Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Unveiling the potentiality of a self-powered CGT chalcopyrite-based photodetector: theoretical insights

Open Access Open Access

Abstract

The article demonstrates the design and modelling of CuGaTe2 direct bandgap (1.18 eV) chalcopyrite-based photodetector (PD), which has superb optical and electronic characteristics and shows remarkable performance on the photodetector. The photodetector has been investigated throughout the work by switching width, carrier and defect densities of particular layers and also the interface defect density of particular interfaces. The various layers have been optimized for the higher performance of the PD. Also, the impression of various device resistances has been analyzed. The JSC and VOC of the heterostructure photodetector is found to be 38.27 mA/cm2 and 0.94 V, in turn. The maximum responsivity, R and detectivity, D* are found to be 0.663A/W and 1.367 × 1016 Jones at a wavelength of 920 nm. The spectral response has a very high value in the range of 800 to 1000 nm light wavelength, which confirmed that this device is capable of detecting the near infrared (NIR) region of light. This work gives important guidance for the manufacture of CGT material-based photodetectors with higher performance.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

A photodetector (PD) is a very special form of optoelectronic device which converts the illuminated light or other electromagnetic radiation from the various type of lights such as UV, visible and infrared radiation into electrical signals. The PD is the core of a contemporary optoelectronic technology, e.g. artificial intelligence (AI), spectroscopy and automated vision. It is sometimes referred to as a photosensor or light detector. The primary job of a photodetector is to transform light signals into electrical currents or voltages that are generally proportionate to the falling light intensity or wavelength. As a consequence of the multi-spectral characteristic of the light signal, the identification of UV, visible and infrared light in the same photodetectors have developed with fast, wideband, ample area, cheap and energy-saving qualities [14]. Photodetectors (PDs) have been extensively employed for a number of commercial and scientific uses, including fiber optic data transmission, ambient surveillance, day-night vision, and bio-chemical detection, medical diagnostics, aviation, target identification, missile warning and other areas [58]. Photodetectors can be divided into a variety of classes based on their detection range, including IR, visible, and UV. An appropriate bandgap material must be used in order to create appropriate photodetectors [810]. The III-V compounds or silicon (Si) are basically used in the high-performance commercial photodetectors. But these PDs face some major problems such as relatively low responsivity and detectivity, high driving voltage and costly manufacturing process. That is why it is needed to enhance the analysis on new materials for having better photodetector performances [6].

The narrow band sensitivity photodetector has been accomplished using organohalide [3,4], organic materials [1113], Mercury selenide (HgSe) quantum dots (QDs) [14], lead sulfide (PbS) QDs [12], GaN/InGaN quantum well [2] and silicon (Si) [15]. Additionally, there has been no research conducted to date on a photodetector that possesses both high speed and energy efficiency and is specific to light signals with a broad wavelength range [1]. Self-powered PD, which can detect light in the absence of an exterior supply of power, have recently sparked a huge attention. By harnessing the photovoltaic (PV) outcome of a pn or Schottky junction when illumined, self-powered devices can function in a solitary state [16]. A driving force for the splitting of highly effective photo-carriers and the generation of sustained photocurrent will be provided by the electric field developed in an effective heterojunction between various materials. Magnificent photoelectric performances including a rapid reaction time, extensive linear operating range and minimal noise are displayed by PDs with self-powered characteristics based on pn junctions and have made notable advancements [17].

In recent years, a widespread research and exploration of Cu-III-VI2 (III = Al, Ga, In; VI = S, Se, Te) chalcopyrite semiconducting compounds are being conducted due to their promising potential for use in PV cells and optoelectronic devices. Within these materials, CuGaTe2 (CGT) has been proven itself as a standard intrinsic p-type direct bandgap chalcopyrite semiconducting compound with high current conductivity and a significant absorption coefficient of light [18]. Its structure is analogous to the zinc blende ZnTe cubic structure, within which Zn and Te are substituted by Cu and Ga atoms in alteration in the place between Cu and Ga elements [19]. Deposition of stoichiometric and homogenous layers is the key need for the production of a device using CGT thin films [20]. Different techniques have been used so far to deposit the CGT thin films, namely, flash evaporation [21,22], co-evaporation [23], and close spaced vapor transport technique (CSVT) employing iodine (I) as a transportation agent [24]. The hindmost method may create thin films with big grains (a few micrometers), huge surface area and cheap cost [25,26]. However, the potentiality of the CuGaTe2 as a photodetector has not been discovered yet. The innocuous and cost-effective CGT with high light absorbance makes this material a promising one in photodetector sector.

Additionally, CdS is a well-known n-type semiconductor that has earned substantial global attention due to its adjustable wide optical bandgap, elevated carrier density and excellent stability even when exposed to continuous light, coupled with high transmission properties [2729]. Also, the CdS can be made by various techniques for the application in heterostructure photodetectors [30].

Moreover, GeS is a wide bandgap group-IV monochalcogenide that attracts researchers for its distinctive characteristics [31]. GeS possesses benefits such as excellent chemical and environmental stability, affordability, abundant presence of earth-friendly elements and good energy band alignment by the adequate bandgap, electron affinity and higher carrier mobility, rendering it a promising substance for use in photovoltaic systems, photodetectors, and lithium-ion batteries [31,32]. The GeS thin films are mainly produced through the fabrication of germanium monosulfide at a temperature of 725 K, keeping the substrates at room temperature (300 K) under a pressure of 10−5 torr according to the literature [33]. That is why p-type GeS has the potential to be used as the back surface field (BSF) layer in the self-powered heterostructure PD.

