Reduced Recombination Current Due to Sputtered CdO Nanolayer at CdS/CdTe Interface

Abstract

In this paper, interface engineering via sputtering of CdO nanolayer at the buffer-CdS/CdTe-absorber interface is demonstrated as an efficient approach to improve the performance of solar cell device. The i-CdO interfacial layer with various thicknesses from 5 nm to 35 nm was deposited by DC magnetron sputtering. Comparative studies on TCO/CdS/CdTe and TCO/CdS/CdO/CdTe interfaces have been conducted by current-voltage, capacitance-voltage and admittance spectroscopy measurements. The current-voltage characteristics of the devices with an area of 0.45 cm2 under 100 mW/cm2 illumination, at the optimum thickness of CdO intermediate layer in the proposed structures, show increases of the short circuit current density and the open circuit voltage by 5% and 25%, respectively. The efficiency improvement of 3.1% of p-i-n cell over p-n cell is observed. Results of the temperature-dependent current-voltage and admittance measurements revealed the removing of the deep level defect with the activation energy of 0.43 eV and the reducing of the ideality factor from 1.9 to 1.8 via buffer/absorber interfacial passivation method. Interface passivation appears to be critical to improve the short circuit current density and the open circuit voltage, and CdO thin film is clearly effective for this purpose.

Share and Cite:

Al-Qassem, A. , Fedorov, V. , Gagara, L. and Potlog, T. (2023) Reduced Recombination Current Due to Sputtered CdO Nanolayer at CdS/CdTe Interface. Materials Sciences and Applications, 14, 186-207. doi: 10.4236/msa.2023.143011.

1. Introduction

CdTe solar cells with certified power conversion efficiency of 22.1% have attracted significant attention due to their excellent photovoltaic properties [1] . But there is still much potential for progress towards the 32% Shockley-Queisser limit [2] . The intensive research works are performed across various components of a photovoltaic device. For CdTe thin film technology, various buffer and interfacial layers, such as CdS, ZnO, Zn1–xSnxO (ZTO) or MgxZn1-xO, (MZO) have been successfully implemented for the interface engineering, but there is yet significant space for R&D in this direction [3] - [8] . A lot of publications [9] [10] [11] revealed that the CdS/CdTe interface is in the cliff category and is not favorable from the band-offset perspective. Therefore, this type of interface recombination will not be considered in this paper. Also, it is shown that polycrystalline CdTe have been limited to hole density on the order of 1014 cm−3, higher carrier recombination in the absorber and interfaces [12] [13] [14] [15] [16] . Among interfaces problems, the one between CdS and CdTe is imperative for efficient extraction, collection and recombination. Shockley-Read-Hall recombination due to defects or mismatches within the crystal lattice can create trap states that are spatially localized and energetically exist between the valence and conduction bands [17] [18] [19] . The mobile free carriers can be captured and bound to the traps. It can also be the case where a hole or an electron in the trap can recombine with a free electron or hole respectively, and then the trap is considered a recombination center. Surface recombination processes are due to defects that exist at semiconductor surfaces and interfaces. These defects create spatially localized states in the energy bands because of broken bonds, extrinsic impurities from the environment or as a result of the semiconductor growth/deposition process. It has been shown in a lot of publications that sulfur and telluride diffusion cannot improve device efficiency, despite minimizing lattice mismatch at the CdS/CdTe interface [20] [21] [22] [23] . This is because the number of misfit dislocations is conserved, and sulfur diffusion shifts the dislocations further into the space charge region, where the SRH recombination rate is higher. A way to increase efficiency of solar cells is by applying an intrinsic semiconductor broadening layer to the depletion region of a p-n junction. Literature survey [24] shows that in the amorphous silicon (a-Si) PV cells, the highest performance was realized by introducing the p-i-n structure. M. Hossain [25] simulated a CdS/CdTe solar cell with i-CdTe layer that improved efficiency to 26.74%. These successes mean the efforts to improve further device performance, and have shifted towards surface physical phenomena, mainly the passivation of surfaces and interfaces. Therefore, in this paper, the lack of interface engineering approach to CdTe will be discussed using sputtered CdO nanolayer. CdO is attracting tremendous attention due to its interesting properties like direct band gap of 2.3 eV [26] . We have found that among metal oxides, high electron mobility of CdO satisfies the essential requirements for a low resistance and high infrared transmission transparent electron layer. We assume that the electron affinity of the CdO film (4.5 eV) is one of the key parameters that tune the device performance. Since the value of ΔEc depends only by the nature of materials that form the heterojunction, the value of the ΔEc will change if the new n-window material is applied and, in the case of CdO, equal with −0.1 eV. The ΔEc = −0.1 eV means there is no barrier against the photo-generated current and the value of the short circuit current density (Jsc) is mainly determined by the band gap of CdTe and is independent of ΔEc value. A new wider thin band gap material, CdO on the traditional CdS window layer will improve the value of the Jsc of CdTe solar cells more than the value of the open circuit voltage. Also, due to the CdO large static dielectric constant (εo = 21.9), it has very high electron mobility because of reduced electron scattering through efficient screening of coulomb potentials of ionized donors. The low electrical resistivity of CdO thin films and high electron concentration of 1020 cm−3 allow high transmittance up to >1200 nm [26] .

2. Fabrication Processes and Characterization Methods

The CdS/CdTe heterojunctions were prepared by successive deposition of CdS and CdTe thin films by close space sublimation method onto glass substrates (2 × 2) cm2 covered with a transparent (~90%) conducting (~10–3 Ω∙m) ITO layer. For CdS and CdTe thin films deposition, CdS and CdTe powders 99.999% purity were purchased from Alfa Aesar, Germany. The effects of the source and substrate temperatures on the structure and electrophysical parameters of CdS and CdTe layers were studied to determine the optimum temperatures: Ts = 580˚C and Tsub = 390˚C for CdS and Ts = 535˚C and Tsub = 450˚C for CdTe. The CdS layer obtained in the optimal technological growth conditions had the thickness of 0.45 µm and the electron concentration of 1.6 × 1018 cm–3. The CdTe thin film had the thickness of 10.8 µm and the hole concentrations of 4.2 × 1014 cm–3. CdO layers were prepared by DC magnetron sputtering of a cadmium target in an oxygen atmosphere at a direct current. For sample preparation, a Cd target of 99.99% purity was used in the sputtering process. The sputtering chamber was evacuated down to 2 × 10−5 mbar. The distance between the target and the substrate was kept constant at 3 cm. The magnetron discharge was done in an O2 atmosphere via independent mass-flow controller. The working total pressure was kept at 5 × 10−3 mbar. The substrate temperature for deposition of CdO thin films was about 90˚C. The CdO layers thickness were controlled by adjusting the deposition time and determined from the calibration curve of the dependence of the CdO layer thickness on the sputtering duration. In all performed experiments the deposition rate was calculated using the dependence of the thickness layer on the deposition time. The CdO film deposition rate was 5 Ǻ/s. At an optimum total pressure of 5 × 10−3 mbar, the CdO films were polycrystalline and nearly stoichiometric with Hall mobility of 51 cm2/V∙s, carrier concentration of 1.5 × 1019 cm−3, with an average optical transmittance of 85% in the wavelength (400 - 1200) nm region and with a value of the band gap of 2.48 eV. The deposited CdO layer encloses the surface of the CdS layer, smoothers the surface relief and helps block impurities diffusion from CdS in the CdTe layer [27] . Using the optimized growth parameter determined for the individual layers, two different heterostructures (HJs) were prepared: TCO/CdS/CdTe and TCO/CdS/CdO/CdTe. Both devices were subjected to postdeposition chemical CdCl2 treatment in a saturated solution of CdCl2 in methanol on the CdTe surface, followed by thermal annealing in air for 30 min at 400˚C temperature. For chemical treatment the CdCl2 powder 99.999% purity was used from Alfa Aesar, Germany. Then, with the aim of illustrating the application of these HJs, solar cell devices were fabricated with both structures by adding a Ni back contact evaporated onto the CdTe surface [26] . Before the deposition of the absorbing layer, a phlogopite mica mask was applied to cover the half of the CdS surface for the deposition of CdO before the deposition of the CdTe (Figure 1, left part).

