Numerical Optimization of Gradient Bandgap Structure ... - MDPI

10
energies Article Numerical Optimization of Gradient Bandgap Structure for CIGS Solar Cell with ZnS Buffer Layer Using Technology Computer-Aided Design Simulation Joonghyun Park 1,2 and Myunghun Shin 3, * 1 School of Electronic Electrical Engineering, College of Information and Communication Engineering, Sungkyunkwan University, Suwon 440-746, Korea; [email protected] 2 Samsung Display Co., Ltd., 1, Samsung-ro, Yongin-si, Gyeonggi-do 17113, Korea 3 School of Electronics and Information Engineering, Korea Aerospace University, Goyang-city, Gyeonggi-do 412-791, Korea * Correspondence: [email protected] or [email protected]; Tel.: +82-2-300-0145 Received: 21 June 2018; Accepted: 5 July 2018; Published: 7 July 2018 Abstract: The band structure characteristics of a copper indium gallium sulfur selenide (Cu(In 1–x Ga x )SeS, CIGS) solar cell incorporating a cadmium-free zinc sulfide (ZnS) buffer layer were investigated using technology computer-aided design simulations. Considering the optical/electrical properties that depend on the Ga content, we numerically demonstrated that the front gradient bandgap enhanced the electron movement over the band-offset of the ZnS interface barrier, and the back gradient bandgap generated a back side field, improving electron transport in the CIGS layer; in addition, the short circuit current density (J SC ) and open circuit voltage (V OC ) improved. The simulation demonstrated that the conversion efficiency of a double graded bandgap cell is higher than with uniform or normal/reverse gradient cells, and V OC strongly correlated with the average bandgap in the space charge region (SCR) of CIGS. After selecting V OC from the SCR, we optimized the band structure of the CIGS cell with a Cd-free ZnS buffer by evaluating J SC and the fill factor. We demonstrated that the cell efficiency of the fabricated cell was more than 15%, which agrees well with the simulated results. Our numerical method can be used to design high-conversion efficiency CIGS cells with a gradient band structure and Cd-free buffer layer. Keywords: CIGS (copper indium gallium selenide); band structure; double graded bandgap; TCAD (technology computer-aided design) simulation; carrier transportation 1. Introduction Copper indium gallium sulfur selenide (Cu(In 1-x Ga x )SeS, CIGS) solar cells have been widely studied because of their various advantages. For thin film-based solar cells, CIGS solar cells can achieve an efficiency of over 20%, which is as high as that of high-quality crystalline silicon-based solar cells at small sizes (~0.25 cm 2 )[1]. CIGS cells fabricated on flexible substrates such as polyimide or titanium foils can be applied to wider areas than conventional Si solar cells [2]. In order to avoid toxicity to the environment, cadmium-free (Cd-free) zinc sulfide (ZnS) layers can be used as buffer layers for the CIGS solar cells [3]. The bandgap is the key feature of solar cells determining the optical and electrical characteristics of semiconductor devices, and in CIGS solar cells, the bandgap (E g ) can be engineered by controlling the gallium or sulfur content. Thus, one of the particular merits of CIGS solar cells is that the bandgap structure can be designed with a Ga or S gradient, and a CIGS solar Energies 2018, 11, 1785; doi:10.3390/en11071785 www.mdpi.com/journal/energies

Transcript of Numerical Optimization of Gradient Bandgap Structure ... - MDPI

energies

Article

Numerical Optimization of Gradient BandgapStructure for CIGS Solar Cell with ZnS Buffer LayerUsing Technology Computer-AidedDesign Simulation

Joonghyun Park 1,2 and Myunghun Shin 3,*1 School of Electronic Electrical Engineering, College of Information and Communication Engineering,

Sungkyunkwan University, Suwon 440-746, Korea; [email protected] Samsung Display Co., Ltd., 1, Samsung-ro, Yongin-si, Gyeonggi-do 17113, Korea3 School of Electronics and Information Engineering, Korea Aerospace University, Goyang-city,

Gyeonggi-do 412-791, Korea* Correspondence: [email protected] or [email protected]; Tel.: +82-2-300-0145

Received: 21 June 2018; Accepted: 5 July 2018; Published: 7 July 2018�����������������

Abstract: The band structure characteristics of a copper indium gallium sulfur selenide(Cu(In1–xGax)SeS, CIGS) solar cell incorporating a cadmium-free zinc sulfide (ZnS) buffer layer wereinvestigated using technology computer-aided design simulations. Considering the optical/electricalproperties that depend on the Ga content, we numerically demonstrated that the front gradientbandgap enhanced the electron movement over the band-offset of the ZnS interface barrier, and theback gradient bandgap generated a back side field, improving electron transport in the CIGSlayer; in addition, the short circuit current density (JSC) and open circuit voltage (VOC) improved.The simulation demonstrated that the conversion efficiency of a double graded bandgap cell is higherthan with uniform or normal/reverse gradient cells, and VOC strongly correlated with the averagebandgap in the space charge region (SCR) of CIGS. After selecting VOC from the SCR, we optimizedthe band structure of the CIGS cell with a Cd-free ZnS buffer by evaluating JSC and the fill factor.We demonstrated that the cell efficiency of the fabricated cell was more than 15%, which agrees wellwith the simulated results. Our numerical method can be used to design high-conversion efficiencyCIGS cells with a gradient band structure and Cd-free buffer layer.

