Open access peer-reviewed chapter

Kesterite Cu2ZnSnS4-xSex Thin Film Solar Cells

Written By

Kaiwen Sun, Fangyang Liu and Xiaojing Hao

Submitted: 11 November 2021 Reviewed: 24 November 2021 Published: 31 December 2021

DOI: 10.5772/intechopen.101744

From the Edited Volume

Thin Films Photovoltaics

Edited by Beddiaf Zaidi and Chander Shekhar

Chapter metrics overview

421 Chapter Downloads

View Full Metrics

Abstract

Kesterite Cu2ZnSnS4-xSex (CZTS) is a promising thin film photovoltaic (PV) material with low cost and nontoxic constitute as well as decent PV properties, being regarded as a PV technology that is truly compatible with terawatt deployment. The kesterite CZTS thin film solar cell has experienced impressive development since its first report in 1996 with power conversion efficiencies (PCEs) of only 0.66% to current highest value of 13.0%, while the understanding of the material, device physics, and loss mechanism is increasingly demanded. This chapter will review the development history of kesterite technology, present the basic material properties, and summarize the loss mechanism and strategies to tackle these problems to date. This chapter will help researchers have brief background knowledge of kesterite CZTS technology and understand the future direction to further propel this new technology forward.

Keywords

  • kesterite
  • Cu2ZnSn(S
  • Se)4
  • CZTS
  • thin film solar cells
  • loss mechanism

1. Introduction

Thin film photovoltaic (PV) technologies such as Cu(In,Ga)Se2 (CIGS) and CdTe have already demonstrated more than 20% power conversion efficiency (PEC) [1, 2] and are at their commercial stage. However, considering the Restriction of Hazardous Substances Directive (RoHS) adopted in European Union [3] and the recent classification of critical raw materials (CRM) by the European Commission [4], emerging thin film PV technologies with RoHS-compliant and CRM-free constituents are increasingly desirable. Kesterite copper-zinc-tin-selenosulfide and related quaternary semiconductor represented by chemical formula Cu2ZnSnS4-xSex is generally accepted as one promising option for low-cost and nontoxic thin film PV.

The family of kesterite Cu2ZnSnS4-xSex includes pure sulfide Cu2ZnSnS4 (CZTS), pure selenide Cu2ZnSnSe4 (CZTSe), and selenosulfide Cu2ZnSn(S,Se)4 (CZTSSe) and other related compound semiconductors. Herein we abbreviate all Cu2ZnSnS4-xSex compounds as CZTS. The formation of CZTS is derived from cation mutations (cross-substitutions) in CuInS2 by replacing two In atoms with one Zn and one Sn [5, 6]. CZTS possess high absorption coefficient of over 104 cm−1, tunable band gap that can range from 1.0 to 1.5 eV to favorably match the solar spectrum, intrinsic p-type conductivity, and a three-dimensional symmetry of carrier transport [7, 8, 9]. These decent photovoltaic properties of CZTS have attracted considerable attention and enabled its rapid development in the last decades. The highest power conversion efficiencies (PCEs) of Cu2ZnSnS4, Cu2ZnSnSe4and Cu2ZnSn(S,Se)4 have been set to 11%, 12.5%, and 13%, respectively [10, 11, 12], which represent the best PCE among the emerging RoHS-compliant and CRM-free inorganic thin-film PV technologies.

In this chapter, the development history and current status of CZTS PV technology will be briefly introduced. The basic physical and chemical properties of CZTS thin film will be described to help readers have a better understanding of the material. The device architecture and absorber processing will be touched; finally. The limiting factors as well as the perspective for future development of this technology will also be reviewed and addressed.

Advertisement

2. The development history of CZTS

The CZTS single crystal was first grown by Nitsche, Sargent, and Wild in 1966 when they tried to prepare a serious of AI2BIICIVX4-type quaternary chalcogenides using iodine vapor transport [13]. The photovoltaic effect of CZTS was exhibited for the first time in 1988 by Ito and Nakazawa on a heterojunction diode consisting of cadmium-tin-oxide transparent conductive layer and CZTS thin film on a stainless steel substrate. The open-circuit voltage was measured to be 165 mV under AM1.5 illumination [14]. The open-circuit voltage was increased to 250 mV after annealing the device in air, and short-circuit current of 0.1 mA/cm2 was achieved [15].

The first CZTS solar cell with PCE of 0.66% was reported by Katagiri et al. in 1996 at PVSEC-9, with the device structure of ZnO:Al/CdS/CZTS/Mo/soda lime glass (SLG) substrate [16]. The CZTS thin film was fabricated by vapor-phase sulfurization of E–B-evaporated precursors [17]. In 1997, Friedlmeier et al. reported the CZTS solar cell with PCE of 2.3% and open-circuit voltage of 570 mV based on thermal evaporation [18]. In 1999, the Katagiri group improved the PCE up to 2.63% [19], after that they did a lot of work on optimizing the CZTS thin film and pushed the efficiency to 5.74% until 2007 [19, 20, 21, 22, 23, 24].

The CZTS solar cell started to gain intensive interest from academic and industry community from 2008 when its efficiency was further boosted to 6.7% by Katagiri et al. and when the CIGS solar cell became mature in its commercialization stage [25]. A lot of research institutes and solar cell manufactures such as Toyota, IBM, NREL (National Renewable Energy Laboratory), Solar Frontier, EMPA (Swiss Federal Laboratories for Materials Science and Technology), HZB (Helmholtz-Zentrum Berlin), ZSW (Centre for Solar Energy and Hydrogen Research Baden-Württemberg), UNSW (University of New South Wales), and so on involved in the development of this technology and significant advances have been achieved in the following decade [10, 26, 27, 28, 29, 30, 31, 32, 33]. During the development period, Se incorporation has attracted significant attention and led to impressive progress in PCE [34, 35, 36, 37]. Moreover, various fabrication methods including vacuum deposition process and non-vacuum process such as solution method, electrochemical deposition, etc., have been developed for fabricating the CZTS thin film [29, 37, 38, 39, 40].

An important milestone in the development of CZTS solar cell is the 10% benchmark efficiency breakthrough achieved by IBM Thomas J. Watson Research Center in 2011 [34], which shows substantial commercial promise for the CZTS-based class of thin film PV materials. This breakthrough also established the leading position of IBM in CZTS PV technology research, represented by a serious of world record efficiencies [38, 39, 41]. The highest PCE for CZTS solar cell has been stagnant at 12.6% for more than 6 years since 2014 when IBM last updated their efficiency breakthrough [39]. Thanks to the better understanding of the CZTS material and the loss mechanism of the device, quite a few groups reported CZTS PCE close to the 12.6% record efficiency in recent years [11, 42, 43, 44, 45, 46]. The recently announced NREL Best Research-Cell Efficiency Chart included the newly refreshed CZTS record efficiency of 13% achieved by Xin et al. from NJUPT [12]. This small step breakthrough comes from great efforts in this research area and hopefully will bring more interests and confidence to the CZTS R&D community.

Advertisement

3. The physical and chemical properties of CZTS thin film

The investigation and understanding of nature of the CZTS thin film are crucial for further developing this technology. Intensive studies have been conducted during the rapid development stage; therefore, the physical and chemical properties of CZTS have been well revealed.

3.1 Crystal structure

As we briefly introduced earlier, CZTS is derived from the cation mutations of CuInS2 (CIS), while both are originated from the binary II–VI semiconductors adopting the cubic zinc-blende (or hexagonal wurtzite) structure as shown in Figure 1.

Figure 1.

Schematic illustration of the origin of CZTS structure. Reproduced from [6] with permission from the Royal Society of Chemistry.

Similarly, the structure of CZTS is also derived from ternary I-III-VI2 compounds. In general, as shown in Figure 2, chalcopyrite (CH) and CuAu-like (CA) structures are two fundamental I-III-VI2 structures that obey the octet rule [5, 47, 48]. Therefore, the quaternary CZTS is well known as two principal structures, kesterite (space group I4, Figure 2d) and stannite-type (space group I42m,Figure 2e), which are derived from CH structure and CA structure, respectively. Other primitive mixed CA structure (PMCA) (space group I42m,Figure 2f) derived from CA structure has also been reported in CZTS [5, 47, 48]. As the most common structures discussed in literature are kesterite and stannite-type, we only focus on these two structures in this section. The two structures are closely related with the main difference of cation arrangement. Both structures are composed of a cubic close-packed lattice of S anions, with half of the tetrahedral interstices occupied by cations. Sn atoms occupy the same fixed positions in both structures, but the Cu and Zn atoms are in different position [49, 50]. In kesterite structure, cation layers of CuZn, CuSn, CuZn, CuZSn alternated at z = 0, ¼, ½, and ¾ respectively (Figure 2d), while in stannite structure, ZnSn and Cu2 layers alternate with each other (Figure 2e). The similarities of structure make it difficult to distinguish kesterite and stannite experimentally by employing common X-ray diffraction and Raman spectroscopy techniques [51, 52]. Only advanced techniques such as neutron powder diffraction analysis are capable to tell them apart [50, 53].

Figure 2.

The crystal structure of (a) zinc-blende ZnS, (b) chalcopyrite CuInS2, (c) CuAu-like CuInS2, (d) Kesterite-type Cu2ZnSnS4, (e) stannite-type Cu2ZnSnS4, and (f) PMCA-Cu2ZnSnS4. Reproduced from [47] with permission from the American Physical Society.