The structure grown in this way, made it possible to obtain two types of structures in a single cycle and measure them separately, despite the fact that all stages of the fabrication were identical.

The surface morphology of the films and chemical composition were investigated using the scanning electron microscope (SEM VEGA TESCAN TS 5130 MM) coupled with energy-dispersive X-ray spectroscope (EDX, INCA OXFORD instruments). The acceleration voltage was 15 kV. AFM studies were performed using the SmartScanTM program with the XE-7 AFM System (Park Systems Corp., Korea) under room temperature. The surface topography images of samples were taken with a piezo scanner and recorded for 5 μm × 5 μm and 10 μm × 10 μm areas. The scanning process was conducted in the tapping mode with a 1 Hz scanning rate. A parabolic diamond cantilever with a radius of curvature approximately 10 nm, a normal force constant of 40 N/m was used for fitting. All analyses, including topography and surface roughness were performed using the XEI-7 Image Processing Program. The device performance was tested with conventional dark/light current-voltage (J-V) characteristics using a standard light source equivalent to AM1.5 (100 mW/cm2), 300 K. The temperature dependent J-V and capacitance-voltage (C-V) characteristics of TCO/CdS/CdTe and TCO/ CdS/CdO/CdTe photovoltaic devices were measured with a KEITHLEY 2400 Source Meter, at frequency of 1 MHz, a.c. amplitude of 30 mV and DC bias scanning from −1.0 V to 1.0 V. The admittance spectroscopy measurements

Figure 1. The cross section of a CdS/CdTe photovoltaic device with and without interfacial CdO nanolayer.

(TAS) were performed at temperature varied from 20 K to 320 K with temperature step of T = 10 K, using a Wayne Kerr 6500B impedance analyzer. The electrical impedance Z and phase angle of the devices were measured as a function of same temperature interval, frequency—varied from 20 Hz to 10 MHz and AC voltage kept as low as 30 mV, to reduce the background noise. Dark measurements were performed at a DC bias of 0 V. The density of the deep states in CdTe absorber and at the related interface was determined from the demarcation energy concept, as a derivative of the capacitance, according to the model described in [28] .

3. Results and Discussion

3.1. AFM and Morphology Studies

The literature investigation shows, that surface and interface in CdTe photovoltaic devices play an important role in the determining of the solar cell efficiency. The morphological characteristics of the CdS thin film were investigated by AFM and SEM. It has been revealed that the morphology of CdS films has a considerable impact on their overall properties. This was especially noted in the devices, wherein the surface roughness and grain boundaries affected the recombination in the films. The 3D AFM image topography was used to investigate CdS topography on different substrates. In Figure 2, 3D AFM images of CdS on glass, on CdTe thin film, on CdTe/CdO interface and CdO on CdS thin film are presented.

The surface of all samples reveals a distinct distribution of grains with a uniform granular topography. The surfaces consist of hundreds nano-sized grains when CdS deposited on CdTe and CdTe/CdO, and the root mean square roughness (Sq)

Figure 2. The AFM topography of CdS thin film surfaces on glass substrate (a), on CdTe thin film (b) and on CdTe/CdO interface (c). The AFM of CdO nanolayer surface deposited on CdS thin film (d).

of the deposited CdS thin films was 284 nm and 222.4 nm, respectively, which indicates the formation of smooth and well-connected grains on the film (Table 1). When CdS is deposited on glass and CdO on CdS, the values of Sq are 31.1 nm and 29.3 nm, respectively. A low influence on Sq was found for CdS and CdO samples in comparison with CdS deposited on CdTe and CdTe/CdO surfaces. The same situation is observed and for the average surface roughness (Sa). For all films, skewness (Ssk) shows negative values that specify more valleys than peaks, and the surface is more planer. We point out that, when CdS is deposited on glass and CdO on CdS, the films surfaces are more homogenous and flatter. This can be explained by an antireflection effect of the CdO layer. As the refractive index of CdO is about 2.5, of CdS is 2.4 and of CdTe about 3.0, there is better matching of the refractive index at the air/CdS/ CdO/CdTe interfaces (Sq = 29.3 nm) compared to the air/CdS/CdTe interface (Sq = 284.0 nm). Moreover, after CdO nanolayer deposition, additional interference fringes in the CdO appears in the CdTe absorption (300 - 1200 nm) region due to the flat CdO on CdS interfaces, which can impact the light intensity in the bulk of the absorber. Moreover, method for interfacial defect passivation could change the interface morphology and modify recombination rate, one would expect higher recombination for a rough interface (Sq = 284.0 nm) rather than a smooth interface (Sq = 29.3 nm).

We conclude that the surface morphology of the films is influenced by both, the nature of the substrate and CdO on CdS provoked a decreasing of the surface roughness.

The literature analysis suggests that an intrinsic interlayer formed at the interface suppress the surface recombination [29] . Figure 3 displays a plain view SEM image showing the morphology of CdO on CdS surface (a) and a cross-section of a CdS/CdO/CdTe heterostructure (b). After the deposition of CdO thin film of about 8 - 11 nm on the CdS, the surface became more homogeneous. The deposited CdO layer encloses the surface of the CdS layer, smoothers the surface relief and helps block impurities diffusion from CdS in the CdTe layer. The SEM cross-section image of CdS/CdO/CdTe heterostructure (Figure 3(b)) shows the big grains for CdTe that not densely packed.

