The structural and electronic properties of Cu(In1−xBx)Se2 as a new photovoltaic material

Yulan Cheng*ab, Kexiang Weia, Ping Xiaa and Quan Baia
aDepartment of Mechanical Engineering, Hunan Institute of Engineering, Xiangtan 411101, China. E-mail: cyl-zjf@163.com
bHunan Province Cooperative Innovation Center for Wind Power Equipment and Energy Conversion, Xiangtan 411101, China

Received 8th July 2015 , Accepted 19th September 2015

First published on 23rd September 2015


Abstract

The newly synthesized solar cell absorber material Cu(In1−xBx)Se2 based on a solvothermal method with a special solvent is investigated using density functional theory. Our results show that Cu(In1−xBx)Se2 has an extremely high formation enthalpy which indicates that it is hard to synthesize using a traditional evaporation method. The band gap of Cu(In1−xBx)Se2 increases with increasing B content with a larger bowing parameter than that of Cu(In1−xGax)Se2. The band alignment of CuInSe2 and CuBSe2 is of type II. And the upshift of the valence band maximum is much smaller than that of the conduction band maximum for CuBSe2, which indicates that pure CuBSe2 can easily be p-type doped but that n-type doping could be more difficult in the alloy with a high B content. Therefore, we expect that the high efficiency Cu(In1−xGax)Se2 solar cell material has a low B content.


I. Introduction

Ternary CuInSe2 possesses a zinc-blend derived chalcopyrite crystal structure and an excellent optical absorption coefficient, and is widely studied as an absorber material in solar cells. The band gap of CuInSe2 is about 1.04 eV, which is smaller than the optimal single-junction value (1.5 eV) according to the Shockley–Queisser model.1 However, the band gap of CuInSe2 can be tuned over a large range through alloying CuInSe2 with its sister semiconductors Cu–III–VI2 (III = Al, Ga, VI = S and Se), since these compounds can form various solid solutions. Among these solid solutions, one outstanding material is Cu(In1−xGax)Se2 (CIGS) which possesses the world record of optical to electrical efficiency in thin film solar cells.2 Changing the Ga content can tune the band gap from 1.04 to 1.68 eV. However, Ga is an expensive raw material, which limits the large scale production of CIGS solar cells. Therefore, it is necessary to explore alternative alloys to CuGaSe2 which can tune the band gap of CuInSe2 to close to 1.5 eV and also have the excellent optical properties of CuInSe2.

Apart from Al and Ga, B also lies in the same column as In in the periodic table. However, there is seldom research about the Cu(In1−xBx)Se2 alloy, since, until recently, the Cu(In1−xBx)Se2 alloy was synthesized through a solvothermal based method.3 In this paper, the authors also created Cu(In1−xBx)Se2 solar cell devices with an efficiency of 2.34%. A theoretical understanding of CuBSe2 as well as the Cu(In1−xBx)Se2 alloy is still lacking. DFT calculations enable us to predict the electronic structure properties of both CuBSe2 and the Cu(In1−xBx)Se2 alloy. We can expect that the band gap of CuBSe2 will be larger than that of CuGaSe2 and CuAlSe2, and a low concentration of B rather than Ga will tune the band gap of CuInSe2 to the ideal value of 1.5 eV. In this study, the miscibility and the compositional dependence of the band gap change as well as band alignments are determined using first-principles calculations.

The electronic structure and the compositional dependence of the physical properties of Cu(In1−xBx)Se2 are studied based on the special quasi-random structure (SQS). We find out that Cu(In1−xBx)Se2 has an extremely high formation enthalpy, which indicates that Cu(In1−xBx)Se2 is hard to synthesize using a traditional evaporation method. The band gap of Cu(In1−xBx)Se2 increases with increasing B content and the band gap bowing parameter is larger than that of Cu(In1−xGax)Se2 because of the larger band gap of CuBSe2 than CuGaSe2. The band alignment of CuInSe2 and CuBSe2 shows that the band offset mainly results from the conduction band maximum (CBM) upshift of CuBSe2, which indicates that n-type doping could be more difficult in the alloy with a high B content. Therefore, the high efficiency Cu(In1−xBx)Se2 solar cells have a low B content compared with that of Cu(In1−xGax)Se2.