Keywords: CIGS (copper indium gallium selenide); band structure; double graded bandgap; TCAD(technology computer-aided design) simulation; carrier transportation

1. Introduction

Copper indium gallium sulfur selenide (Cu(In1−xGax)SeS, CIGS) solar cells have been widelystudied because of their various advantages. For thin film-based solar cells, CIGS solar cells canachieve an efficiency of over 20%, which is as high as that of high-quality crystalline silicon-basedsolar cells at small sizes (~0.25 cm2) [1]. CIGS cells fabricated on flexible substrates such as polyimideor titanium foils can be applied to wider areas than conventional Si solar cells [2]. In order to avoidtoxicity to the environment, cadmium-free (Cd-free) zinc sulfide (ZnS) layers can be used as bufferlayers for the CIGS solar cells [3]. The bandgap is the key feature of solar cells determining the opticaland electrical characteristics of semiconductor devices, and in CIGS solar cells, the bandgap (Eg) canbe engineered by controlling the gallium or sulfur content. Thus, one of the particular merits of CIGSsolar cells is that the bandgap structure can be designed with a Ga or S gradient, and a CIGS solar

Energies 2018, 11, 1785; doi:10.3390/en11071785 www.mdpi.com/journal/energies

Energies 2018, 11, 1785 2 of 10

cell with an optimized bandgap profile can be fabricated through elemental substitution during thedeposition and annealing processes.

The Ga/(Ga + In) ratio, the parameter controlling the bandgap of CIGS with the Ga content,and its depth profile in the CIGS layer were studied to increase the efficiency, and the correlationbetween the CIGS bandgap structure and cell efficiency was analyzed using the numerical simulationsin several studies. Gloeckler and Sites, and Song et al. reported that the improvement in back gradingin the CIGS solar cell could potentially increase the efficiency when the thickness of the absorberdecreased [4–6]. In 2004 and 2010, Song et al. reported various bandgap profiles in the CIGS layer andsuggested an optimum bandgap structure, the double graded bandgap (DGB, bandgap gradings onboth front and back sides of the CIGS absorption layer) profile using the AMPS-1D and DESSIS 2Dsimulators [5,6]; the front grading improved the open circuit voltage (VOC) without reducing the shortcircuit current density (JSC) and the back grading improved the VOC. Dullweber et al. reported theeffect of back grading on the CIGS layer, which suppressed the carrier recombination and resulted inan enhancement in the VOC by up to 0.09 V [7].

In this work, we demonstrate a numerical procedure to design a gradient band structure for CIGSsolar cells with a Cd-free ZnS buffer layer and demonstrate bandgap-optimization to fabricate the CIGSsolar cells using a two-stage process. It is well known in CIGS solar cells the band alignment betweenabsorber layer and buffer layer is so important [8]. However, this paper focus on bandgap alignmentof absorber layer. Using a technology computer-aided design simulation (TCAD), we investigated theeffect of the gradient bandgap near the interface of the ZnS buffer layer on the conversion efficiencyparameters of the CIGS cells, which has not been fully proven yet. The conversion efficiency of theCIGS solar cells is also evaluated not only for the front gradient bandgap (FGB) and back gradientbandgap (BGB), but also for various other types of gradient bandgap profiles. In the simulation,in order to reduce the performance difference in the design and fabrication, we use the band structuralparameters obtained from the measurements and practical data from the literature, which matchwell with our device. This work will help us understand the engineering of the bandgap profileand management of the defect density related to the Ga/(Ga + In) ratio, which is key to achievinghigh efficiencies.

2. Experiment and Simulations

In this work, the CIGS solar cells were fabricated by a two-step process, the deposition of a metallicprecursor and annealing in a Se and S atmosphere. The general process is as follows. A 300-nm-thickMo layer is deposited as a back-contact metal by DC magnetron sputtering on a soda lime glasssubstrate. Cu, In, and Ga alloyed metals are deposited by a DC sputter as a precursor on the Molayer. The CIGS precursor is annealed in H2Se and H2S gas sequentially. By co-sputtering Cu-Ga, andin targets, we controlled the Ga profile in the precursor, and by adjusting the partial pressure of H2Seand H2S gases, we controlled Se, S contents in the CIGS cells. CIGS bandgap and profiling the Ga ratiowere analyzed using various material analysis methods, and reported in detail in other studies [9,10].Following these steps, the metallic precursor is reacted and diffused at high temperatures, and thedesigned gradient bandgap profile in the CIGS layer can only form with the proper precursor andthermal process conditions. After the annealing process, the CIGS absorption layer with a thicknessof about 1.8 µm is formed. An approximately 30-nm-thick ZnS buffer layer is deposited by chemicalbath deposition (CBD) on the CIGS layer. A 30-nm-thick undoped zinc oxide buffer layer (passivatingthe vacancy in ZnS layer) and 1.2-µm-thick boron-doped ZnO layer (front transparent conductingoxide, the front contact metal layer) are deposited sequentially by low-pressure chemical vapordeposition without any vacuum break. The device structure, and the scanning electron microscope(SEM) image of fabricated CIGS solar cell are shown in Figure 1a. Statistical approach is important forevaluating the performance of solar cells [11]. When the basic experimental conditions were satisfied,the characteristics and performance parameters of fabricated CIGS cells showed statistically very stablewith deviation of 10% or less. More details on the fabrication process are reported elsewhere [9,10].

