Muhammad Abiyyu Kenichi Purbayanto*a,
Subrata Ghosh
ab,
Dorota Moszczyńskac,
Carlo S. Casari
b and
Agnieszka Maria Jastrzębska*a
aFaculty of Mechatronics, Warsaw University of Technology, św. Andrzeja Boboli 8, Warsaw, 02-525, Poland. E-mail: muhammad_abiyyu.kenichi.dokt@pw.edu.pl; agnieszka.jastrzebska@pw.edu.pl
bDepartment of Energy, Micro and Nanostructured Materials Laboratory—NanoLab, Politecnico di Milano, via Lambruschini, Milano, 20156, Italy
cFaculty of Materials Science and Engineering, Warsaw University of Technology, Wołoska 141, Warsaw, 02-507, Poland
First published on 10th July 2025
The rapid advancement of nanomaterial-based thin-film processing has significantly contributed to the development of multifunctional optoelectronic devices. Among novel nanomaterials, MXenes, 2D transition metal carbides, nitrides, and carbonitrides have garnered substantial attention due to their high optical transparency, tunable optical properties, and excellent electrochemical performance. In particular, niobium carbide (Nb2CTx) MXene holds great promise for photoelectrochemical photodetectors (PEC PDs) due to its narrow-band photodetection capability, solution-processing, and stability under light irradiation. However, current Nb2CTx-based and 2D-based PEC PDs, in general, suffer from low photocurrent density, limited optical transparency, and poor environmental stability, hindering their practical applications. In this study, we developed a polymeric binder-free transparent reduced Nb2CTx/graphene oxide (r-Nb2CTx/GO) heterostructured thin film using a facile layer-by-layer technique. Incorporating reduced GO not only assists in improving the electrical conductivity of the heterostructure but also serves as a binder for MXene flakes. We systematically investigate the physicochemical properties of the film, its photodetection, and electrochemical performance. The optimized film exhibits outstanding transparency (70% at 550 nm), narrow-band photodetection response in the ultraviolet region, an excellent photoresponsivity of 50.21 μA W−1, and high environmental stability. Altogether, this study paves the way for developing Nb2CTx-based heterostructures for highly sensitive and environmentally stable transparent PEC PDs.
New conceptsFabricating all-solution-processed thin films based on van der Waals materials holds great potential for constructing large-scale, low-cost, and high-performance optoelectronic devices. However, achieving solution-processed thin films that simultaneously offer high electrical conductivity and excellent transparency remains a significant challenge. By leveraging the outstanding hydrophilicity of both graphene oxide (GO) and Nb2CTx MXenes, we develop an innovative approach to fabricate heterostructures via a layer-by-layer (LbL) assembly technique. The resulting Nb2CTx/GO heterostructures are subsequently reduced using L-ascorbic acid at a low temperature (90 °C). This strategy yields reduced r-Nb2CTx/GO (r-Nb2CTx/GO) heterostructures exhibiting remarkable oxidation and cycling stability, as well as tunable optical transparency by varying the number of LbL cycles. Simultaneously, the electrical conductivity of the films can be freely adjusted, balancing the trade-off with optical transparency. Using this method, we further developed a transparent photoelectrochemical photodetector operating in the UV region, achieving a transparency of 70% at 550 nm and a high photoresponsivity of 50.21 μA W−1. The superior performance of our novel heterostructure is attributed to several key factors: (i) a moderate and optimized electrical conductivity of the r-Nb2CTx/GO films, facilitating efficient electron transport to the current collector, (ii) effective charge separation and transfer across the r-Nb2CTx/GO interface, and (iii) enhanced light absorption of r-Nb2CTx/GO compared to standalone Nb2CTx. |
Recently, 2D-based photoelectrochemical photodetectors (PEC PDs) have become an emerging class of photodetectors due to their simple preparation, cost-effectiveness, and high sensitivity.16 The photodetection mechanism of PEC PDs is different from that of conventional solid-state PDs. In the case of PEC PDs, both physical (light absorption, photogenerated charge separation, and transport) and chemical processes (redox reactions) play important roles in regulating the photoresponse performance.16 Additionally, the interaction of electrolyte with the material interface in PEC PDs tunes the unique photoconductivity behaviour that depends on the wavelength of light.17
In this sense, Nb2CTx has received significant attention as an active material for PEC PDs due to its high specific surface area and narrow-band photoelectric response.18,19 For example, Gao et al. explored the potential of Nb2CTx as the PEC PD sensitive to 350–400 nm, with a narrow-band photoresponse. Its photodetection performance could be easily adjusted by controlling the external parameters, such as electrolyte concentrations, bias potential, and light power.19 However, the performance of these PEC PDs is still far from the merit achieved by their solid-state counterparts, with the responsivity (under 400 nm light) being only 3.74 μA W−1 (0.6 V bias voltage and 1 M KOH).