Also, voids and a non-uniformity of the CdS and CdTe films thickness at the

Table 1. The AFM surface parameters of CdS on different substrates and CdO on CdS surface.

The root means square roughness—Sq, The average surface roughness—Sa, Skewness—Ssk, Bearing Mean diameter—D.

Figure 3. The SEM image of CdO deposited on CdS (a) thin film and cross-section image of TCO/CdS/CdO/CdTe structure (b). The insets in cross-section image of a CdS/CdO/CdTe heterostructure present the non-uniform accumulation of CdO at the interface.

interface are observed. The cross-section inset confirms the non-uniform accumulation of CdO at the interface of CdS/CdTe. The presence of the void’s at CdS internal surface (Figure 3(b)) indicates that the CdO penetrate within the voids and reduce the imperfections on its surface. The evolutions of the elemental distribution across the depth of CdS/CdTe and CdS/CdO/CdTe devices are shown in Figure 4.

The EDX elemental distribution profiles of both CdTe devices show the Cd-L, Te-L, S-K lines. The EDX signals of Cd (Lα1) and S (Kα1) at 3.1 eV and 2.3 eV were observed. In addition, CdS/CdO/CdTe devices show O-K line (yellow) at 0.5 eV. The results of EDX line scans of cross-section of the CdS/CdO/CdTe device show non-uniform distribution of Cd-L (blue) and Te-L (green) lines in the elemental profile compared to CdS/CdTe heterostructure, where distinct regions can be seen. The image of CdS/CdO/CdTe device reveals complex interfacial chemistry, with a bilayer in place of the CdS. CdO film on the CdS layer suppresses the S interdiffusion from CdS into CdTe film and prevents the formation of solid solution with a lower bandgap that results in poor quantum efficiency.

3.2. The Impact of Sputtered CdO Nanolayer on Device Performance

The J-V characteristics for the best SnO2/CdS/CdTe and SnO2/CdS/CdO/CdTe devices under 100 mW∙cm−2 illuminations, 300 K are shown in Figure 5(a).

The photovoltaic parameters, the fill factor (FF), short circuit current density (Jsc), open circuit voltage (Voc) and the efficiency of the conversion (η), estimated from these dependencies are presented in Table 2. Comparison of the photovoltaic parameters of solar cells based on CdTe revealed that solar cells with i-layer are the best. To further show the utility of i-CdO approach on short circuit current density, FF and open circuit voltage, a number of nine CdS/CdTe structures with different thicknesses of CdO insulating layer were compared. According to

Figure 4. EDX line scan of cross section of Cd, S, Te and O profile, CdS/CdTe, (left), CdS/CdO/CdTe (right).

Figure 5. The J-V characteristics of the CdS/CdTe and CdS/CdO/CdTe heterostructures under 100 mW/cm2, 300 K (a) and dependence of Jsc in function of the CdO insulating layer thickness (b).

Table 2. The photovoltaic parameters of CdS/CdTe and CdS/CdO/CdTe devices.

Figure 5(b), for a thicker i-layer than 11 nm, Jsc decreases because degrades the drift electric field for carrier transport.

Through the theoretical analysis it is shown that the insertion of a thin insulating layer between the two semiconductors of an n+-p structure improves the efficiency in the same way as for a metal-insulator semiconductor (MIS) solar cell, if the discontinuity in the conduction bands constitutes a barrier for excited electrons. In our devices, the CdO interfacial layer with thickness greater than 11 nm modifies the morphology and composition of CdS and either plays no role or is disadvantageous.

For the open circuit voltage (Figure 6(a)), a sharp increase is observed at 5 - 11 nm thicknesses of the CdO layer and at thickness greater than 11 nm, it remains constant at 0.79 V. From these dependencies, for p-i-n heterojunction devices it follows that the optimal thickness of i-CdO is between 8 nm and 11 nm. It can be assumed that when CdO is too thin, it acts as a tunneling layer for the current, while when it is thicker, it acts as a blocking layer. So, sputtered i-CdO nanolayer as an interface engineering approach led to an enhancement of both Jsc and Voc by 5% and 25%, respectively. The series (Rs) and shunt (Rsh) resistances in this paper were calculated from the line slope of the J-V curves at the open-circuit voltage and short-circuit current points, respectively. The values of the Rs for both CdS/CdTe and CdS/CdO/CdTe devices are equally with 8.8 × 102 Ω∙cm2 and 1.1 × 102 Ω∙cm2, respectively, while the shunt resistances reached the values of 0.54 × 103 Ω∙cm2 and 0.5 × 103 Ω∙cm2. The Rs value for CdS/CdO/CdTe device was decreased. The Rsh value for both devices has about the same value and smaller than of the series resistance because of the non-idealities, mainly impurities near the p-n heterojunction, which cause partial shorting of the heterojunction.

Low shunt resistance causes power losses in the photovoltaic devices by providing an alternate current path for the light-generated current. Such a diversion reduces the amount of current flowing through the photovoltaic device junction and reduces the voltage. The low values of shunt resistance usually result from manufacturing defects.

Analysis of the dark current-voltage characteristics (Figure 6(b)) show that the introduction of i-CdO nanolayer between the CdS and CdTe layers leads to a shift in the I-V curves but not of the changing of the current transport mechanism. The built-in voltage increases from 0.51 V to 0.65 V but the reverse saturation

Figure 6. Dependencies of Voc, FF parameters of CdS/CdO/CdTe devices in function of the CdO insulating layer thickness (a) and the dark I-V characteristics of CdS/CdTe and CdS/CdO/CdTe heterojunctions at room temperature (b).

current decreases from 7.5 × 10−10 A to 1.5 × 10−10 A. The increase in the built-in voltage of the devices with i-CdO nanolayer is due to the change of the work function of the n-doped electron transport layer. Finally, we can conclude that the same model current mechanism can be applied for both CdS/CdTe and CdS/CdO/CdTe heterojunctions solar cells as the diode quality factor is n = 1.9 and n = 1.8, respectively. The decrease of n by 0.1 for CdS/CdO/CdTe solar cell is indication of the enhancement of open circuit voltage and current density values, with 0.11 V and 1.9 mA/cm2, respectively. It is therefore suggested that the recombination mechanism of CdS/CdTe photovoltaic cells is primarily SRH mechanism, achieved by deep trap states in space charge region of CdTe. The energy level can provide both electrons and holes. When n varies from 1 to 2, defects play the role of deep trap level. When the change of n is getting closer to 1, the recombination mechanism begins to fall in inter-band recombination, mostly in the neutral region. Surface recombination will make the change of n value more than 2. The recombination at grain boundaries should also be considered, causing the loss of carrier collection and reducing the voltage value. At the CdS/CdTe interface, there is a 9.7% lattice mismatch between the cubic CdS and the cubic CdTe. Such a high lattice mismatch causes considerable defects at the CdS/CdTe interface. Therefore, both devices exhibit SRH recombination as the dominant current transfer mechanism across the interface because the large conduction (ΔEC = (0.23 … 0.3) eV and valence (ΔEC = (0.67 and 0.74) eV bands offsets block thermal currents [30] . ΔEC is a vital factor affecting carrier transport and recombination.