In this work, a self-powered n-CdS/p-CuGaTe2/p + -GeS heterostructure has been proposed as a photodetector which is computationally inspected by the mostly used one-dimensional software SCAPS-1D. In the structure, CdS and GeS act as the buffer and BSF layers, respectively. The influence of width, density of defects and doping density on the operation of the schemed photodetector has been probed rigorously. The simulation results illustrate a promising route to further expand the understanding of CuGaTe2-based optoelectronic devices.

2. Structure of the device and computation

The one-dimensional SCAPS-1D (version 3.3.09) simulator has been employed for simulating and analyzing the characteristics of the schemed CGT heterostructure photodetector [34,35]. Its dependability has been thoroughly tested, and it has been effectively applied in real-world situations. The outputs of this simulation are comparable to those of the experimental investigation according to the literatures [36,37]. Basically, SCAPS-1D addresses three equation sets, namely Poisson’s equation, hole and electron continuity equations while adhering to specified boundary conditions [38]. Equations (1)–3 represented these three equations, respectively given below and can be solved in a self-consisted manner by iteration method:

$$\displaystyle{{\partial ^2{\rm \Psi }} \over {\partial {\rm x}^2}} + \displaystyle{{\rm q} \over {\rm \varepsilon }}{\rm \; }\left[ {{\rm p}\left( {\rm x} \right)-{\rm n}\left( {\rm x} \right) + {\rm N}_{\rm D} + {\rm N}_{\rm A} + {\rm \rho }_{\rm p}-{\rm \; }{\rm \rho }_{\rm n}} \right] = 0$$
$$\frac{1}{\textrm{q}}\frac{{\partial {\textrm{J}_\textrm{p}}}}{{\partial \textrm{x}}} = {\textrm{G}_{\textrm{op}}} - \textrm{R}(\textrm{x} )$$
$$\frac{1}{\textrm{q}}\frac{{\partial {\textrm{J}_\textrm{n}}}}{{\partial \textrm{x}}} ={-} {\textrm{G}_{\textrm{op}}} + \textrm{R}(\textrm{x} )$$
where, the letters q, ε and $\mathrm{\Psi }$ stands for electronic charge, dielectric constant, electrostatic potential, in turn. The symbols ${\rho _p}$ and ${\rho _n}$ render the smearing of hole and electron, in turn; while ${N_D}$ and ${N_A}$ denote the densities of electron and hole, in turn. Furthermore, ${J_p}$ and ${J_n}$ stand for the hole and electron currents, respectively; ${G_{op}}$ and R express the optical carrier production and recombination rates. Also, notations p and n are used for the free acceptor and donor concentrations. The position coordinate x serves as the function for each of these parameters.

Figure 1(a) illustrates the architectural arrangement of the n-CdS/p-CuGaTe2/p + -GeS heterostructure NIR range photodetector where Mo and Al act as anode and cathode, respectively and Fig. 1(b) delimitates the consecutive band structure of every layer. In this structure, the narrow bandgap material CuGaTe2 is used as an absorber layer. Also, the ITO (Indium doped tin oxide), n-CdS (Cadmium Sulfide) and p + -GeS (Germanium Sulfide) are used as transparent conductive oxide layer, buffer layer and BSF in the photodetector structure.

 figure: Fig. 1.

Fig. 1. (a) Architectural arrangement and (b) band schematic of the n-CdS/p-CuGaTe2/p + -GeS photodetector.

Download Full Size | PDF

Responsivity, R (A/W) and Detectivity, D* (Jones) are two crucial factors to consider for evaluating a device's performance and the photodetector’s sensitivity. The R and D* might be determined by the subsequent equations [39]:

$$R = \textrm{}\frac{{\mathrm{q\eta \lambda }}}{{\textrm{hc}}}$$
$${\textrm{D}^\ast } = \frac{\textrm{R}}{{\sqrt {2q{J_0}} }}$$
where, η denote the quantum efficiency (QE), λ stands for the light wavelength, h indicates the Planck's constant, c denotes the speed of light (3 × 108 m/s), q denotes the electron charge (1.6 × 10−19 C) and J0 denote the reverse saturation or dark current.

An overview of the various physical properties of the materials employed in this work is presented in Table 1 that have been derived from prior reports and established theories. For simplifying the numerical analysis, the electron and hole thermal velocity of 107 cms-1 were chosen. On the other hand, Table 2 shows detailed insights into interface properties and defect density. For all the studies, AM 1.5 G illumination with 1000 W/m2 from the ITO layer side were utilized for optimizing the PD device.

Tables Icon

Table 1. The physical specifications used for each layer of the CGT photodetector

Tables Icon

Table 2. The specifications used for the individual interface of the CGT photodetector

3. Results and discussion

3.1 Impression of the thickness of specific layers on the CGT photodetector

3.1.1 Significance of the CGT absorber layer

The breadth of the CGT photon absorbing layer has been switched from 0.2 to 2 µm to determine the optimum thickness, keeping other parameters unchanged as alluded in Table 1 which is shown in the Fig. 2.

 figure: Fig. 2.

Fig. 2. The variation of different performance parameters of CGT PD with CGT thickness: (a) VOC and JSC, (b) QE, (c) Responsivity and (d) Detectivity.