Energies 2018, 11, 1785 3 of 10

The effects of the bandgap gradient on the CIGS solar cell conversion efficiency are investigatedusing a TCAD simulator (Synopsys Inc., Mountain View, USA). First, we characterized the simulationparameters such as thickness, carrier density, and material properties. Then, the thickness of eachlayer in the CIGS solar cell was measured by cross-sectional scanning electron microscopy and thecarrier density in the CIGS layer was estimated by the capacitive-voltage method. The elementalcomposition in the CIGS layer was measured using secondary ion mass spectroscopy to calculate theGa/(Ga + In) ratio as the gradient bandgap parameter. In order to apply the optical characteristicsto the simulation, we modeled the absorption coefficient as a function of the Ga/(Ga + In) ratio byadopting the results from [12]. Because the bandgaps of the CIGS layers can be varied from 1.04 eV to1.68 eV according to the Ga content, their optical parameters were also calculated with the bandgap,which was linearly interpolated. It is well known that the energy level, density, and shape of the defectstates are important factors determining the minority carrier lifetime; particularly, VOC is stronglydependent on the minority carrier lifetime. Hanna et al. reported that the defect states in the CIGSlayer consist of two types; the acceptor-like states of the Cu antisite and donor-like states of the Cuinterstitial or In antisite [13]. Of these, we primarily considered the Cu antisite states in the simulationbecause the acceptor-like state in the p-type semiconductor is a dominant factor. The defect densityof the CIGS layer as a function of Ga composition was also applied to the simulation. We referredto model B of the results of Frisk et al. [14]. We used a linear DGB structure of the CIGS layer witha ZnS buffer as [4–6], where the minimum bandgap of 1.1 eV is located at a depth of 0.25 µm in theCIGS from the interface of the ZnS buffer, and the maximum bandgap is 1.5 eV at a depth of 1.8 µmcontacting the Mo layer as shown in Figure 1b. From the reference structure, we investigated theeffects of the FGB and BGB on the CIGS solar cells.

Energies 2017, 10, x FOR PEER REVIEW 3 of 10

The effects of the bandgap gradient on the CIGS solar cell conversion efficiency are investigated

using a TCAD simulator (Synopsys Inc., Mountain View, USA). First, we characterized the

simulation parameters such as thickness, carrier density, and material properties. Then, the

thickness of each layer in the CIGS solar cell was measured by cross-sectional scanning electron

microscopy and the carrier density in the CIGS layer was estimated by the capacitive-voltage

method. The elemental composition in the CIGS layer was measured using secondary ion mass

spectroscopy to calculate the Ga/(Ga + In) ratio as the gradient bandgap parameter. In order to apply

the optical characteristics to the simulation, we modeled the absorption coefficient as a function of

the Ga/(Ga + In) ratio by adopting the results from [12]. Because the bandgaps of the CIGS layers can

be varied from 1.04 eV to 1.68 eV according to the Ga content, their optical parameters were also

calculated with the bandgap, which was linearly interpolated. It is well known that the energy level,

density, and shape of the defect states are important factors determining the minority carrier

lifetime; particularly, VOC is strongly dependent on the minority carrier lifetime. Hanna et al.

reported that the defect states in the CIGS layer consist of two types; the acceptor-like states of the

Cu antisite and donor-like states of the Cu interstitial or In antisite [13]. Of these, we primarily

considered the Cu antisite states in the simulation because the acceptor-like state in the p-type

semiconductor is a dominant factor. The defect density of the CIGS layer as a function of Ga

composition was also applied to the simulation. We referred to model B of the results of Frisk et al.

[14]. We used a linear DGB structure of the CIGS layer with a ZnS buffer as [4–6], where the

minimum bandgap of 1.1 eV is located at a depth of 0.25 μm in the CIGS from the interface of the

ZnS buffer, and the maximum bandgap is 1.5 eV at a depth of 1.8 μm contacting the Mo layer as

shown in Figure 1 (b). From the reference structure, we investigated the effects of the FGB and BGB

on the CIGS solar cells.

(a) (b)

Figure 1. Device structure: (a) scanning electron microscope (SEM) image of fabricated CIGS solar

cell, and (b) schematic band structure for TCAD simulation, where FGB, BGB, and SCR are front

gradient bandgap region, back gradient bandgap region, and space charge region, respectively.

3. Results and Discussion

In our CIGS solar cell, the Ga gradients were built on both the front and back of the CIGS layer.

In order to apply the Ga gradient effect to the device simulation, we used the results of Wei et al.; the

bandgap gradient effect of the CIGS layer by the variation of the Ga content mainly appears as an

increase in the conduction band derived from first-principles bandgap theory [13]. In the simulation,

the gradient bandgap profiles and electron affinities with various bandgaps were modeled

numerically using Equations (1) and (2):

Eg (x) = −0.375 x2 + 1.1075 x + 0.98 (1)

Figure 1. Device structure: (a) scanning electron microscope (SEM) image of fabricated CIGS solar cell,and (b) schematic band structure for TCAD simulation, where FGB, BGB, and SCR are front gradientbandgap region, back gradient bandgap region, and space charge region, respectively.

3. Results and Discussion

In our CIGS solar cell, the Ga gradients were built on both the front and back of the CIGS layer.In order to apply the Ga gradient effect to the device simulation, we used the results of Wei et al.;the bandgap gradient effect of the CIGS layer by the variation of the Ga content mainly appears as anincrease in the conduction band derived from first-principles bandgap theory [13]. In the simulation,the gradient bandgap profiles and electron affinities with various bandgaps were modeled numericallyusing Equations (1) and (2):

Eg (x) = −0.375 x2 + 1.1075 x + 0.98 (1)

Energies 2018, 11, 1785 4 of 10

Chi (x) = −0.475 x2 + 0.595 x + 4.48 (2)

where Eg (x) and Chi (x) are the bandgap and electron affinity, respectively, and x is the Ga/(Ga + In)ratio in the CIGS layer. The defect model based on [15] is empirical, but it fits well to our experimentalresults. The other parameters were selected in a reasonable range or referred from a previous paper [16].The main simulation parameters used in our simulations are listed in Table 1.