CZTS usually exists as kesterite-type structure, which is more stable thermodynamically than the stannite-type [54]. This is in agreement with the experimental observation [21, 50, 55, 56]. Plenty of theoretical studies have also confirmed that the kesterite-type structure is the ground state structure in CZTS [5, 47, 48, 57, 58, 59]. This is also the reason that CZTS is named after Kesterite because it crystallizes as kesterite structure. However, the energy difference between the kesterite and the stannite-type structure is rather small [5, 47, 48, 57, 58, 59]. This indicates that kesterite structure should be formed under equilibrium growth conditions, but both phases may exist, especially when growth method and conditions are changed, it should be relatively easy to grow materials with mixed phases. This may also partly explain the existence of disorder structure with more random distribution of the Cu and Zn on the cation positions [60], which will be discussed in more details in the following sections.

3.2 Electronic band structure

Kesterite CZTS has a similar tetrahedral bonds geometry as traditional group I–V, III–V, and II–VI semiconductors. It obeys Lewis’ octet rule with eight electrons around each anion atom (S or Se); the four bonds of each anion therefore form together a close valence shell [48]. However, there are fundamental differences between the Cu-based quaternary compound and the group I–V, III–V, and II–VI binary semiconductors. First of all, the bonds in CZTS involve Cu-d–anion-p hybridized antibonding states. This weakens the bonds in CZTS. Second, Cu in CZTS has only one valence s-like electron while the group III–V (group II–VI) semiconductors have three (two) valence s-like electrons for their cations.

Several first principle studies have revealed the electronic band structure of CZTS [47, 48, 58, 59, 61, 62]. An example is as shown in Figure 3, the calculated band structures of kesterite Cu2ZnSnS4 and Cu2ZnSnSe4 from different methods are presented. Overall, the kesterite CZTS materials are all direct-gap semiconductors. Different calculation methods generated slightly different values, for example, generalized gradient approximation (GGA) method calculated the band gap energy for Cu2ZnSnS4 and Cu2ZnSnSe4 as 1.56 eV and 1.05 eV, respectively. Corresponding band gap values from hybrid functional calculations (HSE06) are 1.47 and 0.90 eV, while the values from the GW0 calculation are 1.57, and 0.72 eV. These values are in agreement with other calculated data and available experimental measurement, converging to the results that Eg ≈ 1.5 eV in CZTS and Eg ≈ 1.0 eV in CZTSe [14, 24, 28, 64, 65, 66, 67, 68, 69, 70, 71, 72] to vary linearly as a function of the Se content x [73] with Eg = 1.47, 1.30, 1.17, 1.01, and 0.90 eV for x = 0, 1/4, 1/2, 3/4, and 1, using the HSE06 potential. This linear relationship also agrees with experimental results [66, 70, 71].

Figure 3.

The electronic band structure of the kesterite structures of Cu2ZnSnS4 and Cu2ZnSnSe4 along four symmetry directions. The energy refers to the VBM (dashed lines). The spin–orbit interaction is included, but the index of the bands refers to spin-independent bands where c1 represents the lowest CB and v1 represents the topmost VB. Different line shapes represent different calculation methods [63]. Reproduced from [59] with permission from American Institute of Physics.

The density of states (DOS) of kesterite Cu2ZnSnS4 and Cu2ZnSnSe4 can also be calculated as illustrated in Figure 4 [63]. It is not surprising to see that Cu2ZnSnS4 and Cu2ZnSnSe4 show comparable DOS as they are in same tetrahedral bond geometries. The DOS of conduction bands in Cu2ZnSnS4 is ~0.5 eV higher than that of Cu2ZnSnSe4 because of large energy gap. It is common that in Cu-based chalcogenides including the quaternary (for example, CIS) and ternary compounds (for example, CZTS), the valence band maximum (VBM) is derived mainly from the hybridization of anion p and Cu d states because Cu has higher d orbital energy than Zn, Ga, In, and Sn [59, 74]. In CZTS, the valence p level of S is lower in energy than Se, thus the VBM of the sulfides is lower than that of the selenides. This difference is reduced by anion p–Cu d overlap (p-d hybridization) because the hybridization is stronger in the shorter Cu–S bond and pushes the antibonding VBM level of the sulfide up relative to that of the selenide. Therefore, the valence band offset between the Cu2ZnSnS4 and Cu2ZnSnSe4 is less than 0.2 eV. The DOS of the CZTS conduction band minimum (CBM) is primarily controlled by the Sn-s and anion p-like states due to the lower s orbital energy of Sn than the other cations [59, 74]. More importantly, in CZTS the lowest CB is separated from the higher energy bands and is therefore more localized in energy. Therefore, the CBM of CZTS is expected to vary depending on the alloying on the Sn-site with other group-IV elements (e.g., Ge) due to this localized CB. The band gap energy can therefore be tailored by cation alloying for an optimized optical efficiency of the materials.

Figure 4.

Atomic and angular momentum resolved DOS of kesterite Cu2ZnSnS4 and Cu2ZnSnSe4 presented with a 70 meV Lorentzian broadening. Reproduced from [59] with permission from American Institute of Physics.

3.3 Composition variation and phase competition

As a quaternary semiconductor, CZTS consists of three metals and up to two chalcogens, which provides a wide range for composition variation and secondary phase formation. However, high-efficiency CZTS thin film solar cells require single-phase kesterite absorber, and some secondary phases have been reported detrimental to the device performance. This section reviews the stable phase region of CZTS and possible secondary phases likely to be formed in the quaternary system.

The 3D CZTS quaternary phase diagram as shown in Figure 5 can be deduced from Cu2S-ZnS-SnS2 pseudo-ternary diagram [77] and CuS-ZnS-SnS pseudo-ternary phase diagram [75] (selenide kesterite system also refers to Cu2SnSe3–SnSe2–ZnSe diagram [78]. In addition, the stability region of the CZTS in the atomic chemical potential space can be calculated as shown in Figure 5 [76, 79]. All these diagrams indicate that volume of the stable CZTS region is small, and the slight deviation (maximum 1–2 at%) outside this space will cause the formation of different secondary phases [5, 80]. The narrow stable window also implies that the composition control and the chemical potential control are crucial for the growth of high-quality and single-phase CZTS absorber. The Zn content is particularly important because the stable region along the μZn axis is much narrower. Moreover, it is well accepted experimentally and theoretically that high-efficiency solar cells need CZTS absorber with Zn-rich and Cu-poor composition [8, 9, 10, 27, 28, 33, 67, 81]. This makes it more difficult to control the secondary phases formation.

Figure 5.

Left: CZTS quaternary phase diagram including the known phases. Reproduced from [75] with permission from American Institute of Physics. Right: The calculated chemical-potential stability diagram of Cu2ZnSnS4 in a 2D Cu-rich plane (the stable 3D region is inset) reproduced from [76] with permission from American Institute of Physics.

According to the phase diagram and literature report, the most common secondary phases in the Cu-Zn-Sn-S/Se system are list in Table 1 [92]. The influence of the secondary phases on the solar cell performance depends on their position in the film as well as their physical properties. For example, Cu2S(Se) phases in the final film may act as shunting path due to both the high conductivity and contact with front and back interfaces. However, Cu2S(Se) is also an important fluxing agent to promote lateral grain growth during the film growth [28]. Generally, If the secondary phase has a lower band gap than the CZTS absorber, it will limit the open-circuit voltage of the solar cell. While secondary phases with higher band gaps than CZTS are less detrimental; however, they can block the transport when present in large amounts [93] or at least increase the series resistance [94]. ZnS(e) is the most likely secondary phase considering the phase diagram and especially when adapting the Zn-rich and Cu-poor condition. Fortunately, ZnS(e) is a large band gap compound and is expected to be rather benign if present in small amounts. It has even been reported that ZnS with similar crystalline structures to CZTS may passivate grain boundaries or heterojunction interface [81, 95, 96] by reducing strain and lowering recombination velocities at the grain interfaces. The tin compounds are unlikely to form because they are usually volatile [97] and will evaporate in most preparation conditions. As shown in Table 1, one unfavorable low band gap secondary phase in both Cu2ZnSnS4 and Cu2ZnSnSe4 system is the ternary Cu2SnS(e)3, which is also one reason for optimal composition range (Zn-rich and Cu-poor) found for the best solar cells. However, the non-homogeneous composition across the CZTS film under normal preparation conditions still provides the possibility to form such detrimental secondary phases [98]. Therefore, how to control the amount and position of the secondary phases in the CZTS absorber needs to be elaborately studied.

CompoundBand gap (eV)Ref.CompoundBand gap (eV)Ref.
Cu2ZnSnS41.5[80]Cu2ZnSnSe41.0[80]
Cu2SnS31.0[82]Cu2SnSe30.8[83]
ZnS3.7[84]ZnSe2.7[84]
SnS2~2.5[85]SnSe21.0–1.6[86]
SnS1.0 indirect, 1.3 direct[87, 88]SnSe1.3[89]
Cu2S1.2[90]Cu2Se1.2[91]

Table 1.

Most common secondary phases in the Cu-Zn-Sn-S/Se system.

3.4 Lattice defects

The formation and properties of lattice defects are important parameters of semiconductor materials and are crucial to the function of photovoltaic devices, because they directly influence the generation, separation, and recombination of electron–hole pairs. The lattice defects (e.g., vacancies, interstitials, antisites) in kesterite CZTS system are complicated because of the increased number of component elements and the similar cation size as well as small chemical mismatch of Cu+ and Zn2+. In this section, the formation and ionization of the lattice defects in kesterite will be briefly reviewed. More importantly, the underlying mechanism of p-type conductivity, Zn-rich and Cu-poor condition growth condition, as well as limiting factors for device performance in kesterite CZTS solar cells from the perspective of lattice defects will be discussed.