II. Computational details

The electronic structure and total energy calculations are performed using density functional theory as implemented in the plane wave VASP code.4 The generalized gradient approximation of the Perdew–Burke–Ernzerhof (PBE-GGA) exchange–correlation functional is used.5,6 The interaction between the core electrons and the valence electrons is included using the standard frozen-core projector augmented-wave PAW potentials.7,8 The cutoff energy for the plane wave basis is set to 300 eV. For the Brillouin zone integration, we used the 4 × 4 × 4 k-point mesh for 64-atom cells.

The valence band offset of ΔEv(CuInSe2/CuBSe2) can be defined as:10,11

 
image file: c5ra13379c-t1.tif(1)

Here,

 
image file: c5ra13379c-t2.tif(2)
 
image file: c5ra13379c-t3.tif(3)
where image file: c5ra13379c-t4.tif and image file: c5ra13379c-t5.tif are the core level to valence band maximum energy separations for CuInSe2 and CuBSe2, respectively. ΔEC,C′ is the core level binding energy difference between CuInSe2 and CuBSe2, which is derived from the calculation of the 1 × 1 (001) oriented (CuInSe2)n/(CuBSe2)n superlattice. We have tested that n = 2 can keep the core level of the innermost layer on each side of the superlattice bulk-like. In our calculations, all the structure parameters of the superlattice are fully relaxed. The conduction band offset ΔEc is defined as:
 
ΔEc = ΔEg + ΔEv (4)
where ΔEg is the experimental band gap difference between CuInSe2 and CuBSe2.

Several methods have been developed to investigate random alloys in the past, for example, virtual crystal approximation (VCA)12 and cluster expansion (CE).13–15 However, VCA is non-structural and CE is time consuming. The SQS method is an alternative approach to investigate a random alloy, and constructs a finite supercell to best match the correlation functions of a standard random alloy.16,17 Therefore, the SQS is a finite supercell, which includes the microstructure and can be adopted for first-principles calculations. The SQS approach is widely used in the study of solid solution alloys with great success.18–20

In this work, we use the SQS to construct the Cu(In1−xBx)Se2 64-atom supercell random alloy with the lattice vectors:

 
[a with combining right harpoon above (vector)]1 = (2, 0, 0)a, [a with combining right harpoon above (vector)]2 = (0, 2, 0)a, [a with combining right harpoon above (vector)]3 = (0, 0, 2η)a (5)
where η = c/2a, and a and c are the short and long lattice constants, respectively, of the tetragonal chalcopyrite cell. The atomic positions for the SQS at x = 0.25 and x = 0.5 are listed in Table 1, and their structural correlation functions [capital Pi, Greek, macron]k,m as well as ideal random alloy correlation functions are listed in Table 2, which shows that the constructed SQSs in this paper are reasonably good.

Table 1 The atomic coordinates and the occupation of the SQS for In and B in Cu(In1−xBx)Se2 at the concentrations x = 0.25 and x = 0.5 used in our work. Only the mixed sublattice coordinates are shown for clarity
Type x = 0.25 Type x = 0.5
Coordinates Coordinates
B 0.00 0.25 0.75 B 0.00 0.25 0.75
B 0.50 0.75 0.75 B 0.50 0.25 0.75
B 0.25 0.25 0.00 B 0.50 0.75 0.75
B 0.25 0.75 0.00 B 0.75 0.50 0.25
In 0.00 0.75 0.75 B 0.25 0.25 0.00
In 0.50 0.25 0.75 B 0.25 0.75 0.00
In 0.25 0.00 0.25 B 0.75 0.25 0.00
In 0.25 0.50 0.25 B 0.00 0.50 0.50
In 0.75 0.00 0.25 In 0.00 0.75 0.75
In 0.75 0.50 0.25 In 0.25 0.00 0.25
In 0.75 0.25 0.00 In 0.25 0.50 0.25
In 0.75 0.75 0.00 In 0.75 0.00 0.25
In 0.00 0.00 0.50 In 0.75 0.75 0.00
In 0.00 0.50 0.50 In 0.00 0.00 0.50
In 0.50 0.00 0.50 In 0.50 0.00 0.50
In 0.50 0.50 0.50 In 0.50 0.50 0.50