To identify the nature of traps from J-V-T measurements, we briefly present analysis of the dark J-V-T characteristics. Figure 7(a), Figure 7(b) presents the Arrhenius plots of the saturation current density (Js) of both devices, measured in the temperature range of 173 K - 331 K. The temperature dependencies revealed in Figure 7(a), Figure 7(b) by the slopes of Arrhenius plots provide useful insight into the presence of two defect levels for CdS/CdTe and one defect

Figure 7. The saturation current density dependency on 1000/T for CdS/CdTe (a) and for CdS/CdO/CdTe devices (b).

level for CdS/CdO/CdTe devices. By analysis of Js evolution with reciprocal temperature, we distinguish regions where Js changes were fitted by linear function. In both CdS/CdTe and CdS/CdO/CdTe devices, in the lower temperature region of Js, the defect level with activation energies of 0.24 eV and 0.25 eV was revealed, while in the higher Js temperature region of CdS/CdTe, device a defect level with activation energies of 0.45 eV is depicted.

The saturation current density reveals that their changes with temperature in CdS/CdTe device follow the rule of recombination as a dominant carrier transport mechanism, via the defect levels located at 0.45 eV and 0.24 eV above the valence band edge of CdTe, while in the CdS/CdO/CdTe device only via the defect level located at 0.25 eV. Therefore, to gain better understanding into the carrier density and deep defects in the CdTe absorber, as well as, the effect of possible defects at the heterojunction interface for the devices with and without CdO interface nanolayer, CV and TAS measurements were conducted. The CV and TAS results were compared only for the reference CdTe/CdS device (i.e., without CdO interface layer) and for the best cell, employing an 8 - 11 nm thick CdO interface nanolayer.

3.3. Carrier Concentration and Defect Density in CdTe Absorber

There are three main interfaces for CdTe thin-film photovoltaic devices: TCO/CdS interface, CdS/CdTe interface, and CdTe/Ni interface. As the main p-n junction of CdTe thin-film photovoltaic devices, the CdS/CdTe interface depletion region is the key in photoelectric conversion of CdTe thin film photovoltaics, and is the core of the entire device. To understand the Jsc and Voc CdS/CdTe photovoltaic devices improvements by engineering of very thin insulating CdO wide-bandgap semiconducting material presented in previous paragraph, J-V-T and C-V measurements at different temperatures has been applied to both CdS/CdTe and CdS/CdO/CdTe devices. Also, admittance measurements C-ω-T give information on the widths of the space charge region of the device and on the contribution of trap levels to the capacitance of the device. High-frequency and low-temperature value of capacitance provide the total width of the depletion region, while the low frequency—the depletion width diminished by the response of traps, which are fast enough to respond to the AC signal of the admittance meter [26] . From bulk capacitance we extract the depletion layer width and the apparent free carrier density according to the equations: [30]

w = ε A C , (1)

N ( W ) = 2 q ε A 3 [ d ( 1 / C 2 ) d V ] 1 (2)

where ε = εoεr, εr is the dielectric constant of CdTe, εo is the permittivity of free space, q is the electron charge, W is width of the depletion region, A is the area of the device. We assume a one-sided device, in which the dominant capacitance response originates from the edge of the depletion region in the more intrinsic layer, and assumes that the temperature is high enough that the carriers can be rapidly transported into and away from the edge of the depletion region at the applied AC frequency. The application of DC bias (V) changes the value of the depletion width in CdS/CdO/CdTe device because at the interface appear a material with the relative permittivity ε different from of CdS and with a higher NCV density concentration. Figure 8(b) shows 1/C2 = f(V) for the CdS/CdO/CdTe structures at 100 kHz and 1 MHz frequencies. Evidently, from this figure, the (C2-V) relation is generally nonlinear. However, each curve can be fitted approximately with a broken line consisting of one linear region. The slopes of the straight lines increase with frequency. The interception of the lower voltage straight lines with the voltage axis results in different values for the built-in voltage. As the NCV density concentration is not constant throughout the depletion layer, the different values for the built-in voltage were obtained. The results from the experimental measurements listed in Table 3 show that at 0.1 MHz frequency the main charge carriers transport processes take place in the depletion region of (5.2 - 5.3) μm width for CdTe/CdS thin film devices. The CdO nanolayer passivates the shallow defects resulting in a higher NCV density concentration at both low and higher frequencies and the carriers transport processes take

Table 3. Electrical parameters of of CdTe/CdS thin film photovoltaic device, without and with a CdO layer at the interface between CdTe and CdS.

Figure 8. Depth carrier density profile versus depletion width of absorber films in CdTe/CdS and CdTe/CdO/CdS devices, calculated from capacitance-voltage (C-V) measurements (vertical dashed lines indicating 0 V) (a) and 1/C2-V dependence for/CdO/CdS best device measured at 100 KHz and 1 MHz, at 300 K (b).

place in the narrower depletion region of 4.7 μm width. If the depletion region width, at 1 MHZ frequency, for one of the CdTe/CdS thin film device slightly increases but for another decrease, then for the device with i-CdO nanolayer for both lower and higher frequencies, the depletion region width decreases. A more interesting feature is with built-in voltages at 1 MHz, for both types of devices, the Vbi values are increased. An increase of about 0.5 V was observed in the case of the device with i-CdO nanolayer. It is well known that in a semiconductor, Vbi equals the potential difference across the depletion region in thermal equilibrium. In thermal equilibrium, no external voltage is applied to the device. Moreover, the Fermi energy is constant throughout the heterostructure and the built-in voltage equals the difference between the Fermi energies, respectively in the buffer and the absorber, divided by the elementary charge. Figure 8(a) presents the defect profiles that show the typical U-shaped form commonly reported in literature [31] [32] , with a minimum at a profiling distance of 2.0 - 5.5 µm. Both profiles exhibit a significant increase in the defect profile toward the junction and toward the bulk of the sample. Trying to understand differences between the low frequency and high frequency profiles, it can be seen that the curves have an almost identical shape, but seem to be shifted by a distance.