The concentration of the lattice defects is determined by their formation energy. In Figure 6, the calculated formation energies of different defects are plotted as functions of the Fermi energy (0 means that the Fermi energy is at VBM, while 1.5 or 1.0 eV means that Fermi energy is at CBM). It is obvious that in both Cu2ZnSnS4 and Cu2ZnSnSe4, CuZn antisite is the lowest energy defects, which is different from the defect properties of their parent compounds (CuInSe2 or CuGaSe2) where the dominant defect is the Cu vacancy VCu [76, 79, 99]. In addition, the formation energies of most acceptor defects are lower than those of donor defects, explaining the intrinsic p-type conductivity observed in literature [21, 48, 64, 84, 100, 101, 102, 103, 104, 105, 106, 107, 108, 109, 110].

Figure 6.

Calculated defect formation energy as a function of the Fermi energy at the thermodynamic chemical-potential point a (from Figure 5right) for Cu2SnZnS4 (left) and Cu2SnZnSe4 (right). For each value of the Fermi energy, only the most stable charge state is plotted, with the filled circles (change of slope) representing a change in charge state (transition energy level). Reproduced from [76, 79] with permission from American Institute of Physics and American Physical Society.

Another important parameter of lattice defects is their ionization (transition) levels, which determines if they can produce free carriers and contribute to the electrical conductivity. The calculated ionization levels of intrinsic defects in the band gap of Cu2ZnSnS4 and Cu2ZnSnSe4 are shown in Figure 7. First of all, in both cases, the dominant defect CuZn has an acceptor level (0.11 eV and 0.15 eV above VBM in Cu2ZnSnSe4 and in Cu2ZnSnS4, respectively) deeper than that of VCu. The relatively deep level of dominant antisite defect in CZTS is not favorable to the device performance because it will limit the open-circuit voltage. This also partially explains why Zn-rich and Cu-poor condition has normally been found to beneficial to solar cell efficiency because it could decrease the formation energy and enhance the population of shallow VCu. The other acceptor defects (e.g., CuSn, ZnSn, VZn, and VSn) have higher formation energy; therefore, they have negligible contribution to the p-type conductivity. However, they may act as recombination centers especially for those deep transition levels such as (4−/3-) and (3−/2-) in band gap as seen in Figure 7.

Figure 7.

The ionization levels of intrinsic defects in the band gaps of Cu2ZnSnS4 (top) and Cu2ZnSnSe4 (bottom). The red bars show the acceptor levels, and the blue bars show the donor levels, with the initial and final charge states labeled in parentheses. Reproduced from [79] with permission from American Physical Society.

In addition to the point defects, various self-compensated defect clusters can be formed in CZTS due to the large amount of low-energy intrinsic defects. Defect compensation in ternary CIS is well known to have electrically benign character because the intrinsic defects undergo self-passivation through the formation of defect complexes such as [2VCu + InCu2+]. Therefore, it is also interesting to see if the same behavior can be observed in quaternary kesterites. CuZn and ZnCu are the lowest-energy acceptor and donor defects, respectively; therefore, antisite pair [CuZn + ZnCu+] also shows extremely low formation energy. Fortunately, its impact on the electronic structure and optical properties is relatively weak. The detrimental defect clusters are those composed of deep level defects, such as SnZn, SnCu, CuSn, and Zni. For example, clusters with SnZn induce large conduction band edge downshift, which could limit the solar cell performance, because the induced states are deep and may trap photo-generated electrons from the high conduction band. [2CuZn + SnZn] clusters could present in high population if in single-phase CZTS (Cu/(Zn + Sn) and Zn/Sn ratios near 1) chemical potential conditions and be detrimental to the solar cell performance. The Zn-rich and Cu-poor condition could prevent the formation of [2CuZn + SnZn] clusters because its formation energy is very sensitive to the chemical potential of Zn. This again partly explains that the Zn-rich and Cu-poor condition is beneficial to solar cell efficiencies from the defect perspective.

Advertisement

4. The path towards high-efficiency kesterite solar cell

4.1 Device architecture

The typical device architecture (Figure 8) of kesterite-based solar cell inherits from its predecessor chalcopyrite CuInSe2 solar cells due to their similar optical and electronic properties. Metal Mo layer with thickness of 500 nm–1000 nm is usually deposited with sputtering on soda lime glass or other substrates as back contact. Then the kesterite CZTS absorber layer is deposited on top of the Mo layer. The p-n junction is formed by p-type CZTS and the following deposited n-type buffer layer. Typical n-type buffer layer in kesterite and chalcopyrite solar cell is 50 nm–100 nm thick CdS layer usually by chemical bath deposition. Alternative buffer materials such as ZnSnO, Zn(O, S), ZnCdS, etc., have been explored, especially in pure sulfide Cu2ZnSnS4 to tackle the unfavorable band alignment found at p-n junction. Next, a 50 nm–100 nm high-resistive intrinsic ZnO (i-ZnO) is deposited. Subsequently, the device structure is completed by deposition of 200 nm–500 nm thick Al-doped ZnO (AZO) or indium tin oxide (ITO) transparent conducting oxide (TCO) layer as the front contact. Ni/Al metal contacts are deposited on the TCO layer for improved current collection. Antireflection coating such as MgF2 is often deposited on top of the cell to reduce the reflection loss.

Figure 8.

Schematic device structure of a typical kesterite Cu2ZnSnS4 thin film solar cell.

Various deposition methods have been investigated for producing high-quality CZTS layer during the development of this technology. These deposition methods are broadly classified as vacuum-based and non-vacuum-based techniques. The vacuum-based methods are usually considered easy to be expanded to commercial scale because of the precise process control. All physical vapor deposition techniques including thermal evaporation, E-beam evaporation [17], sputtering [10, 33, 81, 111], pulsed laser deposition [112] fall into this category. The non-vacuum-based methods are always regarded as low-cost, high-throughput techniques and feasible in roll-to-roll production. These methods usually involve chemical or solution process, such as electrochemical deposition [40], nanoparticle-based synthesis [113, 114], sol–gel spin coating [44, 115], chemical bath deposition (CBD) [116], successive ion layer adsorption and reaction (SILAR) [117], screen printing [118], etc.

4.2 Loss mechanism and solutions

4.2.1 Open-circuit voltage (VOC)

In kesterite CZTS solar cells, it has been widely accepted that VOC loss accounts for the majority (more than 50% [119]) of the efficiency gap between the best performance device and the theoretical limit (i.e., Shockley–Queisser radiative limit [120]). VOC loss is determined by the recombination path in the device. The dominant recombination can occur in the bulk of the absorber in the quasi-neutral zone as band-to-band recombination, via defects, or in the space charge region. Another important recombination path can be located at the heterojunction interface between the buffer/window layer and absorber.

Therefore, the absorber problems discussed in Section 3 such as abundant point defects and defect clusters (i.e., cation disordering), composition variation with narrow stable region are believed to contribute to the recombination. The Cu and Zn substitution with low energy causes a large population of antisite defects such as CuZn and ZnCu and related defects complexes. Consequently, severe electrostatic potential fluctuation and associated band tailing can be observed [121]. In addition, the microinhomogeneities in composition, nonuniform strain, as well as formation of secondary phases lead to band gap fluctuations, which are also detrimental to VOC. Moreover, acceptor-like CuZn defects at the interface will cause Fermi level pinning to a low energy level [122], thus reducing the band bending in the absorber and weakening the electric field. To alleviate the recombination occurring in absorber layer, great efforts have been made over recent years. One solution that has been intensively investigated is cation substitutions because the introduction of larger size cations could result in better cationic ordering, thereby reducing the defects formed because of the Cu and Zn substitution. Using Ag as a substitute for Cu [115], Cd as a substitute for Zn [123, 124], and Ge substitute for Sn [125]are popular choices as they are picked from the same cation groups. As shown in Figure 9, Promising progress has been made in this direction both experimentally and theoretically, demonstrating several remarkable efficiencies and mechanism of the defect emerging with cation substitution. Another possible solution to tackle the cationic disorder and related band tailing, as well as to activate the shallower defect is deliberate control of the synthesis condition, for example, post annealing the CZTS absorber within the critical temperature range of 200250°C and reasonable time range of 14 hours will improve the CZTS lattice ordering and increase the band gap [126, 127]. Control of the precursor fabrication process such as the metal stack order, the valence states of the Sn source, has also been reported effective in obtaining less defective absorber and corresponding high-efficiency device [128, 129]. Modification of the sulfurization or selenization condition could also suppress the formation of detrimental intrinsic defects and defect clusters by creating a desirable local chemical environment [11].

Figure 9.

(a) The band diagrams of CZTSSe and Ag-graded (Cu1 − xAgx)2ZnSn(S,Se)4 (CAZTSSe) solar cells under AM 1.5G 100 mW cm2 illumination and the J–V curves of CAZTSSe solar cells with different Ag composition gradients from the CdS interface to the Mo substrate. (b)J–V curves of CZTS and champion Cu2ZnxCd1–xSnS4 (CZCTS) devices and absorption coefficients and Urbach energy (EU) obtained from PDS measurement. (c)J–V curves for the champion CZTSSe and CZTGeSSe (30% Ge) solar cells. Reproduced from [115, 124, 125] with permission from Royal Society of Chemistry, American Chemical Society and John Wiley and Sons.