To further improve the photodetection performance of Nb2CTx, vdW heterostructures were fabricated to widen the response spectra and prolong the charge carrier lifetime.13,20 For example, Ren et al. developed Nb2CTx/SnS2 PEC PDs with an enhanced photocurrent density of up to 250-fold (3.75 μA cm−2) compared to that of pure Nb2CTx (0.015 μA cm−2).13
Despite the advancement in Nb2CTx-based PEC PDs, several challenges must be tackled to unlock their full potential. For instance, the currently developed Nb2CTx and, more broadly, MXene-based PEC PDs lack optical transparency and robust adhesion to the substrate, as the films are simply fabricated using drop-casting methods.13,19,21,22 Moreover, MXenes suffer from oxidative degradation, which can degrade the device's performance over time, hindering their practical applications.23 Additionally, polymers were often used as a protective layer as well as binders, which may decrease the electrical conductivity of MXene layers.24 Recently, layer-by-layer (LbL) assembly of MXene has been developed to fabricate thin films with high adhesion and easily adjusted optoelectronic properties.25 However, to the best of our knowledge, no study to date has reported the LbL assembly of a transparent thin film for the application as PEC PDs.
In this study, we successfully fabricated polymeric binder-free, transparent, and stable reduced Nb2CTx/GO (r-Nb2CTx/GO) PEC PDs using LbL assembly via a spin-coating technique. We utilized GO nanoflakes as they can facilitate a strong adhesion to the substrate and interfacial and large surface area contacts with MXenes, facilitating a strong adhesion to the substrate.26 Additionally, using reduced GO in heterostructure formulation gives several benefits, including good barrier properties for mitigating oxidation issues,27 facilitating efficient photogenerated interfacial charge transfer,28 and improving electrical conductivity.26
Therefore, we performed a detailed investigation of the morphological, structural, electrochemical, and optoelectronic properties of r-Nb2CTx/GO to design PEC PDs in a reasonable manner. While compared to pure Nb2CTx, we achieved 6.5-fold enhanced photoresponsivity of r-Nb2CTx/GO PEC PDs under ultraviolet (UV) irradiation, coupled with narrow-band photodetection features while maintaining optical transparency (70% at 550 nm). These findings substantially contribute to the development of Nb2CTx-based heterostructures for optoelectronic applications.