Table 2 Atomic correlation functions [capital Pi, Greek, macron]k,m of the SQS used in our calculation at concentrations x = 0.25 and 0.5, and compared with the ideal values (2x − 1)k of the random alloy
  [capital Pi, Greek, macron]2,1 [capital Pi, Greek, macron]2,2 [capital Pi, Greek, macron]2,3 [capital Pi, Greek, macron]2,4 [capital Pi, Greek, macron]3,1 [capital Pi, Greek, macron]4,1
x = 0.25
SQS 1/4 1/4 1/4 0 0 0
Random 1/4 1/4 1/4 1/4 −1/8 1/16
[thin space (1/6-em)]
x = 0.5
SQS 0 0 −1/8 −1/4 −1/4 −1/16
Random 0 0 0 0 0 0


III. Results and discussion

In the chalcopyrite CuInSe2, each Se anion coordinates two Cu and two In, while each cation (Cu or In) is tetrahedrally coordinated by four Se atoms. The computed and experimental lattice parameters21,22 and the anion displacement for CuInSe2 and CuBSe2 are given in Table 3. For CuInSe2 our theoretical structure parameters are in good agreement with the experimental ones. For CuBSe2, there is no experimental data about its structure and electronic structure. To see the electronic structure of CuBSe2 clearly, we plotted the total and partial density of states (DOS) for the 16-atom CuBSe2 as shown in Fig. 1. From the partial DOS, we can see that the upper valence band is derived mainly from the hybridization of the Se 4p and Cu 3d states, similar to those of CuInSe2 and CuGaSe2.22 The low conduction band derives mainly from the hybridization of the Se 4s, 4p, 3d and B 2s, 2p states. The calculated band gaps for CuBSe2 and CuInSe2 are 1.506 eV and 0.02 eV, respectively, which are smaller compared with the experimental values since the GGA functional cannot well describe the hybridization of the Cu 3d states and Se 4p states and therefore GGA investigations usually underestimate the band gap of semiconductors. However, the general chemical trend of the band gap variation is reproduced in the GGA calculations, i.e., the band gap of CuBSe2 is larger than that of CuInSe2. Much work has shown that the hybrid functional (HSE06 functional) describes the localized Cu 3d orbitals more correctly than (semi)local-density functionals23–25 by substituting part of the short range PBE-GGA exchange energy with the short range Hartree–Fock (HF) exchange energy. We also use the HSE06 functional to evaluate the band gaps of CuInSe2 and CuBSe2, which yield the values of 1.0 eV and 3.14 eV.
Table 3 The comparison of the crystal constants a, η = c/2a, and u and the band gaps from the experimental values and GGA optimized results for both CuInSe2 and CuBSe2
  CuInSe2 CuBSe2
Exp./GGA GGA
a (Å) 5.782/5.865 5.337
η 1.004/1.006 0.911
u 0.224/0.219 0.202
Gap (eV) 1.04/0.02 1.506



image file: c5ra13379c-f1.tif
Fig. 1 The total and partial density of states of CuBSe2.