An apparent depth dependence of the resulting doping concentration is often attributed to charge contributions of deep trap levels because their charge state depends on the band bending. The minimum NCV densities concentration is in the range of ~9 × 1013 cm−3 to ~1 × 1014 cm−3 and increases toward negative bias voltages, that is, toward higher apparent depth within the CdTe. The NCV densities were determined from the bottom of the U-shaped curve at zero bias. For the CdS/CdO/CdTe devices there is an observable increase in the NCV density profile inside the device and a decrease of a value of the width in comparison with non-engineering of the interface devices. Also, when the depletion region enters the front and the back-contact regions, the NCV concentration there rapidly rises. Another observation is a difference between the built-in voltages (Vbi). From one side, this suggests that performance degradation of CdS/CdTe without not-engineering interface can arise from lower NCV density due to CdCl2 treatment.

According to the previous investigation [33] , CdCl2 activation treatment is responsible for the formation of a VCd shallow acceptor complex, which leads to an increase in the doping densities in CdTe devices.

However, the excess of Cl can lead to the formation of self-compensating ClTe donors [34] [35] . This effect could explain the decrease in the NCV density concentration which can be directly translated into the VOC decreases, as presented previously. Through the deposition of i-CdO nanolayer when the device is reverse biased the majority charge carriers are stopped to push away from the junction and less ions are left at the junction. This beneficial situation reduced the available VCd (shallow acceptor centers) and the doping density of the CdS increases, which lead to an increase in VOC. Since C-V technique gives information about the kinetics of carrier transport, then with AS, the characteristics of the majority carrier trapping defects can be determined for solar cell materials and devices. If deeper carrier traps are present, the band bending in the SCR causes the Fermi level EF to cross the trap level Et at some distance from the interface, at the crossing point xt shown in Figure 9.

An applied small AC voltage with the frequency f = ω/2π causes the electric charge accumulated by traps to oscillate in the vicinity of crossing point xt. In the case of low frequency, the trap related capacitance is Ct = Clf, where Clf is the low frequency capacitance, while at high frequency measurements give the junction capacitance, i.e. depletion area capacitance Cd = Chf, where Chf is the high frequency capacitance. Accordingly, in the case of a single majority carrier trap level, the total junction capacitance can be described by the equation:

C d + C l f C d 1 + ω 2 τ 2 (3)

where τ is the characteristic trapping time that depends of the trap density Nt [36] . Figure 10 presents the C-f-T curves measured in the temperature range from 50 K to 320 K. The values of the capacitances for the device without i-CdO nanolayer are slightly smaller and temperature dependent in comparison with of the device with nanolayer at the interface.

The inflection frequency ωo can be obtained from the analysis of the first derivative of the capacitance, dC/dω, which demonstrates a maximum at the frequency ωo. A small AC oscillating voltage with the frequency 0.2 MHz causes the electric charge accumulated by traps to oscillate in the vicinity of this crossing point. The trapped electric charge follows the applied voltage oscillations and contributes to the total capacitance only if their frequency does not exceed the trap characteristic frequency ωo. In the case of a small trap concentration Nτ, the characteristic frequency ωo for the charging and discharging of defect levels is ωo = 1/τ [37] . The temperature dependence of the inflection frequency ωo is determined by the following equation [37] :

ω o ( T ) = 2 e t ( T ) = 2 N C , V v t h σ n , p exp ( E A k T ) = 2 ξ 0 T 2 exp ( E A k T ) (4)

Figure 9. Energy band diagram of solar cell for reverse bias conditions (at τ = ∞) consisting of a CdTe absorber with interface states and deep levels. The symbols represent their usual meanings.

Figure 10. C-T-f curves of CdTe/CdS thin film solar cell, without (a) and with a CdO layer (b) at the interface between CdTe and CdS.

where et is the emission rate, NC,V is the effective density of states in the conduction and valence band, νth is the thermal velocity of the minority carriers at the interface, σn,p is the capture cross-section for electrons and holes, EA = Et E is the activation energy of the defect level Et with respect to the band edge E, k is the Boltzmann constant, and ξ0 covers all the temperature independent parameters [38] . The activation energy of a defect level EA was obtained from the temperature frequency dependence of the capacitance spectra i.e., from the Arrhenius plot of the quantity ln(ωo/T2) versus 103/T. From the Arrhenius plot of the ln(ωo/T2) versus 103/T presented in Figure 11 were estimated the activation energy of the defect levels. According to the TAS theory [38] , the majority carrier to the frequency of the AC signal is limited by the modified dielectric relaxation frequency (ωo = W/d∙εr∙ρ), where ρ and εr are the material resistivity and relative permittivity, respectively, W is the depletion width, and d is the absorber thickness. For ω < ωo, the CdTe in the quasineutral region behaves as a conductor with a bulk conductance, while for ω > ωo, it behaves like a capacitor with a bulk capacitance. The extracted activation energies for CdS/CdTe are 0.43 eV and 0.20 eV.

Also, using admittance spectroscopy technique, authors of [39] [40] [41] [42] found a thermal ionization energy of 0.23(3) eV in CdTe attributed to the VCd defects, which is in good agreement with first principles investigations. In CdTe, Cd vacancies act as acceptors with two ionization energy levels, the first ionization state V C d ( 0 / 1 ) and the second ionization state V C d ( 1 / 2 ) have been reported [43] [44] . The activation energy of 0.20 eV defect level could be located above the valence band maximum for both heterojunction devices. According to the theory, wide-band gap semiconductors show self-compensation, that is the spontaneous formation of acceptor-type intrinsic defects to compensate or passivate donor-type.

The CdCl2 activation step is standard and vital process in fabricating high

Figure 11. The Arrhenius plot of the ln(ωo/T2) versus 1000/T of CdS/CdTe thin film solar cells, without (a) and with a CdO layer (b) at the interface.

efficiency PV devices. From a theoretical point of view, chlorine is predicted to promote further p-type doping of CdTe via a shallow donor ClTe defect with an ionization energy below the conduction band.

So, defect energies extracted from the admittance spectroscopy for CdS/CdTe devices (0.43 eV and 0.20 eV), CdS/CdO/CdTe (0.22 eV) and J-V-T (0.45 eV and 0.24 eV), (0.25 eV) measurements, respectively are in good agreement. The experimentally measured activation energies values from plot of capacitance transients and J-V-T techniques are in good agreement and therefore, we assume that the activation energy of 0.43 eV occurs due to the formation of cadmium vacancies VCd and acceptor complexes VCd – Cl, which continuously shift Fermi level towards valence band and decrease resistivity. Several publications [45] [46] [47] have been reported a deep trap in the range 0.43 - 0.54 eV above the valence band maximum [48] [49] [50] [51] . These experimental values are in good agreement with our value highlighted by TAS measurements in CdS/CdTe device without interface engineering and belonging to the complex (ClTeVCd)−1 that consists of a C l T e + 1 impurity and a V C d 2 defect that are below to acceptor (ClTe VCd)−1 complex (AX-center).

