Study of the Growth and Dislocation Blocking Mechanisms in InxGa1−xAs Buffer Layer for Growing High-Quality In0.5Ga0.5P, In0.3Ga0.7As, and In0.52Ga0.48As on Misoriented GaAs Substrate for Inverted Metamorphic Multijunction Solar Cell Application
Abstract
Effects of growth conditions and buffer structures on crystal quality of 1.9-eV In0.5G0.5P, 1-eV In0.3Ga0.7As, and 0.75-eV In0.52Ga0.48As materials on misoriented GaAs substrate for inverted metamorphic solar cell grown by metalorganic chemical vapor deposition have been studied. Large lattice mismatch issue between the two lower bandgap 1.0-eV In0.3Ga0.7As and 0.75-eV In0.52Ga0.48As epilayers and the substrate has been resolved by using optimized step-graded buffer layers. Threading dislocation blocking mechanisms have been studied and discussed. It was indicated that threading dislocations have been significantly blocked in the designed InGaAs buffer layers through annihilation reactions between threading dislocations or through the formation of misfit dislocations. As a result, smooth surface In0.3Ga0.7As and In0.52Ga0.48As epifilms with threading dislocation density of about 1 × 106 cm−2 were obtained. For the growth of InGaP, the surface morphology, crystal quality, ordering parameter, and InGaP composition were significantly affected by growth temperature, V/III ratio, and III/III ratio. The almost lattice match to the substrate, high crystal quality, and smooth surface of In0.495Ga0.505P were obtained at growth temperature of 675°C and V/III ratio above 150. The ordering parameter was strongly dependent on growth temperature. Photoluminescence measurement indicated that the bandgap of the InGaP epilayer is 1.89 eV, indicating that the solid composition in the InGaP film is more disordered.
1 Introduction
In the past year, researchers have reported record-breaking conventional 3-junction solar cells with conversion efficiencies of up to 41.6% (King et al. 2012), and scientists are now looking at overcoming the next performance barrier of 50% (Leite et al. 2013). Despite the high efficiency, it is well known that the 3-junction cell bandgap combination compromises the ideal solar spectrum splitting in favor of subcell material quality, hence limiting the achievable efficiency. Current match for such cells is a major constraint for further improvement of efficiencies. Future terrestrial cells will likely feature four or more junctions with performance potential capable of reaching over 50% efficiency at concentration (Law et al. 2010). The 4-, 5-, or 6-junction concentrator cells trade lower current densities for higher voltage and divide the solar spectrum more efficiently. For example, theoretical calculations indicate an ideal efficiency of over 59% for a 4-junction cell with bandgap combination of 1.9/1.42/1.02/0.75 eV (Law et al. 2010). In these designs, III/V materials which have energy bandgap about 1.0–1.05 and 0.7–0.75 eV play very important roles (Sherif et al. 2005). InGaAs seems to be the best choice of material for this range of bandgap. However, the 2.4% and 4% lattice mismatch between 1.0-eV In0.3Ga0.7As and 0.75-eV In0.52Ga0.48As and GaAs substrates, respectively, imposes difficulty to grow high-quality material. It is thus more challenging and even more difficult to grow a high bandgap subcell above it. In recent years, researchers have tried to apply advance technologies including metamorphic growth on Ge (King et al. 2007), novel lattice-matched designs (Green et al. 2013), inverted metamorphic (IMM) growth (Geisz et al. 2007), or wafer bonding technique (Bett et al. 2013) to increase the efficiency of solar cell. Table 1 summarizes recent results of III/V solar cell performance and material growth for solar cell application. Although 4-junction solar cell using wafer bonding technique has achieved new record efficiency of 44.7% (AM1.5D, 297 suns), the structure, however, consists of two dual junctions GaInP/GaAs and GaInAsP/GaInAs grown on two expensive GaAs and InP substrates, which may result in high production cost (Bett et al. 2013). IMM growth solution has been developed to overcome the current matching and lattice-mismatch issues in multijunction cells, and 40.8% (AM1.5D, 326 suns) efficiency of 3J-IMM In0.49Ga0.51P/In0.04Ga0.96As/In0.37Ga0.63As has been reported (Geisz et al. 2008). Similar to metamorphic growth, a graded InGaAs buffer layer is utilized to change the lattice constant of subsequent growth. However, in this method the solar cell is grown upside down with the top cell InGaP, lattice matched to the substrate grown first, followed by the middle cell that is either GaAs or mismatched GaInAs grown lattice matched to the substrate. The bottom (most highly lattice mismatched) cells are grown last. By this approach, the top cells will be maintained as defect free, while limiting the defects from the buffer region into the lower cells which are not as easily affected as the top subcell. Additionally, Tatavari et al. have developed a full-wafer epitaxial liftoff technology that allows to separate the grown structure and expensive substrate (Tatavarti et al. 2009). This advanced technology not only allows to reduce cell cost due to the possibility of substrate reuse multiple times but also to lower the total weight of IMM multijunction cells. A schematic 4-junction IMM solar cell structure is illustrated in Figure 1(a). Recently, an efficiency of 34.5% (AM0, 1 sun) for a 4-junction IMM has been achieved (Patel et al. 2012). However, the efficiency of the 4-junction IMM is still lower than prediction. The lower performance was likely attributed to high dislocation density, which causes degradation of the performance of device, in the subcells caused by large lattice mismatch between the subcells and substrate. Therefore, it is necessary to find out an effective method to block dislocations from propagating to active layers in such lattice-mismatched InGaAs/GaAs material for future multijunction IMM solar cell application.