The formation enthalpy describes the miscibility of the alloys. For CuIn1−xBxSe2 the formation enthalpy is defined as:

 
ΔH(x) = E(x) − (1 − x) ECuInSe2xECuBSe2 (6)
where ECuInSe2 and ECuBSe2 represent the total energy of pure CuInSe2 and CuBSe2, and E(x) is the total energy of the alloy at a composition x. The relationship between the alloy formation enthalpy and its composition x often obeys a quadratic function, i.e.,
 
ΔH(x) = (1 − xH(0) + xΔH(1) + Ωx(1 − x) (7)
where Ω is the interaction parameter indicating the solubility of the alloy. The calculated formation enthalpies of the Cu(In1−xBx)Se2 alloy at five different compositions, namely x = 0, 0.25, 0.5, 0.75 and 1, are shown in Fig. 2 with blue triangles. The positive formation enthalpies indicate that the alloy prefers phase segregation into CuInSe2 and CuBSe2 at zero temperature, and mixing In and B cations costs additional energy to form the random alloy. By using eqn (7) to fit the calculated formation enthalpies, we get the interaction parameter Ω = 927 meV per atom. This value is much larger than that of Cu(In1−xGax)Se2 (11 meV per atom).26 The extremely large interaction parameter suggests that thousands of Kelvin are needed to transform the two compounds into an alloy, which is mainly caused by the large chemical mismatch between the In and B atoms and the inert nature of the B element. As a result, Cu(In1−xBx)Se2 can hardly be synthesized using a traditional evaporation method. This is also the reason why, until recently, the Cu(In1−xBx)Se2 alloy was only synthesized using a solvothermal-based method with a special solvent. The band gap of CuBSe2 is larger than that of CuGaSe2, so less B than Ga is needed to tune CuInSe2 to the optimized band gap value.


image file: c5ra13379c-f2.tif
Fig. 2 The calculated formation enthalpy for the Cu(In1−xBx)Se2 alloy as a function of the alloy composition x. The fitting curve according to eqn (7) is also given with the interaction parameter Ω.

Since the Cu(In1−xBx)Se2 alloy is already synthesized, we will show the band gap changes depending on the composition. Usually, the dependence of the band gap on the composition can be described using the following equation for random semiconductor alloys:

 
Eg(x) = (1 − x)Eg(CuInSe2) + xEg(CuBSe2) − bx(1 − x) (8)
where Eg denotes the band gap, and b is the bowing parameter. The band gaps for the Cu(In1−xBx)Se2 alloy at x = 0, 0.25, 0.5, 0.75 and 1 are calculated using the PBE functional and the fitted bowing parameter b ∼1.68 eV is obtained. It is true that the PBE functional often underestimates the band gaps of semiconductors, but the calculated bowing parameters are accurate because the bowing parameters depend on the band gap difference, so the band gap errors induced by the PBE functional are systematically canceled in the calculation. The results shown in Fig. 3 are the shifted band gaps using the scissor operator, so that the band gap at x = 0 is equal to the experimental value of about 1.04 eV and the band gap at x = 1 is equal to the HSE06 band gap value of about 3.14 eV. After correction, the band gaps at different compositions can be compared directly with the experimental values.


image file: c5ra13379c-f3.tif
Fig. 3 The band gap of the Cu(In1−xBx)Se2 alloy as a function of the composition x. The band gap shifted using the scissor operator so the band gap at x = 0 agrees with the experimental value and at x = 1 is equal to the HSE06 calculated value.

From Fig. 3 we can see the monotonically increasing nature of the band gap of the Cu(In1−xBx)Se2 alloy with the composition parameter x. The bowing parameter of Cu(In1−xBx)Se2 is larger than that of the Cu(In1−xGax)Se2 alloy (0.1–0.3 eV)9,10, which is caused by the larger size and chemical mismatch between In and B compared to In and Ga. However, the absolute bowing parameter value is large, indicating that the Cu(In1−xBx)Se2 alloy has poor tolerance to the chemical and size difference of the mixed cations. Because the band gap of CuBSe2 is larger than that of CuInSe2, only 30% of B is needed to alloy into CuInSe2 to tune the Cu(In1−xBx)Se2 band gap to about 1.5 eV. For comparison, 30% Ga can tune the band gap of CuIn1−xGaxSe2 to a smaller value of about 1.1 eV.