In term of the interface recombination path, it is mainly caused by the unpassivated charge defects at the heterojunction and/or the undesirable band alignment between the CZTS absorber and conventional CdS buffer layer. The band alignment problem is more prominent in high band gap pure sulfide Cu2ZnSnS4 because of its higher conduction band position. A serious of alternative wide band gap buffer layer materials including ZnCdS [33], Zn(O,S) [130, 131] and ZnSnO [132, 133] have been reported effective in mitigating the band alignment issues. The charge defects at the interface can be reduced by introducing ultrathin layer that has better lattice match with CZTS [10, 81] or alloying Ag near the heterojunction interface to from intrinsic or weak n-type (Cu,Ag)2ZnSn(S,Se)4 [115, 134]. Dielectric layers have also been investigated for passivating the interface defects, thereby suppressing the interface recombination [111, 135].

4.2.2 Short-circuit current density (JSC)

While the VOC loss contributes dominant performance loss in kestrite solar cells, JSC also represents a significant limiting factor for efficiency increases. The JSC loss is mainly caused by two reasons: one is the light or photons that is reflected or absorbed by layers above the CZTS (such as the buffer or window layers as shown in Figure 8), which cannot generate electron–hole pairs. The other is the low carrier collection efficiency due to the short diffusion length or narrow depletion width. The first problem can be addressed by introducing antireflection coating and optimizing the optical designs of the top layers above CZTS to increase the fraction of incident light reaching CZTS. The solution has been proven feasible when applying wide band gap buffer layer [111, 133], reducing the thickness and roughness of the top layers [119]. The typical External Quantum Efficiency (EQE) curve of CZTS usually shows relative lack of long wavelength response [9], which might be related to the low carrier lifetime. This low lifetime could also be the results of high defect density in the absorber layer or could simply be a consequence of high recombination loss at the back contact or at the front interface. Therefore, the solutions for suppressing the formation of detrimental defects disused in VOC loss above are also good for the JSC improvement In addition, band gap grading has also been developed in CZTS to increase the EQE by controlling the [S]/[S + Se] ratio at the surface and back surface [136].

4.2.3 Fill factor (FF)

FF is another significant parameter deficit in kesterite solar cell, which is mainly limited by the series resistance (RS) in the device. Careful study has been conducted to identify the origin of the RS, which implies that nonohmic back contact contributes greatly [137]. The nonohmic back contact could be resulted from Schottky barrier formation and/or secondary phase formation at the CZTS/Mo back interface. It is especially severe in high band gap CZTS because of the concentration decrease. Furthermore, the presence of low band gap secondary phases in CZTS could act as shunting pathway, which limits the FF and is therefore undesirable in the device as well. Deliberate investigation and control of the chemical reaction at the back interface have been proposed to modify the back interface microstructure and improve the device performance including FF [138, 139]. Introducing barrier layer at the CZTS/Mo interface has also been attempted in improving the back interface quality [140, 141].

Advertisement

5. Conclusion

Kesterite CZTS material is an emerging and promising PV technology, which could have the opportunity to realize low-cost and high-volume thin film solar cell production. After undergoing a rapid development in last decades, the technology has attracted increasing attention and demonstrated high potential in high efficiency. Experimental and theoretical studies reveal that the quaternary material exits in kesterite structure, similar to the chalcogenide CuInS2. It has a narrow stable region regarding the chemical potential, which makes the formation of secondary phases very easy. The intrinsic defects characteristic in the CZTS is complex where the deep level antisites defects and defect cluster could be prevalent. These unique properties make it struggling in achieving high-efficiency kesterite solar cells. The sensitive defects environment causes undesirable band tailing, electrostatic potential fluctuations, etc., which become recombination path and limit the open-circuit voltage, as well as other PV performance. The formation of secondary phases can lead to high serious resistance, shunting path depends on the band gap of the phase, which will also limit the device performance. In addition, the heterojunction interface and the Mo/CZTS back interface could also contribute to the performance loss because of the unfavorable band alignment, unpassivated defects, or detrimental reaction during high-temperature annealing. The solutions and approaches for tackling these loss mechanisms have been reviewed at the end of the chapter. Finally, in order to build a successful kesterite CZTS technology, combining the advanced approaches summarized in this chapter with further exploring the material synthesis and device physics would pave a path for higher efficiency.

Advertisement

Acknowledgments

This work was supported by the National Key R&D Program of China (No. 2018YFE0203400), the Science and Technology Innovation Program of Hunan Province (No. 2020RC2005), Australian Renewable Energy Agency (ARENA, 2017/RND006), and International Postdoctoral Exchange Fellowship Program (YJ20200207). X.H. acknowledges Australian Research Council (ARC) Future Fellowship (FT190100756). K.S. acknowledges the support from Australian Government through the Australian Renewable Energy Agency (ARENA) and the Australian Centre of Advanced Photovoltaics (ACAP, Grant No.1-SRI001).