Download Full Size | PDF

From Fig. 2(a), it is observed that the VOC of the device insignificantly lower with a value of 0.93 V and the JSC are and 30.34 mA/cm2 for the breadth of the absorber layer of 0.2 µm. With the variation of the thickness, the VOC slightly increases whereas JSC significantly increases up to the thickness of 0.6 µm. At this thickness, the values of VOC and JSC are 0.94 V and 38.27 mA/cm2, respectively. The limited light absorption by the thin absorber layer leads to a reduced photocurrent, despite the effective transport of most photo-carriers to their respective electrodes that lessens the VOC [5]. However, the JSC noticeably rises with the rise in width as more and more electron-hole pairs are created by the absorption of more light in the layer. Whenever the width of the CGT layer further escalates, the value of VOC slightly declines and the value of JSC saturates. The reason for decreasing the VOC is that the thicker absorber layer also raises the reverse saturation current of the device that reduces VOC. The JSC almost saturates when the thickness further increases as because the recombination starts to rule over the absorbing photons [5,8]. For this reason, the breadth of the CGT layer is set to 0.6 µm for further computations.

The Table 3 shows the variation of VOC and JSC values with thickness of absorber layer ranging from 0.2 to 2 µm. Although, the JSC slightly escalates with thickness after 0.6 µm, but the VOC diminishes. Therefore, the optimal thickness of the CGT layer has been set to 0.6 µm for further study.

Tables Icon

Table 3. Variation of VOC and JSC correlating with thickness of CGT absorber

Figure 2(b) demonstrated the variation of Quantum Efficiency (QE) with wavelength for different values of thickness of CGT absorber. It is seen that the QE is much less for the widht up to 0.4 µm. But for the absorber breadth of 0.6 to 2 µm, a very small change in the QE is noticed. This is because of the recombination between the carriers at the thicker absorber layer results in saturated JSC and therefore the QE does not change so much [8].

Figure 2(c) and (d) illustrates the R and D* of the PD with the variation of wavelength for various thicknesses of CGT layer. The responsivity and detectivity advance by the escalation of thickness from 0.2 to 2 µm of the absorber. It is seen from the graph the thickness imposes a noticeable effect on the R and D* at the wavelengths near the bandgap edge of the CGT absorber. This phenomenon is occurred as the photon energies near the bandgap edge are the preconditioned energies for the effective absorption that provide a route to the higher photocurrent [6]. Herein, at the optimized thickness of 0.6 µm, the maximum responsivity and detectivity are 0.663 A/W and 1.367 × 1016 Jones, in turn at the photon wavelength of 920 nm and which indicates that this photodetector is useful for near-infrared (NIR) region.

3.1.2 Significance of the CdS and GeS layer thickness

The significance of the variation of thickness of the buffer and BSF layers in the heterostructure photodetector is depicted in Fig. 3(a) and (b), respectively. The thickness of these layers has been varied from 0.1 to 1 µm. From the figure, it can be seen that the effect of the thickness of the GeS and CdS is not noticeable. The VOC and JSC are almost unchanged with the change of the thicknesses of these layers. This is happened because the lifetime of the minority carriers and diffusion length are relatively high compared to the thicknesses of the CdS and GeS, therefore carrier recombination in the layer does not significantly impact on the device action [48]. The optimized thickness for the CdS and GeS layer are 0.1 and 0.2 µm, respectively that attain the voltage of 0.94 V and current of 38.27 mA/cm2.

 figure: Fig. 3.

Fig. 3. The variation of VOC and JSC of the CGT PD with the width of (a) CdS, and (b) GeS buffer and BSF layers, respectively.

Download Full Size | PDF

3.2 Impact of the doping of specific layers on the photodetector

The doping density (acceptor of donor) of the absorber, buffer and BSF layer is varied from 1011 to 1020, 1013 to 1022 and 1013 to 1022 cm-3, respectively to observe the consequence of doping on the voltage and current of the CGT photodetector.

The CGT PD device exhibits almost constant voltage and current nature of the with the shallow acceptor doping of the CuGaTe2 layer in the span from 1011 to 1016 cm-3 as observed in Fig. 4(a). Further increment on the doping concentration results in decreased VOC and JSC that may result due to the rise in recombination current and increase in diode parameter. After the carrier density of 1018 cm-3, the VOC is increased and it results to 1 V which may happen due to the rise in built-in voltage as well as decrease in diode ideality parameter [5,48]. However, some previous reports indicate that the absorber doping concentration should not exceed the experimental value of 1018 cm-3 [49]. Therefore, the optimum shallow acceptor doping concentration is chosen as 1015 cm-3 for further calculation by conserving all parameters steady as shown in Table 1.

 figure: Fig. 4.

Fig. 4. The deviation of VOC and JSC with the carrier concentration of (a) absorber, (b) CdS Buffer, (c) GeS BSF layers, respectively and (d) Total recombination profile.

Download Full Size | PDF

From Fig. 4(b) and (c) delineate the impression of doping density of the buffer and BSF layers. The impression of donor concentration of the CdS buffer is not noticeable for JSC, but it shows an effect on the VOC. The VOC is almost steady up to the doping concentration of 1018 cm-3. But further increment in doping of buffer results in higher VOC. This is happened because the higher doping concentration of the buffer layer results in higher built-in voltage across the CdS/CGT heterojunction and thus enhances the device performance and reduced the non-radiative recombination [8]. The consequence of acceptor doping in the BSF layer is not so much noticeable in the case of VOC and JSC. However, to lower the interface recombination at the CuGaTe2/GeS interface, high BSF doping is very much appreciable to generate a robust electric field across the interface that obstructs the movement of minority charge carriers in the particular interface direction as well as reduce surface recombination [8,48]. However, the BSF and buffer layer doping concentration has been taken as 1018 cm-3 for further calculations by holding all the parameters stable as delimitated in Table 1.