To understand how the band gap increases with increasing B composition in CuIn1−xBxSe2, we calculate the band offsets of CuInSe2 and CuBSe2. The valence band maximum (VBM) offsets of the two compounds are small (the offset is less than 0.02 eV), but the CBM shifts up significantly from CuInSe2 to CuBSe2, as shown in Fig. 4. As a result, the CBM offset is 2.1 eV between CuInSe2 and CuBSe2. The small valence band offset is related to the states of the VBM of these compounds which primarily consists of the Se 4p and Cu 3d orbitals; therefore, the substitution of In with Ga or B does not change the VBM states and energies significantly. However, the CBM mainly consists of the antibonding state of the hybridization between the group III (In, B) s and p orbitals and Se 4s, 4p and 3d orbitals. Because the In–Se bond length is larger than the B–Se bond length, the repulsion between In and Se is smaller than between B and Se; as a result, the antibonding CBM states of CuBSe2 are pushed to a higher energy level than those of CuInSe2. Similarly, the CBM energy of CuGaSe2 is lower than that of CuBSe2.


image file: c5ra13379c-f4.tif
Fig. 4 The band alignments of CuInSe2, CuGaSe2 and CuBSe2 in eV.

The calculated band alignments in Fig. 4 show that the band alignment is of type II. The band gap of CuGaSe2 and CuBSe2 increasing with the Ga and B composition is mainly contributed by the conduction band upshift, with a much smaller contribution from the valence band downshift. With this finding, the electrical conductivity of these alloys can be predicted. According to the doping limit rule,27 a material is difficult to be n-type doped if the CBM energy is high and difficult to be p-type doped if the VBM energy is low. The experiments and theoretical calculations have shown that the CuIn1−xGaxSe2 samples have an intrinsic p-type conductivity but are hard to be n-type doped with high Ga concentrations, due to the compensation of acceptor defects.28–30 The small valence band offset indicates that the Cu(In1−xBx)Se2 alloy has an intrinsic p-type conductivity, and the much higher CBM of CuBSe2 compared to that of CuInSe2 suggests that n-type doping could be more difficult in a Cu(In1−xBx)Se2 alloy as the band gap increases.

IV. Conclusion

Using first-principles calculations, we have investigated the electronic properties of CuBSe2 as well as the Cu(In1−xBx)Se2 random alloy as a function of the composition x. Our results show that Cu(In1−xBx)Se2 has an extremely high formation enthalpy which indicates that it is hard to synthesize using a traditional evaporation method. The band gap increases monotonically with increasing B content, and the bowing parameter is larger for Cu(In1−xBx)Se2 than that for Cu(In1−xGax)Se2 due to the larger size and chemical mismatch between In–B and In–Ga. The band gap increases with B content, which primarily comes from the large conduction band upshift. This indicates that Cu(In1−xBx)Se2 shows an intrinsic p-type conductivity similar to CuInSe2 and their n-type doping could be difficult as the B content increases. Based on these results, we predict that high efficiency Cu(In1−xBx)Se2 solar cells should have low B content.

Acknowledgements

This work was supported by the Hunan Provincial joint fund for young talent training project under Contract No. 2015JJ6022 and the National Nature Science Foundation of China under Contract No. 11472103.