Table 1. Primary parameters of the materials in CIGS solar cell.

Parameters Unit Absorber p-CI1 − xGxS Buffer n-ZnS Window n-ZnO

Thickness nm 1800 30 1200Bandgap eV −0.375 x2 + 1.1075 x + 0.98 3.0 3.4

Electron affinity eV −0.475 x2 + 0.595 x + 4.48 4.7 4.5Carrier density /cm3 1.5 × 1016 1.5 × 1018 1.0 × 1018

Dielectric constant 13.6 9 9Electron/Hole mobility cm2/(V·s) 100/25 100/25 100/25

Defect state parameters

Density /cm3 Variable with Ga ratio 1.0 × 1018 1.0 × 1017

Peak level eV Mid-gap Mid-gap Mid-gapStandard deviation eV 0.1 0.1 0.1

Electron capture cross-section cm2 5.0 × 10−13 1.0 × 10−17 1.0 × 10−12

Hole capture cross-section cm2 1.0 × 10 −15 1.0 × 10 −12 1.0 × 10 −15

The electron and hole pairs generated by sunlight are separated and transferred to the oppositeside. The movement of the carriers is greatly affected by the bandgap profile in the solar cells.The electron and hole pairs can be moved easily in the space charge region (SCR) where a high internalelectric field is applied. Especially, the transportation of minority carriers in the CIGS absorptionlayer, and their movement over the band barrier at the interface of ZnS (due to electron affinity),are important causes for the conversion efficiency of the CIGS solar cells with ZnS buffer layers.Thus, the well-designed FGB of the CIGS layer could become the decisive factor determining theimprovement in carrier transportation in the SCR. In particular cases, the FGB could disturb thecarrier collection. Therefore, we investigated the conversion efficiency of the CIGS solar cell for theextent of the band gradient of the FGB structures. Figure 2 shows the conduction band and electronconcentration profiles in the SCR of the CIGS at the steady state when the bandgap energies at thefront edge are 1.1 eV, 1.18 eV, 1.24 eV, and 1.3 eV. The conduction band offset (CBO)—the mismatchedalignment in the conduction band between the buffer and CIGS absorber layers—was measured fromthe J–V distortion curves under red light and dark illuminations [17]. We note that the shapes of theconduction band in Figure 2 were obtained from the J–V distortion measurement and the changes inCBO are very small for the front side bandgaps in our CIGS layers. The simulated results are presentedin Figure 3 for the conversion efficiency parameters. As the front side bandgap increases from 1.1 eVto 1.37 eV, VOC gradually improves from 0.614 V to 0.648 V, but JSC and the fill factor (FF) decreases.As a result, the optimal front side bandgap is 1.18 eV in Figure 3. This can be attributed to the fact thatthe conduction band bending of the FGB helps the electrons to move over the band-offset of the ZnSinterface barrier in the SCR for the low front side bandgap. However, for the high front side bandgap,a conduction band barrier forms near the interface of the ZnS buffer layer and disturbs the electroncarrier transportation at the forward bias, so JSC and FF decreases gradually, as shown in Figure 3.

Energies 2018, 11, 1785 5 of 10Energies 2017, 10, x FOR PEER REVIEW 5 of 10

Figure 2. Conduction band diagram and electron concentration at the position in CIGS cell as a

function of the FGB.

Figure 3. Simulated solar cell conversion efficiency as a function of the front side bandgap: JSC, VOC,

FF, Efficiency of CIGS solar cell with ZnS buffer layer as a function of front side bandgap.

The BGB for reducing the recombination at the back side was also investigated using the

simulation. While the carriers generated near the front side can be quickly collected by the high

internal electric field in the SCR, the carriers generated near the back side (absorbing

long-wavelength photons) are slowly transferred and collected through carrier diffusion. The back

side field of the BGB can assist in the carrier transportation to increase JSC and, consequently, VOC.

Figure 4 shows the distributions of the electric field and extinction coefficient near the Mo interface

Figure 2. Conduction band diagram and electron concentration at the position in CIGS cell as a functionof the FGB.

Energies 2017, 10, x FOR PEER REVIEW 5 of 10

Figure 2. Conduction band diagram and electron concentration at the position in CIGS cell as a

function of the FGB.

Figure 3. Simulated solar cell conversion efficiency as a function of the front side bandgap: JSC, VOC,

FF, Efficiency of CIGS solar cell with ZnS buffer layer as a function of front side bandgap.

The BGB for reducing the recombination at the back side was also investigated using the

simulation. While the carriers generated near the front side can be quickly collected by the high

internal electric field in the SCR, the carriers generated near the back side (absorbing

long-wavelength photons) are slowly transferred and collected through carrier diffusion. The back

side field of the BGB can assist in the carrier transportation to increase JSC and, consequently, VOC.

Figure 4 shows the distributions of the electric field and extinction coefficient near the Mo interface

Figure 3. Simulated solar cell conversion efficiency as a function of the front side bandgap: JSC, VOC,FF, Efficiency of CIGS solar cell with ZnS buffer layer as a function of front side bandgap.