Figure 4(d) illustrates the recombination profile due to the doping. It is seen from the figure that, the interface as well as bulk recombinations increase in the window/absorber interface as well CdS window and CuGaTe2 absorber layers with doping up to 1018 cm-3. These recombinations raise the dark current and as a result the VOC of the device decreases [5,48]. However, the increase in dark current may also insignificantly decreases the JSC. The recombination is lower in the span of 1018 to 1020 cm-3 of doping that results in the increasing VOC. Therefore, the VOC and JSC exhibit similar behavior for alteration in carriers in CdS and CGT layers. However, no such effect can be seen in the BSF.

3.3 Dominance of the defect density of specific layers on the photodetector

3.3.1 Significance of the CGT absorber layer

The defect density of CGT photon absorbing layer has an essential role in the operation of the heterostructure photodetector. The functioning of the photodetector is affected by the bulk defects of the constituent layers. The bulk or volume defects of CGT is altered from 1010 to 1019 cm-3.

The dominance of volume defects of the CGT layer on JSC and VOC is delineated in Fig. 5(a). The VOC of the PD device is almost constant at 0.94 V for the defects in the span from 1010 to 1014 cm-3. And, the JSC is almost constant from the defects of 1010 cm-3 to 1017 cm-3 and that is 38.27 mA/cm2. This is occurred as the lifetime of the carriers and length of diffusion are higher in this defect range that results lower recombination current in the device [48]. Further increment of the defect density results a rapid decrease in VOC and JSC and therefore the decrease in the PD performance. That is because the escalating defect concentration of the absorber layer ramifications in huge recombination of photo-carriers and this recombination vitally diminished the performance of the photovoltaic device [5,48].

 figure: Fig. 5.

Fig. 5. The impact of bulk defects of CGT absorber on (a) VOC and JSC, (b) QE, (c) Responsivity and (d) Detectivity of the CGT photodetector.

Download Full Size | PDF

Figure 5(b) delineates the fluctuation of the quantum efficiency (QE) of the CGT photodetector with the density of defects of the CGT chalcopyrite. Up to the defects of 1017 cm-3, the QE is almost constant at a value of 89.63%. Further increment of defects significantly decreases the QE of the PD device. The higher order of defects significantly reduces the carrier lifetimes and diffusion length and increase carrier recombinations in the CGT layer that reduce the JSC and therefore the QE of the PD [8,48].

Figure 5(c) and (d) demonstrate the responsivity and detectivity graph for different amounts of defects in the CGT layer. The density of defects in the CGT absorber has been from 1010 to 1019 cm-3. The R and D* determine the photo detection region of the CGT heterostructure photodetector. Up to the defects level of 1017 cm-3, the values of R and D* are almost constant at 0.663 A/W and 1.367 × 1016 Jones, respectively as can be perceived from the figures. Further increment of defects highly reduces the responsivity and detectivity of the PD. This is happened because the rate of recombinations at the localized energy states due to higher defects increases that reduce the performance of the photodetector for detecting the near infrared (NIR) region [5].

Therefore, density of defects at the order of 1014 cm-3 has been appraised for CGT absorber as the optimized defects that are feasible to achieve from the manufacturing point of view.

3.3.2 Significance of the CdS and GeS layer

Figure 6 illustrates the impression of the bulk defects of CdS buffer and GeS BSF layers, respectively on the operation of n-CdS/p-CGT/p + -GeS photodetector.

 figure: Fig. 6.

Fig. 6. Variation of VOC and JSC of CGT PD with the bulk defects of (a) CdS buffer and (b) GeS BSF layers, respectively.

Download Full Size | PDF

It is noticed from Fig. 6(a) that the buffer layer doesn’t show so much change on VOC and JSC with respect to the volume defects in the layer. The constant values of VOC and JSC are found to be 0.94 V and 38.27 mA/cm2, respectively. However, rise in defects of CdS layer hugely affects the performance of the PD. That is because the recombination rate increases owing to the rise in local defect states in the CdS film that can turn out the performance [5]. However, at a very high defect of 1019 cm-3, the VOC rapidly increases that may results from the decrease in diode parameters with defects [48].

It is discerned from Fig. 6(b) that the functioning of the bulk defects in GeS BSF is not so much noticeable till to the value of 1018 cm-3. The VOC and JSC almost constant at 0.94 V and 38.27 mA/cm2, respectively. However, further surge in defects of the GeS BSF sharply decrease the VOC and JSC due to the decrease in lifetime and diffusion length of carriers [48].

However, the bulk defects of the order of 1014 cm-3 have been considered for both the CdS and GeS layers for the higher performance of the heterostructure CGT photodetector.

3.4 Role of interface defects on the photodetector

Figure 7 illustrates the role of defect density of CdS/CGT and CGT/GeS interfaces on VOC and JSC, respectively. The interface defect density of these interfaces has been varied from 107 to 1016 cm-3 for determining the consequence on VOC and JSC.

 figure: Fig. 7.

Fig. 7. Fluctuation of VOC and JSC with interface defects at (a) CdS/CGT and (b) CGT/GeS hetero-interfaces, respectively.

