Two-dimensional (2D) noble-metal dichalcogenides exhibit exceptionally strong thickness-dependent bandgaps, which can be leveraged in a wide variety of device applications. A detailed study of their optical (e.g., optical bandgaps) and electrical properties (e.g., mobilities) is important in determining potential future applications of these materials. In this work, we perform detailed optical and electrical characterization of 2D PdSe2 nanoflakes mechanically exfoliated from a single-crystalline source. Layer-dependent bandgap analysis from optical absorption results indicates that this material is an indirect semiconductor with bandgaps of approximately 1.37 and 0.50 eV for the monolayer and bulk, respectively. Spectral photoresponse measurements further confirm these bandgap values. Moreover, temperature-dependent electrical measurements of a 6.8-nm-thick PdSe2 flake-based transistor show effective electron mobilities of 130 and 520 cm2 V−1 s−1 at 300 K and 77 K, respectively. Finally, we demonstrate that PdSe2 can be utilized for short-wave infrared photodetectors. A room-temperature specific detectivity (D*) of 1.8 × 1010 cm Hz1/2 W−1 at 1 μm with a band edge at 1.94 μm is achieved on a 6.8-nm-thick PdSe2 flake-based photodetector.
Two-dimensional (2D) transition metal dichalcogenides (TMDCs) have attracted considerable attention and have been explored for numerous applications including electronics, optoelectronics, catalysis, and sensors in the last decade.1 Motivated by the attractive properties and wide applications of TMDCs, researchers have explored other 2D two-element systems including the 2D noble-metal dichalcogenides (NMDCs) with the general formula MX2 (where M = Pt, Pd and X = S, Se, Te).2 NMDCs have recently been synthesized and demonstrated to exhibit air stability and strongly thickness-dependent bandgaps. For example, PtSe2 and PdS2 exhibit a sharp thickness-modulated semiconductor-to-metal transition,3–5 whereas PtS2 also shows a strong thickness dependence but is a narrow-bandgap indirect semiconductor in bulk.6 Both PtTe2 and PdTe2 are semimetals,7,8 with the latter exhibiting superconductivity.9,10
Palladium diselenide (PdSe2), a NMDC, has also been synthesized and explored in the last few years. Field-effect transistors (FETs) based on PdSe2 flakes show electron mobilities up to ∼200 cm2 V−1 s−1.11 PdSe2 was experimentally reported to have a thickness-dependent bandgap, in which monolayer PdSe2 has a bandgap of ∼1.4 eV and bulk PdSe2 is metallic with a bandgap approaching 0 eV.12 More recently, PdSe2 has been investigated for infrared detectors.13 In this work, we systematically study the optical and electrical properties of PdSe2 with an emphasis on its thickness-dependent bandgap to determine future applications of this material. Our results show that PdSe2 is an indirect semiconductor with a monolayer bandgap of 1.37 eV and a bulk bandgap of 0.5 eV (in contrast to previous works which have predicted metallic bulk), as shown by optical absorption and photoresponse measurements. Furthermore, the temperature-dependent electrical measurements of a 6.8-nm-thick PdSe2 flake-based transistor show effective electron mobilities of 130 and 520 cm2 V−1 s−1 at 300 K and 77 K, respectively. Owing to its relatively small bandgap, thick PdSe2 can be utilized for short-wave infrared (SWIR) photodetectors. A room-temperature peak specific detectivity (D*) of 6.4 × 1010 cm Hz1/2 W−1 at 0.7 μm with a band edge at 1.94 μm is achieved on a 6.8-nm-thick PdSe2 flake-based photodetector.