The BGB for reducing the recombination at the back side was also investigated using thesimulation. While the carriers generated near the front side can be quickly collected by the highinternal electric field in the SCR, the carriers generated near the back side (absorbing long-wavelengthphotons) are slowly transferred and collected through carrier diffusion. The back side field of the BGB

Energies 2018, 11, 1785 6 of 10

can assist in the carrier transportation to increase JSC and, consequently, VOC. Figure 4 shows thedistributions of the electric field and extinction coefficient near the Mo interface layer as a function ofthe back side bandgap, and it numerically shows that the BGB generates an additional back side fieldto the front side, improving the electron transport. The simulated conversion efficiency parameters ofthe BGB CIGS cells are plotted in Figure 5 for the back side bandgap at the Mo interface. The solar cellconversion efficiency by bandgap gradient is combination effects of back side electrical field, extinctioncoefficient and defect density. As the back side bandgap increases from 1.32 eV to 1.57 eV, the gradientin the conduction band significantly increases the electric field from 2027 V/cm to 3197 V/cm, reducingthe probability for carrier recombination at the back side, so that the solar cell conversion efficiencycan be improved. As the bandgap (due to Ga/(Ga + In) ratio) increases, the extinction coefficient(the absorption coefficient of CIGS) of Figure 4 decreases, however, JSC improves by 0.4 mA/cm2

because of the back side field in the BGB, as shown in Figure 5. Additionally, FF consistently degradeswith the increase in the bandgap because the influence of the defect density also increases. The backside bandgap grading should be optimized. The optimized value of the BGB is about 1.5 eV forachieving high efficiencies.

We evaluated the conversion efficiency of the CIGS solar cells for four different bandgap gradientprofiles, as shown in Figure 6. In addition to the DGB (the optimized one from the results of Figures 2and 4), a uniform (UNI) cell with no bandgap gradient, normal (NOR) cell only with the BGB, andreverse (REV) cell only with the FGB were simulated. Capacitance-voltage profiling (at 100 kHz, 50 mV)for the fabricated CIGS cell (DGB) was performed based on the approach in [9,10] acceptor (Na forp-CIGS) and donor (Nd for n-ZnS) concentrations were about 1.5 × 1016/cm3, and 1.5 × 1018/cm3,the depth of SCR was about 0.3 µm and capacitance was about 38.7 nF/cm2 at 0 V. Thus, for thecomparison, we intentionally set the simulation parameters so the SCR ended at a depth of nearly0.3 µm in the CIGS, and the average bandgap of the entire CIGS layer is about 1.27 eV for all thebandgap gradient structures shown in Figure 6 [18]. The conversion efficiencies of these solar cellsare summarized in Table 2. Even though the average bandgap of the CIGS layer is approximately thesame, the conversion efficiency of these solar cells is considerably different; the highest efficiency isobtained with the DGB and the lowest value is obtained with the REV solar cells.

Energies 2017, 10, x FOR PEER REVIEW 6 of 10

layer as a function of the back side bandgap, and it numerically shows that the BGB generates an

additional back side field to the front side, improving the electron transport. The simulated

conversion efficiency parameters of the BGB CIGS cells are plotted in Figure 5 for the back side

bandgap at the Mo interface. The solar cell conversion efficiency by bandgap gradient is

combination effects of back side electrical field, extinction coefficient and defect density. As the back

side bandgap increases from 1.32 eV to 1.57 eV, the gradient in the conduction band significantly

increases the electric field from 2027 V/cm to 3197 V/cm, reducing the probability for carrier

recombination at the back side, so that the solar cell conversion efficiency can be improved. As the

bandgap (due to Ga/(Ga + In) ratio) increases, the extinction coefficient (the absorption coefficient of

CIGS) of Figure 4 decreases, however, JSC improves by 0.4 mA/cm2 because of the back side field in

the BGB, as shown in Figure 5. Additionally, FF consistently degrades with the increase in the

bandgap because the influence of the defect density also increases. The back side bandgap grading

should be optimized. The optimized value of the BGB is about 1.5 eV for achieving high efficiencies.

We evaluated the conversion efficiency of the CIGS solar cells for four different bandgap

gradient profiles, as shown in Figure 6. In addition to the DGB (the optimized one from the results of

Figure 2 and 4), a uniform (UNI) cell with no bandgap gradient, normal (NOR) cell only with the

BGB, and reverse (REV) cell only with the FGB were simulated. Capacitance-voltage profiling (at 100

kHz, 50 mV) for the fabricated CIGS cell (DGB) was performed based on the approach in [9,10]

acceptor (Na for p-CIGS) and donor (Nd for n-ZnS) concentrations were about 1.5 × 1016/cm3, and 1.5 ×

1018/cm3, the depth of SCR was about 0.3 μm and capacitance was about 38.7 nF/cm2 at 0 V. Thus, for

the comparison, we intentionally set the simulation parameters so the SCR ended at a depth of

nearly 0.3 μm in the CIGS, and the average bandgap of the entire CIGS layer is about 1.27 eV for all

the bandgap gradient structures shown in Figure 6 [18]. The conversion efficiencies of these solar

cells are summarized in Table 2. Even though the average bandgap of the CIGS layer is

approximately the same, the conversion efficiency of these solar cells is considerably different; the

highest efficiency is obtained with the DGB and the lowest value is obtained with the REV solar cells.

Figure 4. Electric field and extinction coefficient at the position in CIGS cell as a function of the back

side bandgap.

Figure 4. Electric field and extinction coefficient at the position in CIGS cell as a function of the backside bandgap.

Energies 2018, 11, 1785 7 of 10Energies 2017, 10, x FOR PEER REVIEW 7 of 10

Figure 5. Simulated solar cell conversion efficiency at the position in CIGS cell as a function of the

back side bandgap: JSC, VOC, FF, Efficiency of CIGS solar cell with ZnS buffer layer as a function of

back side bandgap.