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Nakamura M, Yamaguchi K, Kimoto Y, Yasaki Y, Kato T, Sugimoto H. Cd-Free Cu(In,Ga)(Se,S)2 thin-film solar cell with record efficiency of 23.35%. IEEE Journal of Photovoltaics. 2019;9(6):1863-1867. DOI: 10.1109/JPHOTOV.2019.2937218
  2. 2. Green MA, Dunlop ED, Hohl-Ebinger J, Yoshita M, Kopidakis N, Hao X. Solar cell efficiency tables (Version 58). Progress in Photovoltaics: Research and Applications. 2021;29(7):657-667. DOI: 10.1002/pip.3444
  3. 3. Commission E. Restriction of Hazardous Substances in Electrical and Electronic Equipment (RoHS). Strasbourg: European Commission; 2011 Available from: https://ec.europa.eu/environment/topics/waste-and-recycling/rohs-directive_en#ecl-inpage-622
  4. 4. Commission E. Critical Raw Materials. Brussels: European Commission; 2011 Available from: https://ec.europa.eu/growth/sectors/raw-materials/areas-specific-interest/critical-raw-materials_en
  5. 5. Walsh A, Chen S, Wei S-H, Gong X-G. Kesterite thin-film solar cells: Advances in materials modelling of Cu2ZnSnS4. Advanced Energy Materials. 2012;2(4):400-409. DOI: 10.1002/aenm.201100630
  6. 6. Zhou H, Hsu W-C, Duan H-S, Bob B, Yang W, Song T-B, et al. CZTS nanocrystals: A promising approach for next generation thin film photovoltaics. Energy & Environmental Science. 2013;6(10):2822-2838. DOI: 10.1039/C3EE41627E
  7. 7. Ito K. An overview of CZTS-based thin-film solar cells. In: Copper Zinc Tin Sulfide-Based Thin-Film Solar Cells. West Sussex, United Kingdom: John Wiley & Sons, Ltd.; 2014. pp. 1-41
  8. 8. Katagiri H, Jimbo K, Maw WS, Oishi K, Yamazaki M, Araki H, et al. Development of CZTS-based thin film solar cells. Thin Solid Films. 2009;517(7):2455-2460. DOI: 10.1016/j.tsf.2008.11.002
  9. 9. Mitzi DB, Gunawan O, Todorov TK, Wang K, Guha S. The path towards a high-performance solution-processed kesterite solar cell. Solar Energy Materials and Solar Cells. 2011;95(6):1421-1436. DOI: 10.1016/j.solmat.2010.11.028
  10. 10. Yan C, Huang J, Sun K, Johnston S, Zhang Y, Sun H, et al. Cu2ZnSnS4 solar cells with over 10% power conversion efficiency enabled by heterojunction heat treatment. Nature Energy. 2018;3(9):764-772. DOI: 10.1038/s41560-018-0206-0
  11. 11. Li J, Huang Y, Huang J, Liang G, Zhang Y, Rey G, et al. Defect control for 12.5% efficiency Cu2ZnSnSe4 Kesterite thin-film solar cells by Engineering of local chemical environment. Advanced Materials. 2020;32(52):2005268. DOI: 10.1002/adma.202005268
  12. 12. NREL. Best Research-Cell Efficiency Chart. Golden, United States: The National Renewable Energy Laboratory; 2021 Available from: https://www.nrel.gov/pv/cell-efficiency.html
  13. 13. Nitsche R, Sargent DF, Wild P. Crystal growth of quaternary 122464 chalcogenides by iodine vapor transport. Journal of Crystal Growth. 1967;1(1):52-53. DOI: 10.1016/0022-0248(67)90009-7
  14. 14. Ito K, Nakazawa T. Electrical and optical properties of Stannite-type quaternary semiconductor thin films. Japanese Journal of Applied Physics. 1988;27(Part 1, No 11):2094-2097.DOI: 10.1143/jjap.27.2094
  15. 15. Ito K, Nakazawa T. Stannite-type photovoltaic thin films. In: Proceedings of the 4th International Conference of Photovoltaic Science and Engineering. Sydney, NSW Australia: Institution of Radio and Electronics Engineers Australia; 1989
  16. 16. Katagiri H, Sasaguchi N, Hando S, Hoshino S, Ohashi J, Yokota T. Preparation and evaluation of Cu2ZnSnS4 thin films by sulfurization of E-B evaporated precursor. In: Technical Digest of the 9th International Conference of Photovoltaic Science and Engineering. Miyazaki, Japan: PVSEC-9; 1996
  17. 17. Katagiri H, Sasaguchi N, Hando S, Hoshino S, Ohashi J, Yokota T. Preparation and evaluation of Cu2ZnSnS4 thin films by sulfurization of E·B evaporated precursors. Solar Energy Materials and Solar Cells. 1997;49(1):407-414. DOI: 10.1016/S0927-0248(97)00119-0
  18. 18. Friedlmeier TM, Wieser N, Walter T, Dittrich H, Schock H. Heterojunctions based on Cu2ZnSnS4 and Cu2ZnSnSe4 thin films. In: Proceedings of the 14th European Conference of Photovoltaic Science and Engineering and Exhibition. Barcelona, Spain: European Photovoltaic Solar Energy Conference; 1997
  19. 19. Katagiri H. Cu2ZnSnS4 thin film solar cells. Thin Solid Films. 2005;480-481:426-432. DOI: 10.1016/j.tsf.2004.11.024
  20. 20. Katagiri H, Ishigaki N, Ishida T, Saito K. Characterization of Cu2ZnSnS4 thin films prepared by vapor phase sulfurization. Japanese Journal of Applied Physics. 2001;40(Part 1, No. 2A):500-504.DOI: 10.1143/jjap.40.500
  21. 21. Katagiri H, Saitoh K, Washio T, Shinohara H, Kurumadani T, Miyajima S. Development of thin film solar cell based on Cu2ZnSnS4 thin films. Solar Energy Materials and Solar Cells. 2001;65(1):141-148. DOI: 10.1016/S0927-0248(00)00088-X
  22. 22. Katagiri H, Jimbo K, Moriya K, Tsuchida K, editors. Solar cell without environmental pollution by using CZTS thin film. 3rd World Conference on Photovoltaic Energy Conversion. Osaka, Japan: IEEE; 2003
  23. 23. Kobayashi T, Jimbo K, Tsuchida K, Shinoda S, Oyanagi T, Katagiri H. Investigation of Cu2ZnSnS4-based thin film solar cells using abundant materials. Japanese Journal of Applied Physics. 2005;44(1B):783-787. DOI: 10.1143/jjap.44.783
  24. 24. Jimbo K, Kimura R, Kamimura T, Yamada S, Maw WS, Araki H, et al. Cu2ZnSnS4-type thin film solar cells using abundant materials. Thin Solid Films. 2007;515(15):5997-5999. DOI: 10.1016/j.tsf.2006.12.103
  25. 25. Katagiri H, Jimbo K, Yamada S, Kamimura T, Maw WS, Fukano T, et al. Enhanced conversion efficiencies of Cu2ZnSnS4-based thin film solar cells by using preferential etching technique. Applied Physics Express. 2008;1:041201. DOI: 10.1143/apex.1.041201
  26. 26. Tajima S, Itoh T, Hazama H, Ohishi K, Asahi R. Improvement of the open-circuit voltage of Cu2ZnSnS4 cells using a two-layered process. In: IEEE 40th Photovoltaic Specialist Conference (PVSC). Denver, CO, USA: IEEE; 2014. DOI: 10.1109/PVSC.2014.6924951
  27. 27. Todorov TK, Reuter KB, Mitzi DB. High-efficiency solar cell with Earth-abundant liquid-processed absorber. Advanced Materials. 2010;22(20):E156-E1E9. DOI: 10.1002/adma.200904155
  28. 28. Repins I, Beall C, Vora N, DeHart C, Kuciauskas D, Dippo P, et al. Co-evaporated Cu2ZnSnSe4 films and devices. Solar Energy Materials and Solar Cells. 2012;101:154-159. DOI: 10.1016/j.solmat.2012.01.008
  29. 29. Sugimoto H, Liao C, Hiroi H, Sakai N, Kato T. Lifetime improvement for high efficiency Cu2ZnSnS4 submodules. In: IEEE 39th Photovoltaic Specialists Conference (PVSC). Tampa, FL, USA: IEEE; 2013. DOI: 10.1109/PVSC.2013.6745135
  30. 30. Haass SG, Diethelm M, Werner M, Bissig B, Romanyuk YE, Tiwari AN. 11.2% Efficient solution processed Kesterite solar cell with a low voltage deficit. Advanced Energy Materials. 2015;5(18):1500712(1-7). DOI: 10.1002/aenm.201500712
  31. 31. Lin X, Kavalakkatt J, Lux-Steiner MC, Ennaoui A. Inkjet-printed Cu2ZnSn(S, Se)4 solar cells. Advanced Science. 2015;2:1500028(1-6). DOI: 10.1002/advs.201500028
  32. 32. Schnabel T, Abzieher T, Friedlmeier TM, Ahlswede E. Solution-based preparation of Cu2ZnSn(S,Se)4 for solar cells—Comparison of SnSe2 and elemental Se as Chalcogen source. IEEE Journal of Photovoltaics. 2015;5(2):670-675. DOI: 10.1109/JPHOTOV.2015.2392935
  33. 33. Sun K, Yan C, Liu F, Huang J, Zhou F, Stride JA, et al. Over 9% efficient Kesterite Cu2ZnSnS4 solar cell fabricated by using Zn1–xCdxS buffer layer. Advanced Energy Materials. 2016;6(12):1600046. DOI: 10.1002/aenm.201600046
  34. 34. Barkhouse DAR, Gunawan O, Gokmen T, Todorov TK, Mitzi DB. Device characteristics of a 10.1% hydrazine-processed Cu2ZnSn(Se,S)4 solar cell. Progress in Photovoltaics: Research and Applications. 2012;20(1):6-11. DOI: 10.1002/pip.1160
  35. 35. Qijie G, Yanyan C, Caspar JV, Farneth WE, Ionkin AS, Johnson LK, et al. A simple solution-based route to high-efficiency CZTSSe thin-film solar cells. In: Photovoltaic Specialists Conference (PVSC), 2012 38th IEEE. Austin, TX, USA: IEEE; 2012. DOI: 10.1109/PVSC.2012.6318213
  36. 36. Yang W, Duan H-S, Bob B, Zhou H, Lei B, Chung C-H, et al. Novel solution processing of high-efficiency Earth-abundant Cu2ZnSn(S,Se)4 solar cells. Advanced Materials. 2012;24(47):6323-6329. DOI: 10.1002/adma.201201785