Taking into account the activation energy values of 0.43 eV and 0.20 eV presented in Figure 11 and assuming the energy diagram of CdTe for the CdS/CdTe device presented in Figure 12, the energetic position of these two defects 0.43 eV (AX-center) and 0.22 eV (acceptor-like defect) are situated in the space charge region. The analysis of the current-voltage characteristics in the higher and lower temperature regions of CdS/CdTe device indicates the presence of the traps with activation energy of 0.45 eV and 0.24 eV, respectively and is assigned to the recombination process as a dominant current flow mechanism. For the device with CdO insulating layer at the interface, the only a trap with activation energy of 0.25 eV was revealed. The identified traps act as a recombination centers for electrons. The TAS investigation revealed the same deep levels in CdS/ CdTe and CdS/CdO/CdTe photovoltaic devices, among which the most dominant

Figure 12. Schematic of the assumed energy diagram of CdTe layer in the CdS/CdTe heterojunction device.

was a trap with an activation energy of 0.20 eV and 0.22 eV. This defect has about the same value as energy states determined by J-V-T analysis. Based on the literature review and TAS system database, we assigned it to the formation of cadmium vacancies VCd. This obviously confirms that VCd defects act as a recombination center and, therefore, have a significant impact on the current flow mechanism in the low temperature region.

The engineering of interface of the CdS/CdTe heterostructure with CdO thicknesses about 8 - 11 nm “kills” the deep (ClTeVCd)1 AX-center and improve both Jsc and Voc parameters, but keeps the defect with the activation energy values of 0.20 eV and 0.22 eV in both CdS/CdTe and CdS/CdO/CdTe devices.

In summary, we can assume that the concentration of free charge carriers is determined by the vacancies of cadmium VCd defects with activation energies of 0.20 eV (CdS/CdTe) and 0.22 eV (CdS/CdO/CdTe) devices that agree well with values of the publications [51] [52] [53] . The typical dopant concentrations are ~1014 to 1015 cm−3 for polycrystalline CdTe which is limited by the number of VCd present in CdTe [54] . The CdS/CdTe interface is one of the critical device regions in CdTe photovoltaic devices and its properties are not clearly understood at this time, and influence on device stability still unknown.

4. Conclusions

In this paper, we demonstrate that engineered nanometer CdO thin films grown by DC magnetron sputtering is an effective interface-passivation strategy for improving CdTe photovoltaic device performance. The CdO film deposition rate was 5 Ǻ/s. The TCO/CdS/CdTe/Ni and TCO/CdS/CdO/CdTe/Ni photovoltaic devices were fabricated by CSS method under optimized fabrication conditions.

The photovoltaic parameters, such as fill factor, the short circuit current density, the open circuit voltage and the energy conversion efficiency estimated from J-V dependencies under illuminations of 100 mW/cm2 for the n-CdS/p-CdTe heterostructure are 0.44, 24.9 mA/cm2, 0.69 V and 7.5%. A CdS/CdTe photovoltaic device with insulating CdO interfacial layer showed power conversion efficiency of 11.6% with JSC, VOC, and FF of 26.8 mA/cm2 0.80 V and 0.55, respectively. The enhancement of both Jsc and Voc by 5% and 25%, respectively is observed.

Analysis of the C-V-T measurements shows an observable increase in the NCV density profile inside the CdS/CdO/CdTe device, a difference between the built-in voltages from 0.83 V to 1.02 V and a decrease of a value of the width from 5.6 µm to 4.7 µm compared to device with non-engineered interface. In addition, after interface engineering via sputtering of CdO nanolayer the removal of the fast recombination deep (ClTeVCd)1 AX-center forming during CdCl2 thermal treatment had place.

Further improvement of the efficiency in CdS/CdO/CdTe devices requires the elimination of the defect with the activation energy around 0.20 eV. For further improvement of CdS/CdTe solar cells efficiency we propose at the back contact to use the material with the high work function, such as graphene, that can also reduce interface carrier recombination. Band alignment measurements using photoelectron spectroscopy and synchrotron techniques will be included in the investigation of CdS/CdO/CdTe devices with contact of graphene to CdTe layer.

Acknowledgements

This work was supported by the Ministry of Education, Culture, and Research of Republic of Moldova, research grant 20.80009.5007.16.

Conflicts of Interest

The authors declare no conflicts of interest regarding the publication of this paper.