X-ray diffraction patterns (XRD) of Nb2CTx show the characteristic (002) peak located at 6.8° (Fig. S3a, ESI†), which is in good agreement with previous Nb2CTx results.30,31 In the case of GO, we observe two major peaks located at 10.8° (001) and 27.2° (002) (Fig. S3b, ESI†).32 These results confirm the high quality of Nb2CTx and GO used in this study. Furthermore, the XRD pattern of r-Nb2CTx/GO thin films shows the characteristic diffraction peaks of both Nb2CTx and GO without the presence of other heterogeneous phases, showing high phase purity and successful formation of the heterostructured thin film (Fig. S4, ESI†). Fig. 2d and e show the surface morphology of LbL r-Nb2CTx/GO thin films, which confirms a uniform decoration of Nb2CTx flakes on transparent GO layers. The Nb2CTx/GO thin film without thermal reduction is shown in Fig. S5 (ESI†). Upon reduction, the Nb2CTx is still tightly connected with GO layers, implying a strong interfacial interaction between these materials (Fig. 2d). The cross-sectional images indicated that the r-Nb2CTx/GO nanoflakes are well aligned parallelly onto the substrate (Fig. 2f). The EDS elemental mapping (Fig. 2g) of the LbL film shows an even and uniform distribution of Nb, C, O, and Cl elements, which are the constituent elements of the heterostructure. For more details, the EDS spectra of r-Nb2CTx/GO thin films and their corresponding quantification are given in Fig. S6 (ESI†). To further confirm the chemical composition and surface chemistry of r-Nb2CTx/GO thin films, we carried out X-ray photoelectron spectroscopy (XPS) measurements. The XPS survey spectra (Fig. S7, ESI†) show the strong contribution of C 1s and O 1s compared to Nb 3d, implying that GO is most noticeable at the surface of the film and covers Nb2CTx flakes. Notably, C 1s spectra are fitted by five fitting components, as shown in Fig. S8a (ESI†). The peak at 284.5 eV corresponds to sp2 C–C bonding, while the peaks located at 285.7, 286.6, 287.8, and 288.9 eV can be assigned to sp3 C–C bonding, C–O, CO, and COOH, respectively.33 The high fraction of sp2 C–C bonding (Table S1, ESI†) shows evidence of GO reduction. At the same time, the low contribution from oxygen-containing functional groups further confirms the partial removal of surface oxygen functionalities of GO.34 The carbide component, which typically comes from Nb2CTx, cannot be clearly identified in C 1s spectra, which might be related to the uniform surface coverage of GO. On the other hand, the Nb 3d signal from Nb2CTx can be fitted by a doublet component of Nb5+ at 207.5 and 210.3 eV (Fig. S8b, ESI†).35 The O 1s spectra can be fitted with four components located at 531.1, 532.6, 533.5, and 535.0 eV, corresponding to C–Nb–Ox, C
O, C–O, and adsorbed H2O at the surface (Fig. S8c, ESI†).36 The occurrence of C–Nb–Ox implies that Nb2CTx in r-Nb2CTx/GO is terminated by
O functional groups. The surface roughness of thin films is critical in optimizing their photodetection performance. AFM topography images (Fig. S9, ESI†) reveal that the r-Nb2CTx/GO film exhibits a low root-mean-square roughness of 3.40 nm, indicating a smooth and continuous surface of the fabricated film. A smoother surface can reduce light scattering and ensure more favorable carrier transport and good optical transparency of the film.4,37