  37. 37. Xin H, Katahara JK, Braly IL, Hillhouse HW. 8% Efficient Cu2ZnSn(S,Se)4 solar cells from redox equilibrated simple precursors in DMSO. Advanced Energy Materials. 2014;4(11):1301823 (1-5). DOI: 10.1002/aenm.201301823
  38. 38. Shin B, Gunawan O, Zhu Y, Bojarczuk NA, Chey SJ, Guha S. Thin film solar cell with 8.4% power conversion efficiency using an earth-abundant Cu2ZnSnS4 absorber. Progress in Photovoltaics: Research and Applications. 2013;21(1):72-76. DOI: 10.1002/pip.1174
  39. 39. Wang W, Winkler MT, Gunawan O, Gokmen T, Todorov TK, Zhu Y, et al. Device characteristics of CZTSSe thin-film solar cells with 12.6% efficiency. Advanced Energy Materials. 2014;4(7): 1301465 (1-5). DOI: 10.1002/aenm.201301465
  40. 40. Ahmed S, Reuter KB, Gunawan O, Guo L, Romankiw LT, Deligianni H. A high efficiency electrodeposited Cu2ZnSnS4 solar cell. Advanced Energy Materials. 2012;2(2):253-259. DOI: 10.1002/aenm.201100526
  41. 41. Todorov TK, Tang J, Bag S, Gunawan O, Gokmen T, Zhu Y, et al. Beyond 11% efficiency: Characteristics of state-of-the-art Cu2ZnSn(S,Se)4 solar cells. Advanced Energy Materials. 2012;2013:34-38. DOI: 10.1002/aenm.201200348
  42. 42. Yang K-J, Son D-H, Sung S-J, Sim J-H, Kim Y-I, Park S-N, et al. A band-gap-graded CZTSSe solar cell with 12.3% efficiency. Journal of Materials Chemistry A. 2016;4(26):10151-10158. DOI: 10.1039/C6TA01558A
  43. 43. Son D-H, Kim S-H, Kim S-Y, Kim Y-I, Sim J-H, Park S-N, et al. Effect of solid-H2S gas reactions on CZTSSe thin film growth and photovoltaic properties of a 12.62% efficiency device. Journal of Materials Chemistry A. 2019;7(44):25279-25289. DOI: 10.1039/C9TA08310C
  44. 44. Su Z, Liang G, Fan P, Luo J, Zheng Z, Xie Z, et al. Device postannealing enabling over 12% efficient solution-processed Cu2ZnSnS4 solar cells with Cd2+ substitution. Advanced Materials. 2020;32(32):2000121. DOI: 10.1002/adma.202000121
  45. 45. Du Y, Wang S, Tian Q , Zhao Y, Chang X, Xiao H, et al. Defect Engineering in Earth-abundant Cu2ZnSn(S,Se)4 photovoltaic materials via Ga3+-doping for over 12% efficient solar cells. Advanced Functional Materials. 2021;31(16):2010325. DOI: 10.1002/adfm.202010325
  46. 46. Gong Y, Qiu R, Niu C, Fu J, Jedlicka E, Giridharagopal R, et al. Ag Incorporation with controlled grain growth enables 12.5% efficient Kesterite solar cell with open circuit voltage reached 64.2% Shockley–Queisser limit. Advanced Functional Materials. 2021;31(24):2101927. DOI: 10.1002/adfm.202101927
  47. 47. Chen S, Gong XG, Walsh A, Wei S-H. Electronic structure and stability of quaternary chalcogenide semiconductors derived from cation cross-substitution of II-VI and I-III-VI compounds. Physical Review B. 2009;79(16):165211. DOI: 10.1103/PhysRevB.79.165211
  48. 48. Chen S, Gong XG, Walsh A, Wei S-H. Crystal and electronic band structure of Cu2ZnSnX4 (X=S and Se) photovoltaic absorbers: First-principles insights. Applied Physics Letters. 2009;94(4):041903. DOI: 10.1063/1.3074499
  49. 49. Hall S, Szymanski J, Stewart J. Kesterite, (Cu< 2)(Zn, Fe) SnS< 4), and stannite, Cu< 2)(Fe, Zn) SnS< 4), structurally similar but distinct minerals. The Canadian Mineralogist. 1978;16(2):131-137
  50. 50. Schorr S, Hoebler H-J, Tovar M. A neutron diffraction study of the stannite-kesterite solid solution series. European Journal of Mineralogy. 2007;19(1):65-73. DOI: 10.1127/0935-1221/2007/0019-0065
  51. 51. Friedelmeier T, Dittrich H, Schock H-W. Growth and characterization of Cu2ZnSnS4 and Cu2ZnSnSe4 thin films for photovoltaic applications. In: Ternary and Multinary Compounds. University of Salford, UK: CRC Press; 1998. p. 345
  52. 52. Bernardini GP, Borrini D, Caneschi A, Di Benedetto F, Gatteschi D, Ristori S, et al. EPR and SQUID magnetometry study of Cu2FeSnS4 (stannite) and Cu2ZnSnS4 (kesterite). Physics and Chemistry of Minerals. 2000;27(7):453-461. DOI: 10.1007/s002690000086
  53. 53. Schorr S. The crystal structure of kesterite type compounds: A neutron and X-ray diffraction study. Solar Energy Materials and Solar Cells. 2011;95(6):1482-1488. DOI: 10.1016/j.solmat.2011.01.002
  54. 54. Hall S, Kissin S, Stewart J. Stannite and kesterite-distinct minerals or components of a solid-solution. In: Acta crystallographica section A. Copenhagen, Denmark: Munksgaard Int Publ Ltd.; Vol. 31. 1975 p. s67
  55. 55. Seol J-S, Lee S-Y, Lee J-C, Nam H-D, Kim K-H. Electrical and optical properties of Cu2ZnSnS4 thin films prepared by rf magnetron sputtering process. Solar Energy Materials and Solar Cells. 2003;75(1):155-162. DOI: 10.1016/S0927-0248(02)00127-7
  56. 56. Babu GS, Kumar YBK, Bhaskar PU, Raja VS. Effect of post-deposition annealing on the growth of Cu2ZnSnSe4 thin films for a solar cell absorber layer. Semiconductor Science and Technology. 2008;23(8):085023. DOI: 10.1088/0268-1242/23/8/085023
  57. 57. Chen S, Walsh A, Luo Y, Yang J-H, Gong XG, Wei S-H. Wurtzite-derived polytypes of kesterite and stannite quaternary chalcogenide semiconductors. Physical Review B. 2010;82(19):195203. DOI: 10.1103/PhysRevB.82.195203
  58. 58. Paier J, Asahi R, Nagoya A, Kresse G. Cu2ZnSnS4 as a potential photovoltaic material: A hybrid Hartree-Fock density functional theory study. Physical Review B. 2009;79(11):115126. DOI: 10.1103/PhysRevB.79.115126
  59. 59. Persson C. Electronic and optical properties of Cu2ZnSnS4 and Cu2ZnSnSe4. Journal of Applied Physics. 2010;107(5):053710. DOI: 10.1063/1.3318468
  60. 60. Schorr S. Structural aspects of adamantine like multinary chalcogenides. Thin Solid Films. 2007;515(15):5985-5991. DOI: 10.1016/j.tsf.2006.12.100
  61. 61. Botti S, Kammerlander D, Marques MAL. Band structures of Cu2ZnSnS4 and Cu2ZnSnSe4 from many-body methods. Applied Physics Letters. 2011;98(24):241915. DOI: 10.1063/1.3600060
  62. 62. Ichimura M, Nakashima Y. Analysis of atomic and electronic structures of Cu2ZnSnS4 based on first-principle calculation. Japanese Journal of Applied Physics. 2009;48(9):090202. DOI: 10.1143/jjap.48.090202
  63. 63. Persson C, Chen R, Zhao H, Kumar M, Huang D. Electronic Structure and Optical Properties from First-Principles Modeling. In: Copper Zinc Tin Sulfide‐Based Thin‐Film Solar Cells. West Sussex, United Kingdom: John Wiley & Sons, Ltd.; 2014. pp. 75-105
  64. 64. Nakayama N, Ito K. Sprayed films of stannite Cu2ZnSnS4. Applied Surface Science. 1996;92:171-175. DOI: 10.1016/0169-4332(95)00225-1
  65. 65. Kamoun N, Bouzouita H, Rezig B. Fabrication and characterization of Cu2ZnSnS4 thin films deposited by spray pyrolysis technique. Thin Solid Films. 2007;515(15):5949-5952. DOI: 10.1016/j.tsf.2006.12.144
  66. 66. Gao F, Yamazoe S, Maeda T, Nakanishi K, Wada T. Structural and optical properties of In-free Cu2ZnSn(S,Se)4 solar cell materials. Japanese Journal of Applied Physics. 2012;51:10NC29. DOI: 10.1143/jjap.51.10nc29
  67. 67. Tanaka K, Fukui Y, Moritake N, Uchiki H. Chemical composition dependence of morphological and optical properties of Cu2ZnSnS4 thin films deposited by sol–gel sulfurization and Cu2ZnSnS4 thin film solar cell efficiency. Solar Energy Materials and Solar Cells. 2011;95(3):838-842. DOI: 10.1016/j.solmat.2010.10.031
  68. 68. Patel M, Mukhopadhyay I, Ray A. Structural, optical and electrical properties of spray-deposited CZTS thin films under a non-equilibrium growth condition. Journal of Physics D: Applied Physics. 2012;45(44):445103. DOI: 10.1088/0022-3727/45/44/445103
  69. 69. Ahn S, Jung S, Gwak J, Cho A, Shin K, Yoon K, et al. Determination of band gap energy (Eg) of Cu2ZnSnSe4 thin films: On the discrepancies of reported band gap values. Applied Physics Letters. 2010;97(2):021905. DOI: 10.1063/1.3457172
  70. 70. Haight R, Barkhouse A, Gunawan O, Shin B, Copel M, Hopstaken M, et al. Band alignment at the Cu2ZnSn(SxSe1−x)4/CdS interface. Applied Physics Letters. 2011;98(25):253502. DOI: 10.1063/1.3600776
  71. 71. He J, Sun L, Chen S, Chen Y, Yang P, Chu J. Composition dependence of structure and optical properties of Cu2ZnSn(S,Se)4 solid solutions: An experimental study. Journal of Alloys and Compounds. 2012;511(1):129-132. DOI: 10.1016/j.jallcom.2011.08.099