References

[1] Zghaibeh, M., Okonkwo, P.C., et al. (2022) CdTe Solar Cells Fabrication and Examination Techniques: A Focused Review. International Journal of Green Energy, 20, 555-570.
https://doi.org/10.1080/15435075.2022.2126943
[2] Romeo, A. and Artegiani, E. (2021) CdTe-Based Thin Film Solar Cells: Past, Present and Future. Energies, 14, Article 1684.
https://doi.org/10.3390/en14061684
[3] Bonnet, D. (1992) The CdTe Thin Film Solar Cell—An Overview. International Journal of Solar Energy, 12, 914-924.
https://doi.org/10.1080/01425919208909746
[4] Britt, J. and Ferekides, C. (1993) Thin-Film CdS/CdTe Solar Cell with 15.8% Efficiency. Applied Physics Letters, 62, Article ID: 2851.
https://doi.org/10.1063/1.109629
[5] Wu, X. (2004) High-Efficiency Polycrystalline CdTe Thin-Film Solar Cells. Solar Energy, 77, 803-814.
https://doi.org/10.1016/j.solener.2004.06.006
[6] Wu, X., Dhere, R., Albin, D. and Gessert, T. (2001) Advances in CdTe R&D at NREL. Proceedings of the NCPV Program Review Meeting, Colorado, 14-17 October 2001, 47.
[7] Gloeckler, M., Sankin, I. and Zhao, Z. (2013) CdTe Solar Cells at the Threshold to 20% Efficiency. IEEE Journal of Photovoltaics, 3, 1389-1393.
https://doi.org/10.1109/JPHOTOV.2013.2278661
[8] Geisthardt, R., Topic, M. and Sites, J. (2015) Status and Potential of CdTe Solar-Cell Efficiency. IEEE Journal of Photovoltaics, 5, 1217-1221.
https://doi.org/10.1109/JPHOTOV.2015.2434594
[9] Artegiani, E., Leoncini, M., Barbato, M., Meneghini, M., Meneghesso, G., Cavallini, M. and Romeo, A. (2019) Analysis of Magnesium Zinc Oxide Layers for High Efficiency CdTe Devices. Thin Solid Films, 672, 22-25.
https://doi.org/10.1016/j.tsf.2019.01.004
[10] Song, T., Kanevce, A. and Sites, J. (2016) Emitter/Absorber Interface of CdTe Solar Cells. Journal of Applied Physics, 119, Article ID: 233104.
https://doi.org/10.1063/1.4953820
[11] Burst, J. (2016) CdTe Solar Cells with Open-Circuit Voltage Breaking the1V Barrier. Nature Energy, 1, Article No. 16015.
https://doi.org/10.1038/nenergy.2016.15
[12] Wilhelm, H., Schock, H. and Scheer, R. (2011) Interface Recombination in Heterojunction Solar Cells: Influence of Buffer Layer Thickness. Journal of Applied Physics, 109, Article ID: 084514.
https://doi.org/10.1063/1.3554409
[13] Sites, J. and Pan, J. (2006) Strategies to Increase CdTe Solar Cell Voltage. Thin Solid Films, 515, Article ID: 6099.
https://doi.org/10.1016/j.tsf.2006.12.147
[14] Munshi, A., Kephart, J., Abbas, A., Shimpi, T., Barth, K., Walls, J., et al. (2018) Polycrystalline CdTe Photovoltaics with Efficiency over 18% through Improved Absorber Passivation and Current Collection. Solar Energy Materials and Solar Cells, 176, 9-18.
https://doi.org/10.1016/j.solmat.2017.11.031
[15] Colegrove, E., Yang, J., Harvey, S., Young, M., Burst, J., Duenow, J., Albin, D., Wei, S. and Metzger, W. (2018) Experimental and Theoretical Comparison of Sb, As, and P Diffusion Mechanisms and Doping in CdTe. Journal of Physics D: Applied Physics, 51, Article ID: 075102.
https://doi.org/10.1088/1361-6463/aaa67e
[16] Kephart, J., McCamy, J., Ma, Z., Ganjoo, A., Alamgir, F. and Sampath, W. (2016) Band Alignment of Front Contact Layers for High-Efficiency CdTe Solar Cells. Solar Energy Materials and Solar Cells, 157, 266-275.
https://doi.org/10.1016/j.solmat.2016.05.050
[17] Kim, S., Suh, J., Kim, T., Hong, J. and Cho, S. (2019) Plasmon-Enhanced Performance of CdS/CdTe Solar Cells Using Au Nanoparticles. Optics Express, 27, 22017-22024. https://doi.org/10.1364/OE.27.022017
[18] Kosyachenko, L.A., et al. (2013) Optical Absorptivity and Recombination Losses: The limitations Imposed by the Thickness of Absorber Layer in CdS/CdTe Solar Cells. Solar Energy Materials and Solar Cells, 114, 179-185.
https://doi.org/10.1016/j.solmat.2013.03.003
[19] Yang, J., Yin, W., Park, J., Ma, J. and Wei, S. (2016) Review on First-Principles Study of Defect Properties of CdTe as a Solar Cell Absorber. Semiconductor Science and Technology, 31, Article ID: 083002.
https://doi.org/10.1088/0268-1242/31/8/083002
[20] Kavanagh, S., Walsh, A. and Scanlon, D. (2021) Rapid Recombination by Cadmium Vacancies in CdTe. ACS Energy Letters, 6, 1392-1398.
https://doi.org/10.1021/acsenergylett.1c00380
[21] Li, C., et al. (2014) S-Te Interdiffusion within Grains and Grain Boundaries in CdTe Solar Cells. IEEE Journal of Photovoltaics, 4, 1636-1643.
https://doi.org/10.1109/JPHOTOV.2014.2351622
[22] Taylor, A., Major, J.D., Kartopu, G., Lamb, D., Duenow, J., Dhere, R.G., et al. (2015) A Comparative Study of Microstructural Stability and Sulphur Diffusion in CdS/CdTe Photovoltaic Devices. Solar Energy Materials and Solar Cells, 141, 341-349.
https://doi.org/10.1016/j.solmat.2015.06.010
[23] Potlog, T. (2016) Thin-Film Photovoltaic Devices Based on A2B6 Compounds. In: Tiginyanu, I., Topala, P. and Ursaki, V., Eds., Nanostructures and Thin Films for Multifunctional Applications, NanoScience and Technology, Springer, Cham, 143-186.
https://doi.org/10.1007/978-3-319-30198-3_5
[24] Kabir, M., Shahahmadi, S., Lim, V., Zaidi, S., Sopian, K. and Amin, N. (2012) Amorphous Silicon Single-Junction Thin-Film Solar Cell Exceeding 10% Efficiency by Design Optimization. International Journal of Photoenergy, 5, Article ID: 460919.
https://doi.org/10.1155/2012/460919
[25] Hossain, M.M., Ul Karim, M.M., Banik, S., Jahan, N.A. and Matin, M.A. (2016) Design of a High Efficiency Ultrathin CdTe/CdS p-i-n Solar Cell with Optimized Thickness and Doping Density of Different Layers. Proceedings of 2016 International Conference on Advances in Electrical, Electronic and Systems Engineering (ICAEES), Putrajaya, 14-16 November 2016, 305-308.
https://doi.org/10.1109/ICAEES.2016.7888058
[26] Gurumurugan, K., Mangalaraj, D., Narayandass, S. and Nakanishi, Y. (1996) DC Reactive Magnetron Sputtered CdO Thin Films. Materials Letters, 28, 307-312.
https://doi.org/10.1016/0167-577X(96)00074-2
[27] Al Qassem, A., Fedorov, V., Suman, V., Gagara, L. and Potlog, T. (2019) Improved Photovoltaic Parameters in CdTe Solar Cells by Insertion of a i-CdO Layer. Proceedings of 36th EU PVSEC 2019 Conference, Marseille, 9-13 September 2019, 673-677.
[28] Li, J.V., Johnston, S.W., Li, X., Albin, D.S., Gessert, T.A. and Levi, D.H. (2010) Discussion of Some “Trap Signatures” Observed by Admittance Spectroscopy in CdTe Thin-Film Solar Cells. Journal of Applied Physics, 108, Article ID: 06450.