Download Full Size | PDF

The interface defect density has significant role on the context of VOC as delimitated in Fig. 7(a) and (b) for the buffer/absorber and absorber/BSF interfaces. When the value of interface defect density for CdS/CGT and CGT/GeS interfaces is up to 1011 cm-3, the voltage is nearly unchanged at a value of 0.94 V. However, the device current is also constant for this variation of interface defects and retains at 38.27 mA/cm2. Whenever the interface defect density is further increased, the VOC is decreased continuously. The significant increment in trap states due to interface defects that act as the recombination centers uplift the reverse saturation and lower the action of the heterostructure photodetector [8]. Therefore, the density of interface defects of 1011 cm-3 is appraised for both the CdS/CGT and CGT/GeS hetero-interfaces for further numerical calculations.

3.5 Impact of resistances on the photodetector

The main energy losses of a device are caused by the resistors. The diode voltage drops due to ohmic resistance in the body that concerns about the constituent layers in optoelectronic components. Herein, the function of the series resistance, RS and shunt resistance, RSh has been observed on the J-V characteristics of the CGT heterostructure photodetector. These two types of resistances work as parasitic elements in optoelectronic devices. The RS and RSh of the device should be minimum and infinite, respectively in optimum conditions [8]. But in practical analysis, the RS appears because of the contact among various layers along with the anterior and rear metal contacts and the recombination in defects causes of the shunt resistances and it is increased with the defect density decreases [8]. Figure 8(a) depicts the impression of RS resistance on the J-V curves of newly designed CGT heterostructure photodetector. Here, the RS resistance is altered from 0 to 10 Ω.cm2. From the figure, it is seen that when RS is at default value which is 0 Ω.cm2, the current density is maximum. When the RS increases significantly, the current density of the device slightly decreases.

 figure: Fig. 8.

Fig. 8. The role of (a) Series Resistance, and (b) Shunt Resistances on the J-V curves of the CGT PD.

Download Full Size | PDF

The functioning of shunt resistance on the J-V characteristics of CGT heterostructure PD is delimitated in Fig. 8(b). The shunt resistance has been altered from 1 × 105 to 5 × 105 Ω.cm2. It is perceived from the figure that the change in shunt resistance in this range does not have any influence on the operation of the PD device. From the analysis, it can be concluded that lofty shunt resistance and lowest series resistance offers higher photodetector performances [8]. That is why, the RS and RSh resistances of default value and 5 × 105 Ω.cm2, in turn have been chosen for the computations.

3.6 Optimized self-powered CGT photodetector

In this section, the overall performance of the CGT photodetector has been discussed for the optimized condition.

Figure 9(a) and (b) depict the current-voltage (J-V) curves without and with the GeS BSF layer under dark and illumination. Under no illumination, the device shows a very small amount of dark current of the order of 7.35 × 10−12 A/cm2 without and with the GeS BSF layer. This is happened because the CGT PD exhibits a normal rectifying J-V characteristics owing to the heterojunction formation at the interface among various layers [5]. But under the illuminations, a very much increment of the current is observed and the value of current is reached to 38.27 mA/cm2. However, the current density is almost same under light with and without the GeS BSF. On the contrary, the voltage is very much increased with the back surface layer in the heterostructure photodetector. Without the GeS BSF, the highest voltage is 0.6 V and the addition of GeS BSF layer promotes the VOC to 0.94 V that makes this structure a promising one to use it as a self-powered photodetector.

 figure: Fig. 9.

Fig. 9. J-V curves of the optimized CGT photodetector: (a) Without BSF and (b) with BSF layer under dark and illumination conditions; (c) QE Spectrum, (d) Responsivity and (e) Detectivity of the optimized photodetector.

Download Full Size | PDF

Figure 9(c), (d) and (e) illustrate the Quantum Efficiency, Responsivity and Detectivity of the optimized heterostructure photodetector. These parameters are very much important to determine the quality, type and performance of photodetector. Responsivity declares the ability to convert the incident photons into current in variety of wavelengths which is very much dependable in QE [6]. The QE is more than 80% and maintains almost similar value in the NIR region. The ceiling value of responsivity is determined to be 0.663 A/W which is attained at a photon wavelength of 920 nm that is in the NIR region. Also, from the 800 nm to 1000 nm, the responsivity has a higher value in the range and thus it shows that this photodetector serves as NIR region photodetector. Again, the detectivity is also a very important characteristics of photodetector devices that signifies in determination of the weakest optical signals and it is basically employed for differentiating the performance to detect weak signals among photodetectors [8]. The maximum detectivity for the designed CGT based heterostructure photodetector determined at 920 nm wavelength of light and the value of D* is 1.367 × 1016 Jones. This photodetector has higher value of detectivity among the light wavelength of 800 nm to 1000 nm. This also indicates that this photodetector is a NIR region photodetector.

The optimum widths and carrier concentration for the CGT, CdS and GeS layers are 0.6, 0.1 and 0.2 µm and 1015, 1018 and 1018 cm-3, respectively.

3.6 Comparison of the performance of different thin film PD

Table 4 shows the comparative analysis of the proposed CuGaTe2 based heterostructure photodetector with other thin film photodetectors. Among the photodetectors, the proposed CGT structure provides a higher value of responsivity and detectivity that will helpful in further investigation towards photodetector.