Raman spectra were measured to further elucidate the crystal structure of MXene. Fig. 3a shows Raman spectra of Nb2CT and r-Nb2CTx/GO thin films. Notably, the MXene region (50–900 cm−1) was fitted to unravel the contribution of each phonon mode (Fig. S9, ESI†). The peaks are assigned based on the experimental computational results of Giordano et al. (Table S2, ESI†).38 The ω1 (159 cm−1) and ω3 (258 cm−1) modes can be assigned to in-plane and out-of-plane Nb2C(OH)2. The ω2 (209 cm−1) mode corresponds to the phonon mode of out-of-plane Nb2CO2. Moreover, the ω4 (395 cm−1) and ω5 (507 cm−1) modes correspond to the phonon mode of the –OH functional group, whereas the ω6 (685 cm−1) mode corresponds to the oxygen functional group. The last peak of ω7 (745 cm−1) is originated from the phonon mode of C atoms. Altogether, Raman spectra confirm Nb2CTx terminated by –OH and O groups without the presence of Raman peaks, originating from oxide species. We do not observe the contribution from the phonon modes of the –F terminal group, and the intensity of –OH-related modes is relatively weak. This observation is in line with the previous study, which showed that the
O group is thermodynamically stable compared to –F and –OH terminal groups on the Nb2CTx surface.39 In the amorphous carbon region (1200–1800 cm−1), we observe broad peaks of D and G bands originating from the disorder-induced peak and in-plane vibration of sp2 carbon, respectively.40 The presence of these bands can be related to the synthesis process as reported by other works.41,42
In r-Nb2CTx/GO thin films, both vibrational modes from Nb2CTx and rGO co-exist, confirming the successful fabrication of the heterostructure. We observe the characteristic D-band (1347 cm−1) and G-band (1594 cm−1) of graphitic carbon with strong intensity. Here, the calculated ID/IG ratio of r-Nb2CTx/GO (1.24) is higher than those of GO (0.78) and Nb2CTx (0.72) (Fig. S10, ESI†), indicating the formation of defects during the reduction process and the decrease of oxidation levels in GO.43 Lower full width at half maximum of the D-band and G-band of r-Nb2CTx/GO than that of Nb2CTx is also observed (Table S3, ESI†), indicating that the dominant features of sp2-carbon come from rGO and also the reduction process induced graphitization of amorphous carbon. Additionally, the characteristic phonon modes of Nb2CTx are preserved without noticeable niobium oxide-related vibrations. Fig. S11 (ESI†) shows the curve fitting of Raman modes for r-Nb2CTx/GO in the MXene region. Most of the peaks in the low-frequency region are blue-shifted compared to Nb2CTx (Table S4, ESI†), which might be caused by interfacial charge redistribution and structural modification of Nb2CTx upon hybridization with rGO.38,44,45 We observe an interesting observation in the surface termination region, where ω6 intensity is markedly increased in comparison to ω7, indicating that the O functional group becomes more dominant in the heterostructure compared to the Nb2CTx counterpart.
To employ the materials for optoelectronic applications, a material should have excellent optical and electrical properties. We monitored the optical transmittance changes on Nb2CTx/GO, acting as a starting film before reduction (Fig. 3b). The optical transmittance of Nb2CTx/GO decreases across the entire measured wavelength range as the number of coating layers increases, indicating that the film absorbs both UV and visible light more. Fig. 3c shows the effect of reduction time on the optical transmittance of Nb2CTx/GO-3, where transparency is decreased by increasing the reduction time. The absorption spectra of Nb2CTx/GO-3 showed two distinct peaks located at 234 and 298 nm (Fig. 3d), which are the fingerprints of GO. The former is originated from π–π* transitions of CC in amorphous carbon, while the latter is connected to n–π* transitions in C
O bonds.46 After reduction, only a single peak located at 271 nm is observed.47 This fact ensures a successful transformation from Nb2CTx/GO to r-Nb2CTx/GO.