  72. 72. Ennaoui A, Lux-Steiner M, Weber A, Abou-Ras D, Kötschau I, Schock HW, et al. Cu2ZnSnS4 thin film solar cells from electroplated precursors: Novel low-cost perspective. Thin Solid Films. 2009;517(7):2511-2514. DOI: 10.1016/j.tsf.2008.11.061
  73. 73. Kumar M, Zhao H, Persson C. Cation vacancies in the alloy compounds of Cu2ZnSn(S1−xSex)4 and CuIn(S1−xSex)2. Thin Solid Films. 2013;535:318-321. DOI: 10.1016/j.tsf.2012.11.063
  74. 74. Chen S, Walsh A, Gong X-G, Wei S-H. Classification of lattice defects in the Kesterite Cu2ZnSnS4 and Cu2ZnSnSe4 Earth-abundant solar cell absorbers. Advanced Materials. 2013;25(11):1522-1539. DOI: 10.1002/adma.201203146
  75. 75. Lund EA, Du H, Hlaing Oo WM, Teeter G, Scarpulla MA. Investigation of combinatorial coevaporated thin film Cu2ZnSnS4 (II): Beneficial cation arrangement in Cu-rich growth. Journal of Applied Physics. 2014;115(17):173503. DOI: 10.1063/1.4871665
  76. 76. Chen S, Gong XG, Walsh A, Wei S-H. Defect physics of the kesterite thin-film solar cell absorber Cu2ZnSnS4. Applied Physics Letters. 2010;96(2):021902. DOI: 10.1063/1.3275796
  77. 77. Olekseyuk ID, Dudchak IV, Piskach LV. Phase equilibria in the Cu2S–ZnS–SnS2 system. Journal of Alloys and Compounds. 2004;368(1):135-143. DOI: 10.1016/j.jallcom.2003.08.084
  78. 78. Dudchak IV, Piskach LV. Phase equilibria in the Cu2SnSe3–SnSe2–ZnSe system. Journal of Alloys and Compounds. 2003;351(1):145-150. DOI: 10.1016/S0925-8388(02)01024-1
  79. 79. Chen S, Yang J-H, Gong XG, Walsh A, Wei S-H. Intrinsic point defects and complexes in the quaternary kesterite semiconductor Cu2ZnSnS4. Physical Review B. 2010;81(24):245204. DOI: 10.1103/PhysRevB.81.245204
  80. 80. Siebentritt S, Schorr S. Kesterites—a challenging material for solar cells. Progress in Photovoltaics: Research and Applications. 2012;20(5):512-519. DOI: 10.1002/pip.2156
  81. 81. Sun K, Huang J, Yan C, Pu A, Liu F, Sun H, et al. Self-assembled nanometer-scale ZnS structure at the CZTS/ZnCdS heterointerface for high-efficiency wide band gap Cu2ZnSnS4 solar cells. Chemistry of Materials. 2018;30(12):4008-4016. DOI: 10.1021/acs.chemmater.8b00009
  82. 82. Berg DM, Djemour R, Gütay L, Zoppi G, Siebentritt S, Dale PJ. Thin film solar cells based on the ternary compound Cu2SnS3. Thin Solid Films. 2012;520(19):6291-6294. DOI: 10.1016/j.tsf.2012.05.085
  83. 83. Marcano G, Rincón C, de Chalbaud LM, Bracho DB, Pérez GS. Crystal growth and structure, electrical, and optical characterization of the semiconductor Cu2SnSe3. Journal of Applied Physics. 2001;90(4):1847-1853. DOI: 10.1063/1.1383984
  84. 84. Yu P, Cardona M. Fundamentals of Semiconductors. 4th ed. Berlin, Heidelberg, Germany: Springer; 2010
  85. 85. Lin Y-T, Shi J-B, Chen Y-C, Chen C-J, Wu P-F. Synthesis and characterization of Tin Disulfide (SnS2) nanowires. Nanoscale Research Letters. 2009;4(7):694. DOI: 10.1007/s11671-009-9299-5
  86. 86. Sava F, Lörinczi A, Popescu M, Socol G, Axente E, Mihailescu IN, et al. Amorphous SnSe2 films. Journal of Optoelectronics and Advanced Materials. 2006;8(4):1367-1371
  87. 87. Vidal J, Lany S, d’Avezac M, Zunger A, Zakutayev A, Francis J, et al. Band-structure, optical properties, and defect physics of the photovoltaic semiconductor SnS. Applied Physics Letters. 2012;100(3):032104. DOI: 10.1063/1.3675880
  88. 88. Sinsermsuksakul P, Heo J, Noh W, Hock AS, Gordon RG. Atomic layer deposition of Tin monosulfide thin films. Advanced Energy Materials. 2011;1(6):1116-1125. DOI: 10.1002/aenm.201100330
  89. 89. Franzman MA, Schlenker CW, Thompson ME, Brutchey RL. Solution-phase synthesis of SnSe nanocrystals for use in solar cells. Journal of the American Chemical Society. 2010;132(12):4060-4061. DOI: 10.1021/ja100249m
  90. 90. Liu G, Schulmeyer T, Brötz J, Klein A, Jaegermann W. Interface properties and band alignment of Cu2S/CdS thin film solar cells. Thin Solid Films. 2003;431-432:477-482. DOI: 10.1016/S0040-6090(03)00190-1
  91. 91. Kashida S, Shimosaka W, Mori M, Yoshimura D. Valence band photoemission study of the copper chalcogenide compounds, Cu2S, Cu2Se and Cu2Te. Journal of Physics and Chemistry of Solids. 2003;12:2357-2363. DOI: 10.1016/S0022-3697(03)00272-5
  92. 92. Siebentritt S. Why are kesterite solar cells not 20% efficient? Thin Solid Films. 2013;535:1-4. DOI: 10.1016/j.tsf.2012.12.089
  93. 93. Timo Wätjen J, Engman J, Edoff M, Platzer-Björkman C. Direct evidence of current blocking by ZnSe in Cu2ZnSnSe4 solar cells. Applied Physics Letters. 2012;100(17):173510. DOI: 10.1063/1.4706256
  94. 94. Redinger A, Mousel M, Wolter MH, Valle N, Siebentritt S. Influence of S/Se ratio on series resistance and on dominant recombination pathway in Cu2ZnSn(SSe)4 thin film solar cells. Thin Solid Films. 2013;535:291-295. DOI: 10.1016/j.tsf.2012.11.111
  95. 95. Mendis BG, Goodman MCJ, Major JD, Taylor AA, Durose K, Halliday DP. The role of secondary phase precipitation on grain boundary electrical activity in Cu2ZnSnS4 (CZTS) photovoltaic absorber layer material. Journal of Applied Physics. 2012;112(12):124508. DOI: 10.1063/1.4769738
  96. 96. Sun K, Yan C, Huang J, Liu F, Li J, Sun H, et al. Beyond 10% efficiency Cu2ZnSnS4 solar cells enabled by modifying the heterojunction interface chemistry. Journal of Materials Chemistry A. 2019;7(48):27289-27296. DOI: 10.1039/C9TA09576D
  97. 97. Redinger A, Siebentritt S. Coevaporation of Cu2ZnSnSe4 thin films. Applied Physics Letters. 2010;97(9):092111. DOI: 10.1063/1.3483760
  98. 98. Mousel M, Redinger A, Djemour R, Arasimowicz M, Valle N, Dale P, et al. HCl and Br2-MeOH etching of Cu2ZnSnSe4 polycrystalline absorbers. Thin Solid Films. 2013;535:83-87. DOI: 10.1016/j.tsf.2012.12.095
  99. 99. Nagoya A, Asahi R, Wahl R, Kresse G. Defect formation and phase stability of Cu2ZnSnS4 photovoltaic material. Physical Review B. 2010;81(11):113202. DOI: 10.1103/PhysRevB.81.113202
  100. 100. Wibowo RA, Kim WS, Lee ES, Munir B, Kim KH. Single step preparation of quaternary Cu2ZnSnSe4 thin films by RF magnetron sputtering from binary chalcogenide targets. Journal of Physics and Chemistry of Solids. 2007;68(10):1908-1913. DOI: 10.1016/j.jpcs.2007.05.022
  101. 101. Zhang X, Shi X, Ye W, Ma C, Wang C. Electrochemical deposition of quaternary Cu2ZnSnS4 thin films as potential solar cell material. Applied Physics A. 2009;94(2):381-386. DOI: 10.1007/s00339-008-4815-5
  102. 102. Tanaka T, Nagatomo T, Kawasaki D, Nishio M, Guo Q , Wakahara A, et al. Preparation of Cu2ZnSnS4 thin films by hybrid sputtering. Journal of Physics and Chemistry of Solids. 2005;66(11):1978-1981. DOI: 10.1016/j.jpcs.2005.09.037
  103. 103. Scragg JJ, Dale PJ, Peter LM, Zoppi G, Forbes I. New routes to sustainable photovoltaics: evaluation of Cu2ZnSnS4 as an alternative absorber material. Physica Status Solidi (b). 2008;245(9):1772-1778. DOI: 10.1002/pssb.200879539
  104. 104. Altosaar M, Raudoja J, Timmo K, Danilson M, Grossberg M, Krustok J, et al. Cu2Zn1–x Cdx Sn(Se1–y Sy)4 solid solutions as absorber materials for solar cells. Physica Status Solidi (a). 2008;205(1):167-170. DOI: 10.1002/pssa.200776839
  105. 105. Oishi K, Saito G, Ebina K, Nagahashi M, Jimbo K, Maw WS, et al. Growth of Cu2ZnSnS4 thin films on Si (100) substrates by multisource evaporation. Thin Solid Films. 2008;517(4):1449-1452. DOI: 10.1016/j.tsf.2008.09.056
  106. 106. Kishore Kumar YB, Suresh Babu G, Uday Bhaskar P, Sundara RV. Preparation and characterization of spray-deposited Cu2ZnSnS4 thin films. Solar Energy Materials and Solar Cells. 2009;93(8):1230-1237. DOI: 10.1016/j.solmat.2009.01.011
  107. 107. Hönes K, Zscherpel E, Scragg J, Siebentritt S. Shallow defects in Cu2ZnSnS4. Physica B: Condensed Matter. 2009;404(23):4949-4952. DOI: 10.1016/j.physb.2009.08.206
  108. 108. Shinde NM, Dubal DP, Dhawale DS, Lokhande CD, Kim JH, Moon JH. Room temperature novel chemical synthesis of Cu2ZnSnS4 (CZTS) absorbing layer for photovoltaic application. Materials Research Bulletin. 2012;47(2):302-307. DOI: 10.1016/j.materresbull.2011.11.020
  109. 109. Miyamoto Y, Tanaka K, Oonuki M, Moritake N, Uchiki H. Optical properties of Cu2ZnSnS4 thin films prepared by Sol–Gel and sulfurization method. Japanese Journal of Applied Physics. 2008;47(1):596-597. DOI: 10.1143/jjap.47.596