https://doi.org/10.1063/1.3475373
[29] Yun, J.H., et al. (2014) Performance Improvement in CdTe Solar Cells by Modifying the CdS/CdTe Interface with a Cd Treatment. Current Applied Physics, 14, 630-635.
https://doi.org/10.1016/j.cap.2013.11.036
[30] Li, D.B., Bista, S.S., Awni, R.A., et al. (2022) 20%-Efficient Polycrystalline Cd(Se,Te) Thin-Film Solar Cells with Compositional Gradient Near the Front Junction. Nature Communications, 13, Article No. 7849.
https://doi.org/10.1038/s41467-022-35442-8
[31] Schroder, D. (2006) Semiconductor Material and Device Characterization. 3rd Edition, Wiley, Hoboken, 63.
https://doi.org/10.1002/0471749095
[32] Castaldini, A., Cavallini, A., Fraboni, B., Fernandez, P. and Piqueras, J. (1998) Deep Energy Levels in CdTe and CdZnTe. Journal of Applied Physics, 83, Article ID: 2121. https://doi.org/10.1063/1.366946
[33] Ochoa-Landín, R., Vigil-Galan, O., Vorobiev, Y.V. and Ramírez-Bon, R. (2009) Chemically-Deposited Te Layers Improving the Parameters of Back Contacts for CdTe Solar Cells. Solar Energy, 83, 134-138.
https://doi.org/10.1016/j.solener.2008.07.005
[34] Raadik, T., Krustok, J., Josepson, R., Hiie, J., Potlog, T. and Spalatu, N. (2013) Temperature Dependent Electroreflectance Study of CdTe Solar Cells. Thin Solid Films, 535, 279-282.
https://doi.org/10.1016/j.tsf.2012.12.083
[35] Bastola, E., Bordovalos, A., LeBlanc, E., Shrestha, N., Reese, M. and Ellingson, R. (2019) Doping of CdTe Using CuCl2 Solution for Highly Efficient Photovoltaic Devices. Proceedings of 2019 IEEE 46th Photovoltaic Specialists Conference (PVSC), Chicago, 16-21 June 2019, 1846-1850.
https://doi.org/10.1109/PVSC40753.2019.8980780
[36] Walter, T., Herberholz, R., Müller, C. and Schock, H.W. (1996) Determination of Defect Distributions from Admittance Measurements and Application to Cu(In,Ga) Se2 Based Heterojunctions. Journal of Applied Physics, 80, 4411-4420.
https://doi.org/10.1063/1.363401
[37] Gilmore, A., Kaydanov, V., Kaydanov, V. and Ohno, T. (2000) Mobility in SnO2:F Thin Polycrystalline Films: Grain Boundary Effect and Scattering in the Grain Bulk. MRS Online Proceedings Library, 666, Article No. 310.
https://doi.org/10.1557/PROC-666-F3.10
[38] Bayhan, H. and Kavasoglu, A. (2003) Admittance and Impedance Spectroscopy on Cu(In,Ga)Se2 Solar Cells. Turkish Journal of Physics, 27, 529-535.
[39] Gul, R., McCloy, J.S., Murugesan, M., Montag, B. and Singh, J. (2022) Cl-Doped CdTe Crystal Growth for Medical Imaging Applications. Crystals, 12, Article 1365.
https://doi.org/10.3390/cryst12101365
[40] Reislohner, U., Grillenberger, J. and Witthuhn, W. (1998) Band-Gap Level of the Cadmium Vacancy in CdTe. Journal of Crystal Growth, 184-185, 1160-1164.
[41] Potter, M.D.G., et al. (2000) A Study of the Effects of Varying Cadmium Chloride Treatment on the Luminescent Properties of CdTe/CdS Thin Film Solar Cells. Thin Solid Films, 361-362, 248-252.
https://doi.org/10.1016/S0040-6090(99)00782-8
[42] Hoschl, P., Grill, R., Franc, J., Moravec, P. and Belas, E. (1993) Cl-Doping of Te-Rich CdTe: Complex Formation, Self-Compensation and Self-Purification from First Principle. Materials Science and Engineering B, 16, 215.
[43] Popovych, V., Virt, I., Sizov, F., Tetyorkin, V., Tsybrii, Z., Darchuk, L., Parfenjuk, O. and Ilashchuk, M. (2007) The Effect of Chlorine Doping Concentration on the Quality of CdTe Single Crystals Grown by the Modified Physical Vapor Transport Method. Journal of Crystal Growth, 308, 63-70.
https://doi.org/10.1016/j.jcrysgro.2007.07.041
[44] Su-Huai, W. and Zhang, S. (2002) Chemical Trends of Defect Formation and Doping Limit in II-VI Semiconductors: The Case of CdTe. Physical Review B, 66, 155211.
https://doi.org/10.1103/PhysRevB.66.155211
[45] Herberholz, R., Igalson, M. and Schock, H. (1998) Distinction between Bulk and Interface States in CuInSe2/CdS/ZnO by Space Charge Spectroscopy. Journal of Applied Physics, 83, Article No. 318.
https://doi.org/10.1063/1.366686
[46] Major, J., Al Turkestani, M., Bowen, L., Brossard, M., Li, C., Lagoudakis, P. and Durose, K. (2016) A Low-Cost Non-Toxic Post-Growth Activation Step for CdTe Solar Cells. Nature Communications, 7, Article ID: 13231.
[47] Castaldini, A., Cavallini, A., Fraboni, B., Fernandez, P. and Piqueras, J. (1996) Comparison of Electrical and Luminescence Data for the A Center in CdTe. Applied Physics Letters, 69, 3510-3512.
https://doi.org/10.1063/1.117228
[48] Kremer, R. and Leigh, W. (1988) Deep Levels in CdTe. Journal of Crystal Growth, 86, 490-496.
https://doi.org/10.1016/0022-0248(90)90764-C
[49] Maghouli, M. and Eshghi, H. (2020) Studying the Effect of Deposition Time on Physical Properties of CdTe Thin Films; Influence of CdTe Electrical Properties on CdS/CdTe Heterojunction Rectifying Behavior. Optik, 218, Article ID: 165132.
https://doi.org/10.1016/j.ijleo.2020.165132
[50] Seymour, F.H., Kaydanov, V. and Ohno, T.R. (2005) Cu and CdCl2 Influence on Defects Detected in CdTe Solar Cells with Admittance Spectroscopy. Applied Physics Letters, 87, Article ID: 153507.
https://doi.org/10.1063/1.2099515
[51] Shepidchenko, A.S., Mirbt, B., Sanyal, A., Hakansson, A. and Klintenberg, M. (2013) Small Hole Polaron in CdTe: Cd-Vacancy Revisited. Journal of Physics: Condensed Matter, 25, Article ID: 415801.
[52] Lindstrom, A., Mirbt, S., Sanyal, B. and Klintenberg, M. (2016) High Resistivity in Undoped CdTe: Carrier Compensation of Te Antisites and Cd Vacancies. Journal of Physics D: Applied Physics, 49, Article ID: 035101.
https://doi.org/10.1088/0022-3727/49/3/035101
[53] Soundararajan, R., Lynn, S., Awadallah, C., Szeles, S. and Wei, H. (2006) Study of Defect Levels in CdTe Using Thermoelectric Effect Spectroscopy. Journal of Electronic Materials, 35, 1333-1340.
https://doi.org/10.1007/s11664-006-0264-0
[54] Becerril, M., Zelaya-Angel, O., Vargas-García, J.-R., Ramírez-Bon, R. and González-Hernández, J. (2001) Effects of Cd Vacancies on the Electrical Properties of Polycrystalline CdTe Sputtered Films. Journal of Physics and Chemistry of Solids, 62, 1081-1085.
https://doi.org/10.1016/S0022-3697(00)00284-5

Copyright © 2024 by authors and Scientific Research Publishing Inc.

Creative Commons License

This work and the related PDF file are licensed under a Creative Commons Attribution 4.0 International License.