To assess the evolution of electrical conductivity upon reduction, we performed I–V measurements of Nb2C2Tx by varying reduction time (Fig. 3e). In particular, the Nb2CTx/GO-3 showed no observable I–V curve owing to low electrical conductivity exceeding the measurement limit of the source meter, hence not shown in the figure. We begin to observe the Ohmic I–V response by performing 30 minutes of reduction, and the electrical conductivity keeps increasing by prolonging the reduction time. The calculated resistance of r-Nb2CTx/GO-3 obtained by 30 minutes reduction is 1.72 GΩ, and it markedly decreased to 85.62 kΩ for 60 minutes of reduction and 32.52 kΩ for 120 minutes of reduction (Fig. S12, ESI†). At the same time, we observe a trade-off between optical transparency and electrical conductivity. It should be noted that at least ∼70% transparency should be maintained to use a material for transparent optical applications.48 Since the transmittance of r-Nb2CTx/GO-3 (70% at 550 nm) falls between that of Nb2CTx (75%) and rGO (61%) (Fig. S13, ESI†), and it shows good electrical conductivity, we choose r-Nb2CTx/GO-3 with 30 minutes of reduction for further analysis. The good optical transparency of the r-Nb2CTx/GO film is also evidenced by digital photography, where the film maintains its transparency after reduction (Fig. 3f). Interestingly, the reduction procedure does not cause peeling of the Nb2CTx/GO thin film, as confirmed from the tape adhesion test (Fig. S14, ESI†). The film shows minimal peeling after testing, even in the absence of a polymeric binder in the fabrication process. This fact cannot be achieved with regular drop-casting or spray-coating methods, where polymeric binders are used.49,50
We then tested the environmental stability by assessing the electrical conductivity of the thin film over time and monitoring the change of Raman spectra after storing the sample under ambient conditions.51,52 This method is more accurate and straightforward than the common method, which measures the change of the optical absorbance of the film.25 Fig. 4a shows that Nb2CTx loses electrical conductivity over time, as indicated by an increase in the resistivity (R/R0). Whereas, r-Nb2CTx/GO shows a negligible change in resistivity, indicating a high oxidation stability of this heterostructure. This high stability is attributed to the greater hydrophobicity of rGO compared to MXenes, along with its low gas and water permeability in ambient air.26,27 Additionally, the stabilization effect could be attributed to the removal of O2− ions in the carbon sites of MXenes using L-ascorbic acids, resulting in lower nucleation sites for metal oxides’ growth.53
From the Raman measurement after aging under ambient conditions, the most striking feature observed in aged-Nb2CTx is the significant amplification of D- and G-bands relative to the peaks in the low-frequency region, indicating the dominance of amorphous carbon (Fig. 4b). An increased graphitization is observed in aged-Nb2CTx, where we observe a blue shifting of D- and G-bands and an increment in ID/IG (Table S5, ESI†).54 Fitting analysis in the low-frequency region (Fig. S15, ESI†) reveals a blueshift in the ω1, ω2, ω3, ω5, and ω6 peaks, suggesting partial oxidation of Nb atoms. In contrast, ω7 exhibits a redshift. These peaks slightly shift by 1–6 cm−1, indicating surface modification upon aging. Furthermore, the ω1/ω3 ratio is an important parameter to quantify the changes in Nb2CTx surface chemistry of Nb2CTx due to degradation. In aged-Nb2CTx, ω1 becomes dominant, resulting in increased ω1/ω3 from 0.89 to 1.04, providing evidence for carbon network disruption (Table S6, ESI†).38
On the other hand, in the case of r-Nb2CTx/GO, we cannot assess the evolution of amorphous carbon originating from MXene upon aging, as it is overshadowed by a strong signal from rGO. However, the degradation study of heterostructures can still be monitored through fitting analysis in the low-frequency region (Fig. S16, ESI†). From the fitting results, we observe a large redshift of all peaks, ranging from 3 to 22 cm−1, which can be attributed to the removal of intercalated water or changes in surface functionalities (Table S4, ESI†).38,55 On the other hand, the relative intensity of ω1 remains largely unchanged, in contrast to aged-Nb2CTx, indicating minimal sample degradation (Table S6, ESI†). Thus, the Raman analysis supports the finding of electrical measurement results, in which the oxidation stability of Nb2CTx is improved by creating an interfacial contact with rGO.