Figure 6. Bandgap energy at the position in CIGS cell for the different bandgap profiles.

Table 2. Performance of the experimented and simulated solar cells with different bandgap profiles.

Types JSC (mA/cm2) VOC (V) FF (%) Efficiency (%) n J0 (mA/cm2)

DGB 34.38 0.632 70.1 15.23 1.54 2.0 × 10−9

NOR 31.65 0.686 68.7 14.90 1.64 1.0 × 10−8

UNI 29.43 0.730 68.6 14.74 1.66 1.67 × 10−8

REV 27.06 0.757 66.3 13.57 1.69 1.4 × 10−8

Figure 5. Simulated solar cell conversion efficiency at the position in CIGS cell as a function of the backside bandgap: JSC, VOC, FF, Efficiency of CIGS solar cell with ZnS buffer layer as a function of backside bandgap.

Energies 2017, 10, x FOR PEER REVIEW 7 of 10

Figure 5. Simulated solar cell conversion efficiency at the position in CIGS cell as a function of the

back side bandgap: JSC, VOC, FF, Efficiency of CIGS solar cell with ZnS buffer layer as a function of

back side bandgap.

Figure 6. Bandgap energy at the position in CIGS cell for the different bandgap profiles.

Table 2. Performance of the experimented and simulated solar cells with different bandgap profiles.

Types JSC (mA/cm2) VOC (V) FF (%) Efficiency (%) n J0 (mA/cm2)

DGB 34.38 0.632 70.1 15.23 1.54 2.0 × 10−9

NOR 31.65 0.686 68.7 14.90 1.64 1.0 × 10−8

UNI 29.43 0.730 68.6 14.74 1.66 1.67 × 10−8

REV 27.06 0.757 66.3 13.57 1.69 1.4 × 10−8

Figure 6. Bandgap energy at the position in CIGS cell for the different bandgap profiles.

Table 2. Performance of the experimented and simulated solar cells with different bandgap profiles.

Types JSC (mA/cm2) VOC (V) FF (%) Efficiency (%) n J0 (mA/cm2)

DGB 34.38 0.632 70.1 15.23 1.54 2.0 × 10−9

NOR 31.65 0.686 68.7 14.90 1.64 1.0 × 10−8

UNI 29.43 0.730 68.6 14.74 1.66 1.67 × 10−8

REV 27.06 0.757 66.3 13.57 1.69 1.4 × 10−8

Energies 2018, 11, 1785 8 of 10

From the results of Table 2, we found that VOC behaves very interestingly; there is a strongcorrelation between VOC and the average bandgap in the SCR rather than that of an entire CIGSlayer. The average SCR bandgaps of the DGB, NOR, UNI, and REV cells are calculated as 1.144 eV,1.212 eV, 1.270 eV, and 1.337 eV, respectively, which are strongly proportional to the VOC of the cells.Because the defect density is modeled as proportional to the bandgap (as a result of the Ga/(Ga+ In) ratio), the defect density also increases with the bandgap: for example, the defect densitiesare 1.1 × 1014 /cm3 and 9.0 × 1014 /cm3 at the bandgap of 1.144 eV and 1.337 eV, respectively [15].However, VOC further increases from 0.632 V to 0.730 V in Table 2 because the influence of the defectdensity is small in our model (in our cells). We note that VOC does not significantly degrade until thedefect density of the SCR is much higher than 1.0 × 1015 /cm3 in our simulation.

As the DGB was determined to be the best bandgap gradient profile and VOC was selected withina desirable range using the average bandgap in SCR, we further investigated JSC and FF. We optimizedthe band structure of the DGB to obtain the best efficiency. To determine the main factor of JSC,we varied the individual parameters to simulate the DGB solar cell. For a ±10% variation in theminimum bandgap with the other parameters fixed, JSC changes by about ±0.7 mA/cm2, and the±10% variation in back side bandgap changes JSC by about ±0.2 mA/cm2. Similarly, for the ±10%variation in the front side bandgap, JSC changes by about ±0.07 mA/cm2. The ±10% variation indefect density in the entire CIGS layer of the DGB structure causes JSC to change by ±1.2 mA/cm2.The FF is known to be affected by the bandgap as well as the defect density. In general, FF graduallydeclines when the front or back side bandgap exceeds a certain value, as shown in Figures 3 and 5.The reduction of FF by the excessive front side bandgap is mainly attributed to the formation of theconduction band barrier obstructing electron collection, while the reduction by the excessive backside bandgap is because the defect traps become deeper as the bandgap increases with the Ga/(Ga+ In) ratio [19,20]. Finally, based on the simulation results, we optimized the DGB structure for theCIGS solar cell incorporating the Cd-free ZnS buffer layer and fabricated the CIGS cells with a size ofapproximately 0.25 cm2. Figure 7 presents the simulated and measured current density-voltage curvesunder 1-sun illumination; the fabricated cell efficiency is about 15.2% and the simulated curve of theCIGS solar cell matches well with the experimental curve.

Energies 2017, 10, x FOR PEER REVIEW 8 of 10

From the results of Table 2, we found that VOC behaves very interestingly; there is a strong

correlation between VOC and the average bandgap in the SCR rather than that of an entire CIGS layer.

The average SCR bandgaps of the DGB, NOR, UNI, and REV cells are calculated as 1.144 eV, 1.212

eV, 1.270 eV, and 1.337 eV, respectively, which are strongly proportional to the VOC of the cells.