The crystal structure of PdSe2 is shown in Figs. 1(a) and 1(b). PdSe2 is a layered 2D material with a theoretical monolayer thickness of ∼4.8 Å. PdSe2 also has pentagonal-structured layers,11,12 as shown in Fig. 1(b). Pentagonal 2D materials have been theoretically proposed14,15 although PdSe2 has been synthesized and experimentally characterized as a 2D material. Figure 1(c) shows the measured X-ray diffraction (XRD) pattern of a bulk PdSe2 crystal grown by chemical vapor transport (CVT), which is consistent with the simulated XRD pattern, confirming the synthesis of bulk single crystal PdSe2. 2D PdSe2 nanoflakes of varying thicknesses were mechanically exfoliated from a large single-crystalline source. Figure 1(d) shows the Raman spectra of PdSe2 flakes of thicknesses ranging from monolayer to bulk. The locations of the Raman modes Ag1-Bg11, Ag2, Bg12, Ag3, and Bg13 and their shifts with the changing thickness are consistent with previous reports on PdSe2.12 Figure 1(e) shows the high-resolution transmission electron microscopy (HR-TEM) image of a 6.8-nm-thick PdSe2 flake. The lattice constant is 0.29 nm, which can be assigned to the (200) planes of the PdSe2 single crystal. The corresponding selected area electron diffraction (SAED) pattern shows a single set of bright diffraction dots [Fig. 1(f)], further confirming the single crystalline nature.
Crystal structure of PdSe2: (a) lateral and (b) top-down view. (c) Measured and simulated XRD spectra of a bulk PdSe2 crystal. (d) Raman spectra of PdSe2 for various thicknesses. (e) HR-TEM image and (f) SAED pattern of a 6.8-nm-thick PdSe2 flake.
Crystal structure of PdSe2: (a) lateral and (b) top-down view. (c) Measured and simulated XRD spectra of a bulk PdSe2 crystal. (d) Raman spectra of PdSe2 for various thicknesses. (e) HR-TEM image and (f) SAED pattern of a 6.8-nm-thick PdSe2 flake.
To perform optical measurements, PdSe2 flakes were exfoliated onto quartz by mechanical exfoliation and gold-mediated exfoliation,16 which can yield large-area mono- and few-layer PdSe2 flakes. Optical transmittance and reflectance measurements were conducted using both a UV-Vis microabsorption setup and a Fourier-transform infrared microscope. The combination of these techniques permits measurements over a wavelength range of 450 nm–15 μm.17 UV-Vis measurements were performed with reference to a blank quartz substrate and a silver mirror [Figs. 2(a) and 2(b)], while FTIR measurements were performed with reference to a blank quartz or KBr substrate and a gold mirror [Figs. 2(c) and 2(d)]. Quartz is optically transparent for visible to SWIR wavelengths, whereas the transparency of KBr extends further a few microns beyond long-wave infrared (LWIR) wavelengths. Note that the reflectance and transmittance of thick PdSe2 flakes show out-of-phase oscillations due to multiple reflections occurring in the 2D film; this destructive and constructive interference behavior is accounted for when calculating the absorption coefficient, as shown more clearly in Fig. S1 in the supplementary material. We did not observe differences in transmittance and reflectance of PdSe2 flakes prepared by mechanical exfoliation and gold-mediated exfoliation. To account for thin-film interference in the PdSe2 flake, we used the transfer matrix method. For transmission and reflection of light by a thin film on a substrate, we define the Fresnel coefficients18
where n0 = 1.00, n1 = n − ik, and n2 = 1.46 are the refractive indices of air, the PdSe2 flake, and quartz, respectively. The reflected and transmitted amplitudes are
where δ1 = (2π/λ)n1d1 and d1 is the thickness of the PdSe2 flake.18 The reflectance and transmittance as ratios of the energy reflected and transmitted to the energy incident are18
(a) Transmission and (b) reflection measurements of a 3.8-nm-thick PdSe2 flake on a quartz substrate in the visible-wavelength range. (c) Transmission and (d) reflection measurements of a 134-nm-thick PdSe2 flake on a KBr substrate in the infrared-wavelength range. (e) Optical indirect bandgap extraction of PdSe2 for various thicknesses. (f) Optical indirect bandgap of PdSe2 as a function of its thickness.
(a) Transmission and (b) reflection measurements of a 3.8-nm-thick PdSe2 flake on a quartz substrate in the visible-wavelength range. (c) Transmission and (d) reflection measurements of a 134-nm-thick PdSe2 flake on a KBr substrate in the infrared-wavelength range. (e) Optical indirect bandgap extraction of PdSe2 for various thicknesses. (f) Optical indirect bandgap of PdSe2 as a function of its thickness.