After successfully fabricating an optically transparent and environmentally stable r-Nb2CTx/GO, we explore its multifunctionality as an active electrode in PEC PDs. The previous study shows that Nb2CTx exhibits a strong absorption from UV to near-infrared regions, making it promising as a building block for various optoelectronic devices.19 First, we performed cyclic voltammetry (CV) measurements to analyze the electrochemical properties of the thin film. The CV response of Nb2CTx shows no observable redox peak, with the current densities increasing with an increase in the scan rate (Fig. 4d). Bare Nb2CTx is resistive in nature as the CV response is not parallel with the voltage-axis and more inclined towards the current-axis, which is in agreement with the electrical properties. Moreover, r-Nb2CTx/GO shows a more rectangular CV response, implying better ionic and electronic transport within the film structure than Nb2CTx alone (Fig. 4e). Additionally, the CV curve retains its quasi-rectangular shape upon scan rate increment. We further calculate the areal capacitance (CA) from the CV curve by taking the linear fitting of current density under various scan rates, as shown in Fig. S17 (ESI†). Notably, r-Nb2CTx/GO showed significant improvement in the CA value (84.21 μF cm−2) compared to Nb2CTx (32.90 μF cm−2). The CA of r-Nb2CTx/GO is larger than other transparent films, such as Ti3C2Tx (75.2 μF cm−2 with 80% optical transparency),56 and graphene (12.4 μF cm−2 with 67% optical transparency).57 A relatively high CA in r-Nb2CTx/GO opens the possibility of being used as a transparent energy-storage device. Furthermore, we examined the stability of r-Nb2CTx/GO by performing continuous CV cycling (Fig. 4f). Here, the film shows good stability, with a capacity retention of 89% after 1000 cycles (Fig. S18, ESI†).
To study the electron transport of the heterostructure, the electrochemical impedance spectroscopy (EIS) spectra have been measured. Fig. 5a shows the Nyquist plot of Nb2CTx and r-Nb2CTx/GO. In the case of Nb2CTx, a large semicircle is observed, indicating charge transfer between the film surface and the electrolyte. The r-Nb2CTx/GO curve depicts the absence of a semicircle in the high-frequency region. However, we observe that the imaginary part of the impedance spectra changes rapidly moving away from the real axis, indicating an excellent electrical conductivity of the film.58 r-Nb2CTx/GO (13.18 Ω) exhibits a lower equivalent series resistance compared to Nb2CTx (17.85 Ω). This positive effect can significantly enhance the photoresponse and stability of r-Nb2CTx/GO as photoelectrochemical PEC PDs. The properties of the films were tested in a three-electrode measurement setup using 0.5 M Na2SO4 as an electrolyte, as illustrated in Fig. 5b. To investigate the performance of PEC PDs, we measured the photocurrent density of Nb2CTx and r-Nb2CTx/GO under UV light irradiation (Fig. 5c). In particular, the r-Nb2CTx/GO heterostructure exhibited a 2.8-fold higher photocurrent (0.65 μA cm−2) with a typical stable on–off cycling under UV irradiation compared to Nb2CTx (0.23 μA cm−2). The high photodetection performance of r-Nb2CTx/GO is attributed to the improved electrical conductivity and light absorption of the film. Thus, heterostructure formation provides efficient light harvesting, accelerated transport of photogenerated charge carriers, and enhanced charge separation efficiency.59,60 Next, to further quantify the photodetection performance of the films, we calculate their photoresponsivity under different light wavelengths (Fig. 5d).
