak-localization magnetoresistance and valley symmetry in graphene
.
Phys Rev Lett
 
2006
;
97
:
146805
.

52.

Jayasingha
 
R
,
Sherehiy
 
A
,
Wu
 
SY
 et al.  
In situ study of hydrogenation of graphene and new phases of localization between metal-insulator transitions
.
Nano Lett
 
2013
;
13
:
5098
105
.

Author notes

Equally contributed to this work.

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse, distribution, and reproduction in any medium, provided the original work is properly cited.

Supplementary data