We then used the Nelder-Mead algorithm to minimize (Rmeas − Rsim)2 + (Tmeas − Tsim)2 to estimate the complex refractive index, n1 = n − ik, of the PdSe2 flake.19,20 Finally, we extracted the optical bandgap of the PdSe2 flake from a Cody plot of (α/hν)1/2 vs hν, where α = 4πk/λ is the absorption coefficient of the PdSe2 flake,19 as shown in Fig. 2(e) for PdSe2 flakes of varying thicknesses. The optical microscopy images of these flakes along with their thicknesses measured via atomic force microscopy (AFM) are shown in Fig. S2 in the supplementary material. Figure 2(f) shows the measured optical indirect bandgap of PdSe2 as a function of its thickness, which agrees with theoretical calculations in terms of the monolayer bandgap, but deviates from previously reported experimental data in terms of the bulk bandgap.12 This experimental deviation is likely due to both accounting for thickness-dependent reflectance (in addition to thickness-dependent transmittance) and the broader spectral range (∼0.35–3.1 eV) measured in this work, which fully encompasses the bandgap range of PdSe2. A previous work on PdSe2 extrapolated optical bandgap values from a measured spectral range of ∼1.3–3.25 eV, which is nearly a whole electron volt beyond the bulk bandgap of PdSe2.12 Accounting for thin-film interference only changes the bandgap by ∼0.1 eV (supplementary material Fig. S1),12 whereas neglecting reflectance and extrapolating far beyond the bandgap can drastically alter the bandgap value (supplementary material Fig. S3). This can be clearly seen for reflection and transmission measurements performed on a 134-nm-thick PdSe2 flake which shows an absorption edge around 2 μm (supplementary material Fig. S4).
Using electron-beam lithography, back-gated PdSe2 field-effect transistors (FETs) were fabricated on 90 nm SiO2/Si with 100-nm-thick nickel (Ni) contacts. The device structure and optical microscopy images of fabricated devices are shown in Fig. 3(a). Figure 3(b) shows the temperature-dependent Id-Vg characteristics of the 6.8-nm-thick PdSe2 device measured at a low drain voltage Vd of 10 mV, indicating that the on/off ratio increases as temperature decreases. The inset shows a plot of the same data on the linear-Id scale, clearly showing that the drain-source current Id,max increases with decreasing temperature. We can therefore expect the mobility of this 6.8-nm-thick PdSe2 flake to reach its maximum at low temperatures. Id-Vg shows ambipolar conduction since the onset of n-type and p-type conduction is roughly centered around Vg = 0 V, suggesting that the material is roughly intrinsic (i.e., low background doping). The ambipolar characteristics of PdSe2 can be drastically tuned by introducing molecular dopants to PdSe2 while annealing under vacuum.11 The same temperature-dependent Id-Vg plot is shown in Fig. S5 in the supplementary material but with a second measurement of the device at T = 300 K taken after 9 months of storage in nitrogen and ambient air, demonstrating the air stability of PdSe2. From linear-Id-Vg plots, we extrapolate room-temperature threshold voltages Vt of 2.25 V and 4.45 V for the 6.8-nm- and 116-nm-thick PdSe2 devices, respectively.
(a) Back-gated PdSe2 field-effect transistor (FET) structure and optical microscopy images of devices with 6.8-nm- and 116-nm-thick PdSe2 flakes. Scale bars are 30 μm. (b) Temperature-dependent Id-Vg characteristics of the 6.8-nm-thick PdSe2 device measured at a low Vd value of 10 mV. The inset shows a plot of the same data on the linear-Id scale. (c) Id-Vd characteristics of the 6.8-nm-thick PdSe2 device inset with those of the 116-nm-thick PdSe2 device. (d) Effective and field-effect mobility of the 6.8-nm-thick PdSe2 device. The dashed line shows a power law fit μeffective ∝ T−γ, where γ = 1.08. (e) Arrhenius plot showing the minimum drain-source current (Id,min) of the 6.8-nm-thick PdSe2 device as a function of temperature to extract the transport bandgap Eg = 0.57 eV. (f) Temperature-dependent Id-Vg of the 116-nm-thick PdSe2 device measured at Vd = 10 mV, showing bulk semiconducting behavior.