The obtained photoresponsivity values for Nb2CT and r-Nb2CTx/GO are summarized in Table S7 (ESI†). Both films exhibited a dominant photoresponsivity under UV light (365 nm). A lower photoresponsivity is also observed under violet (395 nm) and blue light (457 nm), where the photoresponse is considerably low beyond this wavelength. Therefore, these results indicate that Nb2CTx and r-Nb2CTx/GO effectively harvest photons in the short wavelength (365 nm), whereas no noticeable photoresponse is observed for wavelengths longer than 457 nm. A similar narrow band (350–400 nm) photoresponse behavior of Nb2CTx was also reported previously.19
Additionally, Gao et al. calculated the work function of –OH-terminated Nb2CTx to be 3.33 eV, corresponding to the critical wavelength of dominant photoresponsivity, i.e., 365 nm.19 The calculated photoresponsivity of Nb2CTx is 7.75 μA W−1 at 365 nm, which is significantly enhanced in r-Nb2CTx/GO, reaching 50.21 μA W−1 (a 6.5-fold enhancement). Note that the photoresponsivity value of r-Nb2CTx/GO is much higher than previously reported Nb2CTx PEC PDs, which exhibited 12.6 μA W−1 at 1 M KOH.19
In order to further explore the performance of the r-Nb2CTx/GO PEC PDs, we measured the photoresponse under UV light irradiation with different light powers, as depicted in Fig. 5e. Here, the photocurrent density of r-Nb2CTx/GO can be increased almost linearly as we increase the illuminated light power, owing to more photogenerated electron and hole pairs.61 In particular, a 2.3-fold photocurrent enhancement can be achieved by increasing the light power from 4.2 to 27.2 mW cm−2. Interestingly, as the light power increases, we start to observe a sharp transient peak at a light power of 20.9 mW cm−2, which further signifies the highest laser power (27.1 mW cm−2). The peak further exponentially drops and stabilizes over time before it falls to a dark current when the light is chopped off. This sharp and instantaneous transient peak is attributed to the capacitive current, in which the PDs are rapidly charged when illuminated by relatively large light power. This behavior can be connected to charge accumulation phenomena at the heterostructure interface.62 Therefore, the photoresponse of r-Nb2CTx/GO can be tuned by varying light power.
To further evaluate the PEC PD performance, we measured linear sweep voltammograms (LSV) at a scan rate of 10 mV s−1. Fig. 5f shows the LSV curves of r-Nb2CTx/GO in the dark and under UV or blue light irradiation. The results indicate that no noticeable redox peaks are observed across the measured potential range. We also observe the presence of anodic photocurrent, with a more pronounced increase in current density under irradiation with UV compared to blue light. Table S8 (ESI†) compares the photoresponse performance of r-Nb2CTx/GO with other state-of-the-art nanomaterial-based PEC PDs. Notably, r-Nb2CTx/GO demonstrates high photoresponsivity while maintaining excellent optical transparency, which is not commonly achieved in conventional PEC PDs, as most developed PEC PDs are opaque in visible light. This combination makes r-Nb2CTx/GO suitable for see-through optoelectronic devices. Although some studies have attempted to develop transparent PEC PDs, they often suffer from limited visible light transmittance (<60%), complex fabrication procedures, and sluggish photoresponsivity.63–65 Herein, the high photodetection performance of r-Nb2CTx/GO can be attributed to several key factors, including (i) a moderate and optimized electrical conductivity of r-Nb2CTx/GO, facilitating faster electron transport to the current collector, (ii) efficient charge separation and transfer across the r-Nb2CTx/GO interface, and (iii) enhanced light absorption of r-Nb2CTx/GO compared to standalone Nb2CTx, as observed in the UV-Vis results.26,45 Altogether, this study demonstrates the promise of r-Nb2CTx/GO thin films for transparent and highly stable PEC PDs in which the photoresponse performance can be easily tuned by changing the external parameters.
In the case of the photoelectrochemical response, the developed r-Nb2CTx/GO PEC PDs show a highly stable cycling performance, as indicated by 89% capacity retention after 1000 cycles. Additionally, the r-Nb2CTx/GO PEC PDs showed narrow-band photosensitivity in the UV region, with responsivity enhanced 6.5-fold compared to that of solo Nb2CTx, which further underscores the importance of adding rGO in facilitating more efficient light harvesting and photogenerated charge separation and transport. Interestingly, the photoresponse behavior of the PDs can be easily tuned by changing the light power, where we start to observe capacitive current at high incident light power. Altogether, our results open a new insight into the exploration of transparent and binder-free MXene-based PEC PDs, paving the way toward their use in advanced optoelectronic devices. The transparency of PEC PDs can be further enhanced without compromising electrical conductivity by controlling MXene surface chemistry and fabricating thin films composed of large-sized (micron-scale) MXene flakes, which help in reducing interflake resistance.