Because the defect density is modeled as proportional to the bandgap (as a result of the Ga/(Ga + In)

ratio), the defect density also increases with the bandgap: for example, the defect densities are 1.1 ×

1014 /cm3 and 9.0 × 1014 /cm3 at the bandgap of 1.144 eV and 1.337 eV, respectively [15]. However, VOC

further increases from 0.632 V to 0.730 V in Table 2 because the influence of the defect density is

small in our model (in our cells). We note that VOC does not significantly degrade until the defect

density of the SCR is much higher than 1.0 × 1015 /cm3 in our simulation.

As the DGB was determined to be the best bandgap gradient profile and VOC was selected

within a desirable range using the average bandgap in SCR, we further investigated JSC and FF. We

optimized the band structure of the DGB to obtain the best efficiency. To determine the main factor

of JSC, we varied the individual parameters to simulate the DGB solar cell. For a ±10% variation in the

minimum bandgap with the other parameters fixed, JSC changes by about ±0.7 mA/cm2, and the ±10%

variation in back side bandgap changes JSC by about ±0.2 mA/cm2. Similarly, for the ±10% variation

in the front side bandgap, JSC changes by about ±0.07 mA/cm2. The ±10% variation in defect density

in the entire CIGS layer of the DGB structure causes JSC to change by ±1.2 mA/cm2. The FF is known

to be affected by the bandgap as well as the defect density. In general, FF gradually declines when

the front or back side bandgap exceeds a certain value, as shown in Figures 3 and 5. The reduction of

FF by the excessive front side bandgap is mainly attributed to the formation of the conduction band

barrier obstructing electron collection, while the reduction by the excessive back side bandgap is

because the defect traps become deeper as the bandgap increases with the Ga/(Ga + In) ratio [19,20].

Finally, based on the simulation results, we optimized the DGB structure for the CIGS solar cell

incorporating the Cd-free ZnS buffer layer and fabricated the CIGS cells with a size of approximately

0.25 cm2. Figure 7 presents the simulated and measured current density-voltage curves under 1-sun

illumination; the fabricated cell efficiency is about 15.2% and the simulated curve of the CIGS solar

cell matches well with the experimental curve.

Figure 7. Light current density-voltage curves of the experimental results (solid symbol) and

simulation results (open symbol).

4. Conclusions

We numerically investigated the characteristics of the gradient bandgap profile in the

absorption layer to evaluate the conversion efficiency of the CIGS solar cell with a Cd-free ZnS

buffer layer using a TCAD simulation. The Ga-related parameters, such as defect density, bandgap,

Figure 7. Light current density-voltage curves of the experimental results (solid symbol) and simulationresults (open symbol).

Energies 2018, 11, 1785 9 of 10

4. Conclusions

We numerically investigated the characteristics of the gradient bandgap profile in the absorptionlayer to evaluate the conversion efficiency of the CIGS solar cell with a Cd-free ZnS buffer layerusing a TCAD simulation. The Ga-related parameters, such as defect density, bandgap, and electronaffinity, can be controlled by the Ga/(Ga + In) ratio in the CIGS and were modeled and presentedfor the simulation. In the simulation, we considered the FGB, and the BGB formed by spatiallyvarying the Ga content. The FGB enhanced the VOC, but the front side bandgap should be carefullydesigned to not disturb the electron transport at the conduction band barrier near the ZnS buffer layer.The BGB improved the electron transport at the back surface, but the excessive back side bandgapcould induce the defect traps to move deeper in the CIGS, resulting in the degradation of the FF ofthe CIGS cell. By simulating the various types of band structures, DGB, NOR, REV, and UNI, wedemonstrated that the DGB was the most effective structure for achieving the highest cell efficiency.We demonstrated that VOC strongly correlated with the average bandgap in the SCR and could beselected within the desirable range using the DGB; we also numerically evaluated JSC and FF for theband structural parameters of the CIGS cell. In this approach, we designed the optimal band structureof the CIGS cell with a Cd-free ZnS buffer layer and demonstrated a cell efficiency of more than 15%for the fabricated CIGS solar cells. The numerical design procedure of this work will be useful fordeveloping high-conversion efficiency CIGS cells incorporating a gradient bandgap structure andCd-free buffer layer.

Author Contributions: All authors contributed to this work. Conceptualization, simulation and analysis, J. Park.;Investigation, and supervision, M. Shin.

Funding: This work was supported by the Korea Institute of Energy Technology Evaluation and Planning (KETEP)and the Ministry of Trade, Industry & Energy (MOTIE) of the Republic of Korea (No. 20143030011530), and alsosupported by the Pioneer Research Center Program through the National Research Foundation of Korea fundedby the Ministry of Science, ICT & Future Planning (NRF-2013M3C1A3065040).

Acknowledgments: This research was supported by the Solar Energy Group of the Samsung Electronics Inc.;the authors acknowledge the technical support and the assistance offered by Dr. Myeong-Woo Kim at theSamsung SDI. The authors also appreciate Dr. Dong-Seop Kim (Vice President of Samsung SDI) for his adviceand leadership.

Conflicts of Interest: The authors declare no conflict of interest.