(a) Back-gated PdSe2 field-effect transistor (FET) structure and optical microscopy images of devices with 6.8-nm- and 116-nm-thick PdSe2 flakes. Scale bars are 30 μm. (b) Temperature-dependent Id-Vg characteristics of the 6.8-nm-thick PdSe2 device measured at a low Vd value of 10 mV. The inset shows a plot of the same data on the linear-Id scale. (c) Id-Vd characteristics of the 6.8-nm-thick PdSe2 device inset with those of the 116-nm-thick PdSe2 device. (d) Effective and field-effect mobility of the 6.8-nm-thick PdSe2 device. The dashed line shows a power law fit μeffective ∝ T−γ, where γ = 1.08. (e) Arrhenius plot showing the minimum drain-source current (Id,min) of the 6.8-nm-thick PdSe2 device as a function of temperature to extract the transport bandgap Eg = 0.57 eV. (f) Temperature-dependent Id-Vg of the 116-nm-thick PdSe2 device measured at Vd = 10 mV, showing bulk semiconducting behavior.
Figure 3(c) shows the Id-Vd characteristics of the 6.8-nm-thick PdSe2 device inset with those of the 116-nm-thick PdSe2 device. Since the drain saturation voltage Vd,sat = Vg − Vt exceeds the range of measurements, we do not observe saturation. Additionally, the linear behavior of the Id-Vd characteristics at low values of Vd suggests minimal contact resistance with small Schottky barrier heights. From these transfer characteristics, we calculated the effective electron mobility, μeff = (dId/dVd)(Lg/wc)[Cox(Vg − Vt − 0.5 Vd)]−1, and field-effect electron mobility, μfe = (dId/dVg)(Lg/wc)(CoxVd)−1, of the 6.8-nm-thick PdSe2 device vs temperature, as shown in Fig. 3(d). The electron mobility of the 6.8-nm-thick PdSe2 device increases with decreasing temperature. The 6.8-nm-thick PdSe2 device exhibits effective electron mobilities of 130 and 520 cm2 V−1 s−1 at 300 K and 77 K, respectively. We can fit the temperature-dependent mobility of this device with a power law, μeff ∝ T−γ, where γ = 1.08, suggesting that the mobility is limited by phonon scattering.21,22 Finally, we extracted a lower bound on the bandgap of a 6.8-nm-thick PdSe2 flake from an Arrhenius plot of the minimum drain-source current Id,min vs the inverse of temperature T by using
where Eg is the transport bandgap and k is the Boltzmann constant [Fig. 3(e)]. We extract a bandgap of 0.57 eV for a 6.8-nm-thick PdSe2 flake. Due to the contribution from trap states, this method is expected to underestimate the bandgap.23 Finally, to further confirm the optical measurements shown in Fig. 2(f), we performed temperature-dependent measurements on a PdSe2 transistor fabricated using a bulk (116-nm-thick) crystal, as shown in Fig. 3(f). Since the thickness of this device is significantly greater than the Debye screening length, the device cannot be fully turned off. Nevertheless, we still observe moderate gate-dependent transport due to the low background doping in PdSe2. This indicates that bulk PdSe2 is semiconducting and is consistent with the absorption measurements shown above.