Tables Icon

Table 4. Comparison table among different types of photodetectors

4. Conclusion

In this simulation, a newly designed n-CdS/p-CuGaTe2/p + -GeS heterostructure-based photodetector which is modelled, computed and optimized with SCAPS-1D. The impression of tunable parameters such as width, carriers and defects of particular layers and also defects at particular interfaces and functioning of series and shunt resistances have been investigated. This heterostructure PD shows a very high JSC of 38.27 mA/cm2 and VOC of 0.94 V and promising photo response. The results show that the responsivity and detectivity have very higher values in the 800 to 1000 nm light wavelength range which illustrate that this device is capable of detecting near infrared (NIR) region of light. The highest values of responsivity and detectivity are 0.663 A/W and 1.367 × 1016 Jones, respectively at a light wavelength of 920 nm. These higher values will make this material a novel and promising candidate in the optoelectronics-based work. This simulation work provides an important direction for designing self-powered, highly sensitive, wide band CGT based cost-effective photodetector in practical applications and it will be a great interest towards researchers for further investigations.

Acknowledgements

The authors are highly indebted to Prof. Dr. Marc Burgelman, University of Gent, Belgium, for supplying SCAPS 1D simulator.

Disclosures

The authors declare no competing financial interest.

Data availability

Simulation details and associated data are available free of charge from authors upon reasonable request.

References

1. M. Patel, M. Kumar, and J. Kim, “Polarity flipping in an isotype heterojunction (p-SnS/p-Si) to enable a broadband wavelength selective energy-efficient photodetector,” J. Mater. Chem. C 6(26), 6899–6904 (2018). [CrossRef]  

2. C. Yang, S. P. Turaga, A. A. Bettiol, et al., “Textured V-Pit green light emitting diode as a wavelength-selective photodetector for fast phosphor-based white light modulation,” ACS Photonics 4(3), 443–448 (2017). [CrossRef]  

3. Q. Lin, A. Armin, P. L. Burn, et al., “Near infrared photodetectors based on sub-gap absorption in organohalide perovskite single crystals,” Laser Photon. Rev. 10(6), 1047–1053 (2016). [CrossRef]  

4. Q. Lin, A. Armin, P. L. Burn, et al., “Filterless narrowband visible photodetectors,” Nat. Photonics 9(10), 687–694 (2015). [CrossRef]  

5. H. Yao and L. Liu, “Design and optimize the performance of self-powered photodetector based on PbS/TiS3 heterostructure by SCAPS-1D,” Nanomaterials 12(3), 325 (2022). [CrossRef]  

6. Z. Hosseini, K. Jafarizadeh, and H. Amanati Manbar, “Assessment of mixed Tin-Lead perovskite as the absorber material for fabrication of highly sensitive broadband photodetector,” J. Optoelectronical Nanostructures 7, 29–48 (2022). [CrossRef]  

7. L. Dou, Y. Yang, J. You, et al., “Solution-processed hybrid perovskite photodetectors with high detectivity,” Nat. Commun. 5(1), 5404 (2014). [CrossRef]  

8. S. Singh, S. Kumar, M. Deo, et al., “Unveiling the potential of organometal halide perovskite materials in enhancing photodetector performance,” Opt. Quantum Electron. 55(10), 846 (2023). [CrossRef]  

9. A. Tahira, R. Mazzaro, F. Rigoni, et al., “A simple and efficient visible light photodetector based on Co3O4/ZnO composite,” Opt. Quantum Electron. 53(9), 534 (2021). [CrossRef]  

10. P. Rakshit and N. R. Das, “Effect of device parameters on improving the quantum efficiency of a lateral Si p–i–n photodetector,” Opt Quantum Electron. 52(8), 371 (2020). [CrossRef]  

11. L. Shen, Y. Fang, H. Wei, et al., “A highly sensitive narrowband nanocomposite photodetector with gain,” Adv. Mater. 28(10), 2043–2048 (2016). [CrossRef]  

12. L. Shen, Y. Zhang, Y. Bai, et al., “A filterless, visible-blind, narrow-band, and near-infrared photodetector with a gain,” Nanoscale 8(26), 12990–12997 (2016). [CrossRef]  

13. W. Wang, F. Zhang, M. Du, et al., “Highly narrowband photomultiplication type organic photodetectors,” Nano Lett. 17(3), 1995–2002 (2017). [CrossRef]  

14. X. Tang, G. fu Wu, and K. W. C. Lai, “Plasmon resonance enhanced colloidal HgSe quantum dot filterless narrowband photodetectors for mid-wave infrared,” J. Mater. Chem. C 5(2), 362–369 (2017). [CrossRef]  

15. B. Yin, H. Zhang, Y. Qiu, et al., “The light-induced pyro-phototronic effect improving a ZnO/NiO/Si heterojunction photodetector for selectively detecting ultraviolet or visible illumination,” Nanoscale 9(44), 17199–17206 (2017). [CrossRef]  

16. Z. Zhang, W. Zhang, Z. Wei, et al., “Dipole-templated homogeneous grain growth of CsPbIBr2 films for efficient self-powered, all-inorganic photodetectors,” Sol. Energy 209, 371–378 (2020). [CrossRef]  

17. M. Long, P. Wang, H. Fang, et al., “Progress, challenges, and opportunities for 2D material based photodetectors,” Adv. Funct. Mater. 29(19), 1803807 (2019). [CrossRef]  