X-ray photoelectron spectroscopy (XPS) measurements were conducted using a PHI 5000 VersaProbe II Scanning ESCA Microprobe (ULVAC-PHI; Japan) spectrometer. The XPS spectra were obtained under the following conditions: monochromatic Al-Kα radiation (hν = 1486.6 eV), an X-ray source operating at 25 W, 15 kV, and a 100 μm spot size, and an area of 500 μm2. High-resolution XPS spectra were collected with an energy step size of 0.1 eV and an analyzer pass energy of 23.5 eV. The obtained XPS spectra were analyzed using CasaXPS software (v.2.3.22, Casa Software Ltd, United Kingdom) using the sensitivity factors native for the hardware. Deconvolution of all high-resolution XPS spectra was performed using a Shirley background and a Gaussian peak shape with 30% Lorentzian character. The Raman spectra of the films were recorded using a Renishaw InVia Raman spectrometer with a 514.5 nm laser and 1800 lines per mm grating spectrometer. We used a 50× microscope objective, 10 seconds of acquisition time, 10 accumulations, and a low laser power of 0.4 mW to avoid laser-induced damage to the samples. The spike caused by cosmic rays was removed using the single-spectrum spike removal algorithm developed by Coca-Lopez et al.67 For Raman peak analysis, the recorded Raman spectra were baseline-subtracted and fitted using a Lorentzian function.68 Multiple spectra were recorded at different spots on the same sample, and the error was calculated as the standard deviation.
Cyclic-voltammetry (CV) was measured within a voltage window of −0.2 to 0.6 V vs. Ag/AgCl with different scan rates ranging from 10 to 100 mV s−1. Before measurements, a conditioning step of 10 cyclic CV was performed with a scan rate of 50 mV s−1. This conditioning was intended to promote the penetration of electrolytes into electrodes.69 The areal capacitance (CA) is estimated by performing the linear fitting of current density vs. scan rate, according to eqn (1):
![]() | (1) |
The photoelectrochemical (PEC) study was conducted by irradiating the thin films with an LED source (Photocube, Photochemical reactor, ThalesNano, Budapest), at various wavelengths, including 365, 395, 457, 500, 523, and 625 nm. The light intensity was calibrated with an optical power meter equipped with a Si sensor (Thorlabs, PM120VA). Chronoamperometry measurements were recorded at 0.6 V vs. Ag/AgCl with an interval of 0.5 seconds. Furthermore, linear sweep voltammetry (LSV) was scanned in an anodic direction with a scan rate of 10 mV s−1 under dark conditions and irradiation.
Footnote |
† Electronic supplementary information (ESI) available: Figures: EDS spectra of Nb2CTx, GO, and r-Nb2CTx/GO. SEM images of Nb2CTx/GO. UV-Vis absorption spectra of Nb2CTx aqueous dispersion. XRD spectra of Nb2CTx, GO, and r-Nb2CTx/GO thin films. AFM images of r-Nb2CTx/GO thin films. XPS survey and high-resolution spectra of r-Nb2CTx/GO. The evolution of electrical resistance of r-Nb2CTx/GO thin films by increasing the reduction time. Optical transmittance comparison of Nb2CTx, rGO, and r-Nb2CTx/GO thin films. The tape adhesion test on r-Nb2CTx/GO thin films. The Lorentzian fit of the Raman spectra of fresh and aged Nb2CTx and r-Nb2CTx/GO thin films. Raman spectra of GO. Current vs scan rate curve. The capacity retention of r-Nb2CTx/GO thin films. Tables: summary of Nb 3d, C 1s, and O 1s XPS signals for r-Nb2CTx/GO thin films. Results from peak fitting for Nb2CTx and r-Nb2CTx/GO. Photoresponsivity of Nb2CTx and r-Nb2CTx/GO thin films under different wavelengths. Comparison of the photoresponse performance for PEC photodetectors. See DOI: https://doi.org/10.1039/d5nh00280j |
This journal is © The Royal Society of Chemistry 2025 |