  110. 110. Prabhakar T, Jampana N. Effect of sodium diffusion on the structural and electrical properties of Cu2ZnSnS4 thin films. Solar Energy Materials and Solar Cells. 2011;95(3):1001-1004. DOI: 10.1016/j.solmat.2010.12.012
  111. 111. Cui X, Sun K, Huang J, Yun JS, Lee C-Y, Yan C, et al. Cd-Free Cu2ZnSnS4 solar cell with an efficiency greater than 10% enabled by Al2O3 passivation layers. Energy & Environmental Science. 2019;12(9):2751-2764. DOI: 10.1039/C9EE01726G
  112. 112. Cazzaniga A, Crovetto A, Yan C, Sun K, Hao X, Ramis Estelrich J, et al. Ultra-thin Cu2ZnSnS4 solar cell by pulsed laser deposition. Solar Energy Materials and Solar Cells. 2017;166:91-99. DOI: 10.1016/j.solmat.2017.03.002
  113. 113. Guo Q , Ford GM, Yang W-C, Walker BC, Stach EA, Hillhouse HW, et al. Fabrication of 7.2% Efficient CZTSSe Solar Cells Using CZTS Nanocrystals. Journal of the American Chemical Society. 2010;132(49):17384-17386. DOI: 10.1021/ja108427b
  114. 114. Cao Y, Denny MS, Caspar JV, Farneth WE, Guo Q , Ionkin AS, et al. High-efficiency solution-processed Cu2ZnSn(S,Se)4 thin-film solar cells prepared from binary and ternary nanoparticles. Journal of the American Chemical Society. 2012;134(38):15644-15647. DOI: 10.1021/ja3057985
  115. 115. Qi Y-F, Kou D-X, Zhou W-H, Zhou Z-J, Tian Q-W, Meng Y-N, et al. Engineering of interface band bending and defects elimination via a Ag-graded active layer for efficient (Cu,Ag)2ZnSn(S,Se)4 solar cells. Energy & Environmental Science. 2017;10(11):2401-2410. DOI: 10.1039/C7EE01405H
  116. 116. Wangperawong A, King JS, Herron SM, Tran BP, Pangan-Okimoto K, Bent SF. Aqueous bath process for deposition of Cu2ZnSnS4 photovoltaic absorbers. Thin Solid Films. 2011;519(8):2488-2492. DOI: 10.1016/j.tsf.2010.11.040
  117. 117. Sun K, Wang A, Su Z, Liu F, Hao X. Enhancing the performance of Cu2ZnSnS4 solar cell fabricated via successive ionic layer adsorption and reaction method by optimizing the annealing process. Solar Energy. 2021;220:204-210. DOI: 10.1016/j.solener.2021.03.033
  118. 118. Zhou Z, Wang Y, Xu D, Zhang Y. Fabrication of Cu2ZnSnS4 screen printed layers for solar cells. Solar Energy Materials and Solar Cells. 2010;94(12):2042-2045. DOI: 10.1016/j.solmat.2010.06.010
  119. 119. Winkler MT, Wang W, Gunawan O, Hovel HJ, Todorov TK, Mitzi DB. Optical designs that improve the efficiency of Cu2ZnSn(S,Se)4 solar cells. Energy & Environmental Science. 2014;7(3):1029-1036. DOI: 10.1039/C3EE42541J
  120. 120. Shockley W, Queisser HJ. Detailed balance limit of efficiency of p-n junction solar cells. Journal of Applied Physics. 1961;32(3):510-519. DOI: 10.1063/1.1736034
  121. 121. Gokmen T, Gunawan O, Todorov TK, Mitzi DB. Band tailing and efficiency limitation in kesterite solar cells. Applied Physics Letters. 2013;103(10):103506. DOI: 10.1063/1.4820250
  122. 122. Chirilă A, Buecheler S, Pianezzi F, Bloesch P, Gretener C, Uhl AR, et al. Highly efficient Cu(In,Ga)Se2 solar cells grown on flexible polymer films. Nature Materials. 2011;10(11):857-861. DOI: 10.1038/nmat3122
  123. 123. Su Z, Tan JMR, Li X, Zeng X, Batabyal SK, Wong LH. Cation substitution of solution-processed Cu2ZnSnS4 thin film solar cell with over 9% efficiency. Advanced Energy Materials. 2015;5(19):1500682. DOI: 10.1002/aenm.201500682
  124. 124. Yan C, Sun K, Huang J, Johnston S, Liu F, Veettil BP, et al. Beyond 11% efficient sulfide Kesterite Cu2ZnxCd1–xSnS4 solar cell: Effects of Cadmium alloying. ACS Energy Letters. 2017;2(4):930-936. DOI: 10.1021/acsenergylett.7b00129
  125. 125. Hages CJ, Levcenco S, Miskin CK, Alsmeier JH, Abou-Ras D, Wilks RG, et al. Improved performance of Ge-alloyed CZTGeSSe thin-film solar cells through control of elemental losses. Progress in Photovoltaics: Research and Applications. 2015;23(3):376-384. DOI: 10.1002/pip.2442
  126. 126. Rey G, Redinger A, Sendler J, Weiss TP, Thevenin M, Guennou M, et al. The band gap of Cu2ZnSnSe4: Effect of order-disorder. Applied Physics Letters. 2014;105(11):112106. DOI: 10.1063/1.4896315
  127. 127. Scragg JJS, Choubrac L, Lafond A, Ericson T, Platzer-Björkman C. A low-temperature order-disorder transition in Cu2ZnSnS4 thin films. Applied Physics Letters. 2014;104(4):041911. DOI: 10.1063/1.4863685
  128. 128. Gong Y, Zhang Y, Zhu Q , Zhou Y, Qiu R, Niu C, et al. Identifying the origin of the Voc deficit of kesterite solar cells from the two grain growth mechanisms induced by Sn2+ and Sn4+ precursors in DMSO solution. Energy & Environmental Science. 2021;14(4):2369-2380. DOI: 10.1039/D0EE03702H
  129. 129. Yang K-J, Kim S, Kim S-Y, Ahn K, Son D-H, Kim S-H, et al. Flexible Cu2ZnSn(S,Se)4 solar cells with over 10% efficiency and methods of enlarging the cell area. Nature Communications. 2019;10(1):2959. DOI: 10.1038/s41467-019-10890-x
  130. 130. Ericson T, Scragg JJ, Hultqvist A, Wätjen JT, Szaniawski P, Törndahl T, et al. Zn(O, S) buffer layers and thickness variations of CdS buffer for Cu2ZnSnS4 solar cells. IEEE Journal of Photovoltaics. 2014;4(1):465-469. DOI: 10.1109/JPHOTOV.2013.2283058
  131. 131. Li J, Huang L, Hou J, Wu X, Niu J, Chen G, et al. Effects of substrate orientation and solution movement in chemical bath deposition on Zn(O,S) buffer layer and Cu(In,Ga)Se2 thin film solar cells. Nano Energy. 2019;58:427-436. DOI: 10.1016/j.nanoen.2019.01.054
  132. 132. Cui X, Sun K, Huang J, Lee C-Y, Yan C, Sun H, et al. Enhanced heterojunction interface quality to achieve 9.3% efficient Cd-free Cu2ZnSnS4 solar cells using atomic layer deposition ZnSnO buffer layer. Chemistry of Materials. 2018;30(21):7860-7871. DOI: 10.1021/acs.chemmater.8b03398
  133. 133. Larsen JK, Larsson F, Törndahl T, Saini N, Riekehr L, Ren Y, et al. Cadmium free Cu2ZnSnS4 solar cells with 9.7% efficiency. Advanced Energy Materials. 2019;9(21):1900439. DOI: 10.1002/aenm.201900439
  134. 134. Yuan Z-K, Chen S, Xiang H, Gong X-G, Walsh A, Park J-S, et al. Engineering solar cell absorbers by exploring the band alignment and defect disparity: The case of Cu- and Ag-based Kesterite compounds. Advanced Functional Materials. 2015;25(43):6733-6743. DOI: 10.1002/adfm.201502272
  135. 135. Sun H, Sun K, Huang J, Yan C, Liu F, Park J, et al. Efficiency enhancement of Kesterite Cu2ZnSnS4 solar cells via solution-processed ultrathin Tin oxide intermediate layer at absorber/buffer interface. ACS Applied Energy Materials. 2018;1(1):154-160. DOI: 10.1021/acsaem.7b00044
  136. 136. Huang TJ, Yin X, Qi G, Gong H. CZTS-based materials and interfaces and their effects on the performance of thin film solar cells. Physica status solidi (RRL) - Rapid Research Letters. 2014;08(09):735-762. DOI: 10.1002/pssr.201409219
  137. 137. Tai KF, Gunawan O, Kuwahara M, Chen S, Mhaisalkar SG, Huan CHA, et al. Fill factor losses in Cu2ZnSn(SxSe1−x)4 solar cells: Insights from physical and electrical characterization of devices and exfoliated films. Advanced Energy Materials. 2016;6(3):1501609. DOI: 10.1002/aenm.201501609
  138. 138. Scragg JJ, Wätjen JT, Edoff M, Ericson T, Kubart T, Platzer-Björkman C. A detrimental reaction at the molybdenum back contact in Cu2ZnSn(S,Se)4 thin-film solar cells. Journal of the American Chemical Society. 2012;134(47):19330-19333. DOI: 10.1021/ja308862n
  139. 139. Scragg JJ, Kubart T, Wätjen JT, Ericson T, Linnarsson MK, Platzer-Björkman C. Effects of back contact instability on Cu2ZnSnS4 devices and processes. Chemistry of Materials. 2013;25(15):3162-3171. DOI: 10.1021/cm4015223
  140. 140. Liu F, Sun K, Li W, Yan C, Cui H, Jiang L, et al. Enhancing the Cu2ZnSnS4 solar cell efficiency by back contact modification: Inserting a thin TiB2 intermediate layer at Cu2ZnSnS4/Mo interface. Applied Physics Letters. 2014;104(5):051105. DOI: 10.1063/1.4863736
  141. 141. Liu F, Huang J, Sun K, Yan C, Shen Y, Park J, et al. Beyond 8% ultrathin kesterite Cu2ZnSnS4 solar cells by interface reaction route controlling and self-organized nanopattern at the back contact. NPG Asia Materials. 2017;9(7):e401. DOI: 10.1038/am.2017.103

Written By

Kaiwen Sun, Fangyang Liu and Xiaojing Hao

Submitted: 11 November 2021 Reviewed: 24 November 2021 Published: 31 December 2021