18. L. Wang, Z. Li, M. Li, et al., “Self-powered filterless narrow-band p-n heterojunction photodetector for low background limited near-infrared image sensor application,” ACS Appl. Mater. Interfaces 12(19), 21845–21853 (2020). [CrossRef]  

19. Z. Zhang, Y. Gao, Y. Wu, et al., “P-type doping of transition metal elements to optimize the thermoelectric properties of CuGaTe2,” Chem. Eng. J. 427, 131807 (2022). [CrossRef]  

20. O. Abounachit, H. Chehouani, and K. Djessas, “Effects of temperature, pressure and pure copper added to source material on the CuGaTe2 deposition using close spaced vapor transport technique,” Thin Solid Films 540, 58–64 (2013). [CrossRef]  

21. M. S. Reddy, K. T. R. Reddy, B. S. Naidu, et al., “Electrical properties of CuGaTe2, thin films,” Mater. Chem. and Phys. 44(3), 281–283 (1996). [CrossRef]  

22. H. Neumann, W. Hörig, E. Reccius, et al., “Growth and optical properties of CuGaTe2 thin films,” Thin Solid Films 61(1), 13–22 (1979). [CrossRef]  

23. P. Guha, S. Roy, S. Chaudhuri, et al., “Synthesis and characterization of CuGaTe2 films prepared by three source co-evaporation technique,” J. Phys. D: Appl. Phys. 35(13), 3091504 (2002). [CrossRef]  

24. G. Massé, L. Yarzhou, and K. Djessas, “Préparation et étude de couches minces de CuXY2 (X = Ga, In; Y = Se, Te) pour applications en cellules solaires,” J. Phys. III France 3(11), 2087–2099 (1993). [CrossRef]  

25. M. Nouiri, K. Djessas, J. L. Gauffier, et al., “Effect of the growth temperature on the structural, morphological and electrical properties of CuIn0.7Ga0.3Se2 layers grown by CSVT technique,” Thin Solid Films 516(20), 7088–7093 (2008). [CrossRef]  

26. G. Massé and K. Djessas, “Thermodynamical study of the preparation of CuInSe2 thin films in vertical closed tube systems,” Thin Solid Films 237(1-2), 129–133 (1994). [CrossRef]  

27. J. Han, G. Fu, V. Krishnakumar, et al., “Studies of CdS/CdTe interface: Comparison of CdS films deposited by close space sublimation and chemical bath deposition techniques,” Thin Solid Films 582, 290–294 (2015). [CrossRef]  

28. T. Sinha, D. Lilhare, and A. Khare, “Temperature-dependent structural, morphological and optical properties of chemical bath-deposited CdS films,” Rare Met. 40(3), 701–707 (2021). [CrossRef]  

29. T. Sinha, L. Verma, and A. Khare, “Variations in photovoltaic parameters of CdTe/CdS thin film solar cells by changing the substrate for the deposition of CdS window layer,” Appl. Phys. A 126(11), 867 (2020). [CrossRef]  

30. S. K. Mostaque, A. Kuddus, M. F. Rahman, et al., “Solution-processed photodetectors, in: Handbook of II-VI Semiconductor-Based Sensors and Radiation Detectors,” Springer International Publishing, Cham: 427–452 (2023).

31. G. Guo and G. Bi, “Effect of tensile strain on the band structure and carrier transport of germanium monosulphide monolayer: a first-principles study,” Micro Nano Lett. 13(5), 600–605 (2018). [CrossRef]  

32. M. Grodzicki, A.K. Tołłoczko, D. Majchrzak, et al., “Band alignments of GeS and GeSe materials,” Crystals 12(10), 1492 (2022). [CrossRef]  

33. A. Stanchev and C. Vodenicharov, “Dielectric properties of thin germanium monosulphide films,” Thin Solid Films 29(1), L13–L16 (1975). [CrossRef]  

34. M. Burgelman, K. Decock, A. Niemegeers, et al., “SCAPS manual.” https://scaps.elis.ugent.be/SCAPS%20manual%20most%20recent.pdf

35. M. Burgelman, K. Decock, S. Khelifi, et al., “Advanced electrical simulation of thin film solar cells,” Thin Solid Films 535, 296–301 (2013). [CrossRef]  

36. L. Huang, X. Sun, C. Li, et al., “Electron transport layer-free planar perovskite solar cells: Further performance enhancement perspective from device simulation,” Sol. Energy Mater. Sol. Cells 157, 1038–1047 (2016). [CrossRef]  

37. K. Decock, P. Zabierowski, and M. Burgelman, “Modeling metastabilities in chalcopyrite-based thin film solar cells,” J. Appl. Phys. 111(4), 043703 (2012). [CrossRef]  

38. M. I. R. Ebon, M. H. Ali, M. D. Haque, et al., “Computational investigation towards highly efficient Sb2Se3 based solar cell with a thin WSe2 BSF layer,” Eng. Res. Express 5(4), 045072 (2023). [CrossRef]  

39. M. Kaifi and S. Gupta, “Simulation of perovskite based solar cell and photodetector using SCAPS software,” Int. J. Eng. Res. Technol. 10, 1778–1786 (2019).