References

1. Green, M.A.; Emery, K.; Hishikawa, Y.; Warta, W.; Dunlop, E.D. Solar cell efficiency tables (version 49).Prog. Photovolt. Res. Appl. 2016, 25, 3–13. [CrossRef]

2. Kessler, F.; Herrmann, D.; Powalla, M. Approaches to flexible CIGS thin-film solar cells. Thin Solid Films2005, 480, 491–498. [CrossRef]

3. Nakada, T.; Mizutani, M.; Hagiwara, Y.; Kunioka, A. High-efficiency Cu(In,Ga)Se2 thin-film solar cells witha CBD-ZnS buffer layer. Solar Energy Mater. Solar Cells 2001, 67, 255–260. [CrossRef]

4. Gloeckler, M.; Sites, J.R. Band-gap grading in Cu(In,Ga)Se2 solar cells. J. Phys. Chem. Solids 2005, 66,1891–1894. [CrossRef]

5. Song, J.; Li, S.S.; Huang, C.H.; Crisalle, O.D.; Anderson, T.J. Device modeling and simulation of theperformance of Cu(In1−x,Gax)Se2 solar cells. Solid-State Electronics 2004, 48, 73–79. [CrossRef]

6. Song, S.H.; Nagaich, K.; Aydil, E.S.; Feist, R.; Haley, R.; Campbell, S.A. Structure optimization for a highefficiency CIGS solar cell. In Proceedings of the 2010 35th IEEE Photovoltaic Specialists Conference, Honolulu,HI, USA, 20–25 June 2010; pp. 2488–2492.

7. Dullweber, T.; Lundberg, O.; Malmstörm, J.; Bodegård, M.; Stolt, L.; Rau, U.; Schock, H.W.; Werner, J.H.Werner Back surface band gap gradings in Cu(In,Ga)Se2 solar cells. Thin Solid Films 2001, 387, 11–13.[CrossRef]

8. Sozzi, G.; Troni, F.; Menozzi, R. On the combined effects of window/buffer and buffer/absorberconduction-band offsets, buffer thickness and doping on thin-film. Sol. Energy Mater. Sol. Cells 2014.[CrossRef]

Energies 2018, 11, 1785 10 of 10

9. Heo, S.; Lee, H.-I.; Song, T.; Park, J.-B.; Ko, D.-S.; Chung, J.; Kim, K.; Kim, S.H.; Yun, D.-J.; Ham, Y.; et al.Direct band gap measurement of Cu(In,Ga)(Se,S)2 thin films using high-resolution reflection electron energyloss spectroscopy. Appl. Phys. Lett. 2015, 106. [CrossRef]

10. Lee, D.; Yang, J.Y.; Kim, Y.-S.; Mo, C.B.; Park, S.; Kim, B.J.; Kim, D.; Nam, J.; Kang, Y. Effects of the Cu/(Ga +In) ratio on the bulk and interface properties of Cu(InGa)(SSe)2 solar cells. Sol. Energy Mater. Sol. Cells 2016,149, 195–203. [CrossRef]

11. Luber, E.J.; Buriak, J.M. Reporting performance in organic photovoltaic devices. ACS Nano 2013, 7, 4708–4714.[CrossRef] [PubMed]

12. Durante Rincón, C.A.; Hernández, E.; Alonso, M.I.; Garriga, M.; Wasim, S.M.; Rincón, C.; León, M. Opticaltransitions near the band edge in bulk CuInxGa1−xSe2 from ellipsometric measurements. Mater. Chem. Phys.2001, 70, 300–304. [CrossRef]

13. Wei, S.-H.; Zhang, S.B.; Zunger, A. Effects of Ga addition to CuInSe2 on its electronic, structural, and defectproperties. Appl. Phys. Lett. 1998, 72, 3199–3201. [CrossRef]

14. Frisk, C.; Platzer-Björkman, C.; Olsson, J.; Szaniawski, P.; Wätjen, J.T.; Fjällström, V.; Salomé, P.; Edoff, M.Optimizing Ga-profiles for highly efficient Cu(In,Ga)Se2 thin film solar cells in simple and complex defectmodels. J. Phys. D Appl. Phys. 2014, 47, 485104–485115. [CrossRef]

15. Hanna, G.; Jasenek, A.; Rau, U.; Schock, H.W. Influence of the Ga-content on the bulk defect densities ofCu(In, Ga)Se2. Thin Solid Films 2001, 387, 71–73. [CrossRef]

16. Gloeckler, M.; Fahrenbruch, A.L.; Sites, J.R. Numerical modeling of CIGS and CdTe solar cells: setting thebaseline. Proc. 3rd World Conf. Photovolt. Energy Conversion 2003, 1, 491–494. [CrossRef]

17. Pudov, A.O.; Kanevce, A.; Al-Thani, H.A.; Sites, J.R.; Hasoon, F.S. Secondary barriers in CdS-CuIn1−xGaxSe2

solar cells. J. Appl. Phys. 2005, 97, 064901. [CrossRef]18. Barkhouse, D.A.R.; Gunawan, O.; Gokmen, T.; Todorov, T.K.; Mitzi, D.B. Device characteristics of a 10.1%

hydrazine-processed Cu2ZnSn(Se,S)4 solar cell. Prog. Photovolt. 2012, 20, 6–11. [CrossRef]19. Heath, J.T.; Cohen, J.D.; Shafarman, W.N.; Liao, D.X.; Rockett, A.A. Effect of Ga content on defect states in

CuIn1−xGaxSe2 photovoltaic devices. Appl. Phys. Lett. 2002, 80, 4540–4542. [CrossRef]20. Cao, Q.; Gunawan, O.; Copel, M.; Reuter, K.B.; Chey, S.J.; Deline, V.R.; Mitzi, D.B. Defects in Cu(In,Ga)Se2

chalcopyrite semiconductors: A comparative study of material properties, defect states, and photovoltaicperformance. Adv. Energy Mater. 2011, 1, 845–853. [CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open accessarticle distributed under the terms and conditions of the Creative Commons Attribution(CC BY) license (http://creativecommons.org/licenses/by/4.0/).