References

  1. W. Shockley and H. J. Queisser, J. Appl. Phys., 1961, 32, 510 CrossRef CAS PubMed.
  2. P. Jackson, D. Hariskos, R. Wuerz, O. Kiowski, A. Bauer, T. Friellmeier and M. Powalla, Phys. Status Solidi, 2015, 9(1), 28–31 CAS.
  3. L. Chen, J. Liao, Y. Chuang and Y. Fu, J. Am. Chem. Soc., 2011, 133, 3704 CrossRef CAS PubMed.
  4. G. Kresse and J. Furthmüller, Comput. Mater. Sci., 1996, 6, 15 CrossRef CAS.
  5. Y. Kumagai, Y. Soda, F. Oba and A. Seko, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 033203 CrossRef.
  6. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS.
  7. P. E. Blöchl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953 CrossRef.
  8. G. Kresse and D. Joubert, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 1758 CrossRef CAS.
  9. T. Tinoco, C. Rincón, M. Quintero and G. S. Pérez, Phys. Status Solidi A, 1991, 124, 427 CrossRef CAS PubMed.
  10. S.-H. Wei and A. Zunger, J. Appl. Phys., 1995, 78, 3846 CrossRef CAS PubMed.
  11. S.-H. Wei and A. Zunger, Appl. Phys. Lett., 1998, 72, 2011 CrossRef CAS PubMed.
  12. L. Nordheim, Ann. Phys., 1931, 9, 607 CrossRef CAS PubMed.
  13. J. M. Sanchez, F. Ducastelle and D. Gratias, Phys. A, 1984, 128, 334 CrossRef.
  14. D. de Fontaine, Solid State Phys., 1994, 47, 33 Search PubMed.
  15. F. Ducastelle, Order and Phase Stability in Alloys, Elsevier Science, New York, 1994 Search PubMed.
  16. A. Zunger, S.-H. Wei, L. G. Ferreira and J. E. Bernard, Phys. Rev. Lett., 1990, 65, 353 CrossRef CAS.
  17. S.-H. Wei, L. G. Ferreira, J. E. Bernard and A. Zunger, Phys. Rev. B: Condens. Matter Mater. Phys., 1990, 42, 9622 CrossRef CAS.
  18. S. Chen, X. G. Gong and S.-H. Wei, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 75, 205209 CrossRef.
  19. S. Chen, A. Walsh, J.-H. Yang, X. G. Gong, L. Sun, P.-X. Yang, J.-H. Chu and S.-H. Wei, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 83, 125201 CrossRef.
  20. Q. Shu, J.-H. Yang, S. Chen, B. Huang, H. Xiang, X. G. Gong and S.-H. Wei, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 87, 205209 Search PubMed.
  21. H. W. Spiess, U. Haeberlen, G. Brandt, A. Rauber and J. Schneider, Phys. Status Solidi, 1974, 8(62), 183 CrossRef PubMed.
  22. J. E. Jaffe and A. Zunger, Phys. Rev. B: Condens. Matter Mater. Phys., 1983, 28, 5822 CrossRef CAS.
  23. L. E. Oikkonen, M. G. Ganchenkova, A. P. Seitsonen and R. M. Nieminen, J. Phys.: Condens. Matter, 2011, 23, 422202 CrossRef CAS PubMed.
  24. L. E. Oikkonen, M. G. Ganchenkova, A. P. Seitsonen and R. M. Nieminen, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 86, 165115 CrossRef.
  25. J. Pohl and K. Albe, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 87, 245203 CrossRef.
  26. S.-H. Wei and A. Zunger, J. Appl. Phys., 1995, 78, 3846 CrossRef CAS PubMed.
  27. S. B. Zhang, S.-H. Wei and A. Zunger, J. Appl. Phys., 1998, 83, 3192 CrossRef CAS PubMed.
  28. C. Persson, Y.-J. Zhao, S. Lany and A. Zunger, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 72, 035211 CrossRef.
  29. S.-H. Wei and S. B. Zhang, J. Phys. Chem. Solids, 2005, 66, 1994 CrossRef CAS PubMed.
  30. S. Lany and A. Zunger, Phys. Rev. Lett., 2008, 100, 016401 CrossRef.

Footnote

PACS numbers: 71.20.Nr, 71.23.k, 73.61.Le.

This journal is © The Royal Society of Chemistry 2015