40. B. K. Mondal, S. K. Mostaque, and J. Hossain, “Theoretical insights into a high-efficiency Sb2Se3-based dual-heterojunction solar cell,” Heliyon 8(3), e09120 (2022). [CrossRef]  

41. M. M. A. Moon, M. H. Ali, M. F. Rahman, et al., “Investigation of thin-film p-BaSi2/n-CdS heterostructure towards semiconducting silicide based high efficiency solar cell,” Phys. Scr. 95(3), 035506 (2020). [CrossRef]  

42. M. H. Ali, A. T. M. Saiful Islam, M. D. Haque, et al., “Numerical analysis of FeSi2 based solar cell with PEDOT:PSS hole transport layer,” Mater. Today Commun. 34, 105387 (2023). [CrossRef]  

43. G. Massé, K. Djessas, and L. Yarzhou, “Study of CuGa(Se,Te)2 bulk materials and thin films,” J. Appl. Phys. 74(2), 1376–1381 (1993). [CrossRef]  

44. K. V. Reddy and J. L. Annapurna, “Optical absorption and electrical conductivity of flash evaporated CuGaTe2 thin films,” Pramana - J. Phys. 26(3), 269–276 (1986). [CrossRef]  

45. W. Hörig, G. Kühn, W. Möller, et al., “Electrical and optical properties of CuGaTe2,” Kristall Und Technik 14(2), 229–242 (1979). [CrossRef]  

46. J. D. Wiley, A. Breitschwerdt, and E. Schönherr, “Optical absorption band edge in single-crystal GeS,” Solid State Commun. 17(3), 355–359 (1975). [CrossRef]  

47. X. Lv, W. Wei, Q. Sun, et al., “Two-dimensional germanium monochalcogenides for photocatalytic water splitting with high carrier mobility,” Appl. Catal. B. 217, 275–284 (2017). [CrossRef]  

48. M. M. Rana, A. T. Abir, S. S. Nushin, et al., “Numerical investigation on the role of ZnTe back surface layer in an efficient CuInS2 thin film solar cell,” Eng. Res. Express 5(4), 045020 (2023). [CrossRef]  

49. S. Sani, A. Usman, A. Bhatranand, et al., “A study on defect, doping, and performance of ETLs (ZnO, TiO2, and IGZO) for the lead-free CsSnCl3 perovskite solar cell by SCAPS-1D framework,” Mater. Today Commun. 38, 107575 (2024). [CrossRef]  

50. Z. Chen, Z. Kang, C. Rao, et al., “Improving Performance of Hybrid Graphene–Perovskite Photodetector by a Scratch Channel,” Adv. Electron. Mater. 5(6), 1900168 (2019). [CrossRef]  

51. H. Zhou, J. Mei, M. Xue, et al., “High-Stability, Self-Powered Perovskite Photodetector Based on a CH3NH3PbI3/GaN Heterojunction with C60 as an Electron Transport Layer,” J. Phys. Chem. C 121(39), 21541–21545 (2017). [CrossRef]  

52. A. M. Afzal, I.-G. Bae, Y. Aggarwal, et al., “Highly efficient self-powered perovskite photodiode with an electron-blocking hole-transport NiOx layer,” Sci. Rep. 11(1), 169 (2021). [CrossRef]  

53. A. Chetia, D. Saikia, and S. Sahu, “Design and optimization of the performance of CsPbI3 based vertical photodetector using SCAPS simulation,” Optik (Munich, Ger.) 269, 169804 (2022). [CrossRef]  

Data availability

Simulation details and associated data are available free of charge from authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (9)

Fig. 1.
Fig. 1. (a) Architectural arrangement and (b) band schematic of the n-CdS/p-CuGaTe2/p + -GeS photodetector.
Fig. 2.
Fig. 2. The variation of different performance parameters of CGT PD with CGT thickness: (a) VOC and JSC, (b) QE, (c) Responsivity and (d) Detectivity.
Fig. 3.
Fig. 3. The variation of VOC and JSC of the CGT PD with the width of (a) CdS, and (b) GeS buffer and BSF layers, respectively.
Fig. 4.
Fig. 4. The deviation of VOC and JSC with the carrier concentration of (a) absorber, (b) CdS Buffer, (c) GeS BSF layers, respectively and (d) Total recombination profile.
Fig. 5.
Fig. 5. The impact of bulk defects of CGT absorber on (a) VOC and JSC, (b) QE, (c) Responsivity and (d) Detectivity of the CGT photodetector.
Fig. 6.
Fig. 6. Variation of VOC and JSC of CGT PD with the bulk defects of (a) CdS buffer and (b) GeS BSF layers, respectively.
Fig. 7.
Fig. 7. Fluctuation of VOC and JSC with interface defects at (a) CdS/CGT and (b) CGT/GeS hetero-interfaces, respectively.
Fig. 8.
Fig. 8. The role of (a) Series Resistance, and (b) Shunt Resistances on the J-V curves of the CGT PD.
Fig. 9.
Fig. 9. J-V curves of the optimized CGT photodetector: (a) Without BSF and (b) with BSF layer under dark and illumination conditions; (c) QE Spectrum, (d) Responsivity and (e) Detectivity of the optimized photodetector.

Tables (4)

Tables Icon

Table 1. The physical specifications used for each layer of the CGT photodetector

Tables Icon

Table 2. The specifications used for the individual interface of the CGT photodetector

Tables Icon

Table 3. Variation of VOC and JSC correlating with thickness of CGT absorber

Tables Icon

Table 4. Comparison table among different types of photodetectors

Equations (5)

Equations on this page are rendered with MathJax. Learn more.

2Ψx2+qε[p(x)n(x)+ND+NA+ρpρn]=0
1qJpx=GopR(x)
1qJnx=Gop+R(x)
R=qηλhc
D=R2qJ0
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.