We next investigate the performance of PdSe2 as a photoconductor utilizing the devices discussed in Fig. 3. Figure 4(a) shows the room-temperature spectral responsivities of the 6.8-nm- and 116-nm-thick PdSe2 devices shown in Fig. 3(a). We measured the spectral responsivity of these devices using an FTIR by replacing the internal detector with the PdSe2 devices and focusing its light source on the devices with a CaF2 lens. The internal deuterated triglycine sulfate (DTGS) in the FTIR was used to find the relative intensity of the light source and a NIST-traceable Ge photodiode to calibrate the illumination intensity of the light source.24 The responsivity is calculated as R(λ) = Iph(λ)/Pin(λ), where Iph is the photocurrent and Pin is the incident power on the device. The specific detectivity (D*) is then calculated using
where is the device area, Δf is the integration time, NEP is the noise equivalent power, and in2 = 2qIdΔf is the squared noise current, where Id is the dark current. Figure 4(b) shows the room-temperature spectral D* of the 6.8-nm- and 116-nm-thick PdSe2 devices. We find a room-temperature D* value of 1.8 × 1010 cm Hz1/2 W−1 at a wavelength of 1 μm for the 6.8-nm-thick PdSe2 device. From a plot of (EQE/hν)1/2 vs hν, where EQE is the external quantum efficiency we extracted 0.59 eV and 0.49 eV as the optical bandgaps for a 6.8-nm- and 116-nm-thick PdSe2 flake, respectively [Figs. 4(c) and 4(d)]. It is important to note that the photoresponse measurements shown were taken at Vg = 0. This was found to give the highest photoresponse and is consistent with the Id-Vg characteristics which show the lowest dark current for Vg near zero.
(a) Room-temperature spectral responsivities and (b) specific detectivities D* of the 6.8-nm- and 116-nm-thick PdSe2 devices shown in Fig. 3(a) measured at a drain voltage Vd of 1 V with zero gate voltage Vg applied. (c) and (d) Indirect bandgap extraction of a 6.8-nm-thick and a 116-nm-thick PdSe2 flake.
(a) Room-temperature spectral responsivities and (b) specific detectivities D* of the 6.8-nm- and 116-nm-thick PdSe2 devices shown in Fig. 3(a) measured at a drain voltage Vd of 1 V with zero gate voltage Vg applied. (c) and (d) Indirect bandgap extraction of a 6.8-nm-thick and a 116-nm-thick PdSe2 flake.
In conclusion, we have performed a systematic study on the optical and electrical properties of PdSe2, a layered two-dimensional pentagonal semiconductor with an indirect bandgap ranging from ∼0.5 to 1.37 eV. We found that a 6.8-nm-thick PdSe2 flake-based transistor has effective electron mobilities of 130 and 520 cm2 V−1 s−1 at 300 K and 77 K, respectively. The same transistor can serve as a SWIR photoconductive detector with a D* value of 1.88 × 1010 cm Hz1/2 W−1 at 1 μm with a cutoff wavelength of 1.94 μm. At 1 μm, PdSe2 has a D* value on the same order of magnitude as 2D tellurium (Te)23 and black phosphorous (bP).24 Te and bP, however, have optimized cutoff wavelengths of 3.4 μm and 4.6 μm, respectively, allowing more coverage of the SWIR band compared to PdSe2. Nonetheless, given the promising preliminary values of D* for our simple PdSe2 device structure, further optimization, e.g., optical cavity substrate engineering to shift the peak D* and cutoff wavelength, may allow the use of PdSe2 for high-performance SWIR photodetectors.23 Most importantly, through different measurements, we have determined that PdSe2 does not undergo a semiconductor to metal transition with the increasing thickness. This finding changes the outlook of the potential application space in which PdSe2 may be used without further manipulating the material's band structure, e.g., by strain25,26 or defect27 engineering.
See the supplementary material for additional material characterization.
Device fabrication and measurements were supported by the Defense Advanced Research Projects Agency under Contract No. HR0011-16-1-0004. Synthesis work was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, Materials Sciences and Engineering Division under Contract No. DE-AC02-05CH11231 within the Electronic Materials Program (KC1201). The work at the Molecular Foundry was supported by the Office of Science, Office of Basic Energy Sciences, of the U.S. Department of Energy under Contract No. DE-AC02-05CH11231. H.Z. thanks the support from ITC via the Hong Kong Branch of National Precious Metals Material Engineering Research Center and the Start-Up Grant from the City University of Hong Kong.