(color online). (a) Schematic 4-junction IMM solar cell. (b) Designed InxGa1−xAs buffer profiles using 6-, 8-, and 10-SG buffer layer for growing In0.3Ga0.7As on GaAs substrate
Summary of material growth and device performance of III/V compound solar cells
Reference | Material growth | Solar cell performance | |||||
Structure | Growth technique | Relaxation degree (%) | Surface roughness (nm) | TDD (cm−2) | V oc (V) | Efficiency (%) | |
SpectroLab (King et al. 2007) | In0.44Ga0.56P/In0.08Ga0.92As/Ge | MM-MOCVD | 100 | – | – | 2.911 | 40.7-AM1.5D under 240 suns |
NREL (Green et al. 2013) | InGaP/GaAs/InGaNAs(Sb) | LM-MBE | 99 | – | – | – | 43.5-AM1.5D under 360 suns |
EMCORE (Patel et al. 2012) | In0.5Ga0.5P/GaAs/In0.3Ga0.7As/In0.6Ga0.4As | IMM-MOCVD | 100 | – | 5 × 106 | 3.236 | 34.5-AM0 under 1 sun |
Fraunhofer (Bett et al. 2013) | 2J (GaInP/GaAs) + 2J (GaInAsP/GaInAs) | MOCVD | – | – | – | 4.165 | 44.7-AM1.5D under 297.3 suns |
N. J. Quitoriano (Quitoriano and Fitzgerald 2007) | In0.54Ga0.46As/GaAs | MM-MOCVD | 99 | 15 | 3.1 × 107 | – | – |
K. E. Lee (Lee and Fitzgerald 2010) | In0.42Ga0.58As/GaAs | MM-MOCVD | 86.9 | 4.8 | 3.6 × 107 | – | – |
G. B. Galiev (Galiev et al. 2014) | In0.64Ga0.36As/GaAs | MM-MBE | 95 | 6 | 5 × 106 | – | – |
In previous work, the effects of different substrate misorientation degrees and growth temperature on material properties of the n++-GaAs/p++-AlGaAs tunneling diode and crystal quality of the In0.3Ga0.7As layer have been investigated. It is found that the misorientation influences both surface roughness and interface properties of the n++-GaAs/p++-AlGaAs tunneling diode. The use of 10° offcut GaAs substrates resulted in reducing oxygen contamination in n++GaAs and p++-AlGaAs layers due to the reduction of sticking coefficient and number of anisotropic sites (Yu et al. 2010). Moreover, the effect of GaAs substrate misorientation on surface morphology and crystal quality of 1-eV In0.3Ga0.7As was affected by the growth mode by changes in surface free energy and also by the formation, gliding, and multiplication of dislocations (Nguyen et al. 2011). The work indicated that a better crystal quality In0.3Ga0.7As layer with threading dislocation (TD) density as low as 1 × 106 cm−2 has been achieved on a 6o offcut substrate using a 10-step-grade (SG) buffer layer. However, the buffer layer thickness of 1.5 μm seems to be quite thick for IMM solar cell application.
This paper reports more detailed experimental results on In0.5Ga0.5P and 1.0- and 0.75-eV InxGa1−xAs for 4 or more junction IMM solar cell application. For InGaP, the experiment aims at optimization of growth conditions for achieving high crystal quality, smooth surface InGaP, which provides better conditions for growing subcells on top. For InGaAs subcell growth, TD blocking mechanisms and buffer structure are studied in detail for reducing TD density in the subcells to about 106 cm−2.
2 Experiment methods
In the experiment, InGaP and InGaAs samples were grown on epi-ready GaAs (001) substrates with 6° offcut toward [110] direction. Metalorganic chemical vapor deposition (MOCVD-EMCORE D180) was used to grow the epilayers. Group-III precursors of trimethylindium and trimethylgallium and group-V precursor of pure arsine (AsH3) and phosphine (PH3) were used. The indium composition and degree of relaxation were determined with a high-resolution X-ray diffractometer (HR-XRD). The surface texture and roughness were examined by atomic force microscopy (AFM). The dislocation densities were characterized by cross-sectional and plan-view transmission electron microscopy (TEM). The bandgap of InGaP was determined using photoluminescence (PL) performed at room temperature.
3 Results and discussion
3.1 The growth of In0.5Ga0.5P cell on GaAs substrate
Due to high cracking temperature of PH3 source, InGaP compound normally is grown at high temperature. Moreover, difference in bond dissociation energy between InP (46.8 kcal/mol) and GaP (58.4 kcal/mol) (Seki, Watanabe, and Matsui 1978), and As and P intermixing may form unnecessary phase at interface between InGaP and GaAs and degrade InGaP quality. Therefore, any change in growth conditions such as growth temperature, V/III ratio, partial pressure, and III(In)/III(Ga) may result in different InGaP composition and its crystal quality. Our experimental results show that better crystal quality and smoother surface In0.5Ga0.5P epifilm was obtained at growth temperature of 675°C and high V/III ratio conditions. As shown in Figure 2(a), the best surface roughness of 0.8 nm was obtained on the sample grown at a temperature of 675°C and V/III ratio of above 150. With the smooth surface InGaP having been achieved, we believe that the further growth of 1.42-eV GaAs mid-cell and other components of IMM structure on top of InGaP will not be affected. The sample also exhibited good interface between InGaP and GaAs substrate as illustrated in Figure 2(b). Lower or higher growth temperature causes worse crystal quality and rougher surface roughness. The reason is that the PH3 cracking efficiency was too low at lower growth temperature causing insufficient phosphorus for InGaP formation. On the other hand, higher growth temperature will increase the desorption rate from surface for In atoms compared to Ga atoms (Hageman et al. 1992). The lack of In during the solidification process will, therefore, affect the crystal quality due to larger lattice mismatch between the epifilm and substrate.

(color online). (a) AFM image, (b) cross-sectional TEM, (c) (004) ω–2θ scan XRD, and (d) PL spectra measured at room temperature of In0.5Ga0.5P grown on GaAs at 675°C and V/III ratio of 150–180
InGaP is lattice matched to GaAs at the Ga content of 51%. A varied composition would cause a tensile or compressive stress that degrades the crystal quality of the InGaP epilayer. Figure 2(c) shows (004) ω–2θ XRD pattern of InGaP epilayer grown on GaAs substrate at 675°C using two different III(In)/III(Ga) ratios. It is seen that almost lattice match and smooth surface InGaP on GaAs substrate can be attained in the range of III(In)/III(Ga) ratio of 1.69–1.81 at the same growth condition. Any change in III(In)/III(Ga) ratio far from this range would result in higher mismatch that causes an increase in the surface roughness of the epilayer. Figure 2(d) shows the PL spectra of the In0.495Ga0.505As layer measured at room temperature. The layer exhibited PL main peak at 656 nm with the full width half maximum value of 17 meV. The bandgap of the epifilm, therefore, was derived from PL peak position to be 1.89 eV. The ordering parameter of 0.48 calculated based on the method developed by Murata, Ho, and Stringfellow (1997) suggested that Ga and In atoms in the InGaP film tend to arrange more randomly. However, it was found that the degree of ordering parameter was increased with increasing growth temperature, suggesting that the bandgap of In0.5Ga0.5P could be tuned just by tuning degree of ordering through changing growth temperature (Zunger and Mahajan 1994), V/III ratio, surfactant treatment (Mori and Fitzgerald 2009), or strain in the epifilm (Novak et al. 2002; Garcia et al. 2008).
3.2 TD blocked in optimized step-graded InxGa1−xAs buffer for growing high crystal quality 1-eV and 0.75-eV InGaAs
We first study the gliding and blocking process in InxGa1−xAs SG buffer layer for the purpose of growth high crystal quality 1.0-eV In0.3Ga0.7As on GaAs substrate. For the purpose of study, three samples were grown using three different buffer structures as shown schematically in Figure 1(b). In the three samples, buffer layers including 6, 8, and 10 individual SG InxGa1−xAs layers with different composition gradients and/or starting indium concentration values at the interface were grown first, before the fixed composition In0.3Ga0.7As epilayers were grown on the top. In the three buffer structures, the thickness of individual layer in the buffer layer was grown exceeding the critical thickness because it is believed that TD density can be reduced through reaction with other TDs or blocked by misfit dislocations (MDs) during the strain relaxation in metamorphic structure (Romanov et al. 1999). Modeling study (Speck et al. 1996) suggested that the use of multiple discrete strained layers, which were grown over their critical thickness, can provide a marked reduction in TD density through annihilation reactions and blocking process (Freund 1990). The critical thickness of individual layer was estimated based on Matthews-Blakeslee model (Matthews and Blakeslee 1974) as stated in eq. [1]:
Equation [1] can be simplified to hc≈|b|/εm because the product of all the other terms in the equation are approximately constant and have magnitude on the order of one, where |b| is the magnitude of Burgers vector b, and εm is misfit strain.
Figure 3(a)–3(c) shows the cross-sectional bright-field TEM images of the three samples grown using 6-, 8-, and 10-SG InxGa1−xAs buffer layers, respectively. Although some TDs have been blocked in the buffer layer for the 600-nm-thick 6-SG InGaAs grade buffer, some TDs, however, still propagate into the fixed composition epilayer and end at the free surface (see cross-sectional TEM micrograph in Figure 3(a)), resulting in higher TD density of about 7 × 106 cm−2, as indicated by plan-view TEM micrograph in Figure 5(b). TDs extended from the bottom and higher misfit strain in the 6-SG buffer layer have worsened the surface morphology of the top surface layer. AFM study (Figure 4(a)) shows the surface morphology with root mean square (RMS) roughness of 2.7 nm for the 6-SG InGaAs buffer. Interestingly, the 8-SG and 10-SG buffer layers, although greater in thickness (1 and 1.5 μm, respectively), showed no obvious TD in the fixed composition In0.3Ga0.7As layer as shown in the plan-view TEM image of Figure 3(b) and 3(c), respectively. It is clear that TDs were blocked and contained within the buffer layers, resulting in almost no TDs extension into the In0.3Ga0.7As epilayer. Better surface roughness was also observed on the two samples as shown in Figure 4(b) and 4(c). Surface roughness with RMS value of 1.8 and 1.5 nm for samples using 8-SG and 10-SG, respectively, were obtained. The lower TD densities in the epilayers were attributed to the lower misfit strain between individual buffer layer and more discrete strained layers in 8- and 10-SG buffer structure. It has been reported that in lattice-mismatched zinc blend crystal, with low misfit strain, strain relaxation occurs primarily by the formation of 60° a/2 〈110〉{111} misfit dislocations (Speck et al. 1996). At the end of every MD, TD segments are generated and often find their way into the active layer (Goldman et al. 1998). If TDs are glissile, then for [001] growth they will glide in one of four {111} slip planes. Thus, there are four unique {111} planes and six unique 〈110〉 directions. Considering that the dislocation Burgers vectors can assume either positive or negative sense, there are 12 possible Burgers vectors. Finally, a glissile dislocation with an a/2 〈110〉 Burgers vector can have its line in one of two possible {111} planes and thus there are a total of 24 specific dislocation Burgers vector/slip plane combinations (Speck et al. 1996). In all possible combinations between TDs, only certain TDs which have antiparallel Burgers vector can fall into annihilation reactions, the remaining combinations may result in new TD segments through fusion reactions and these new TD segments tend to move up to active layer. Therefore, the purpose of multiple discrete layers in metamorphic growth is to increase the probability of annihilation reactions between TDs. As can be seen in Figure 3(b) and 3(c), both annihilation and fusion reactions between TDs in the buffers were observed. Simultaneously, mobile TD blocked by MD process was also observed. Figure 5(a) presents a plan-view TEM image observed at two InxGa1−xAs buffer interface showing a number of dislocation intersection events. It is observed that when dislocation intersections occur, a variety of interactions can take place. As marked at points B and C, where two pairs of a/2 〈110〉 MDs fell into interaction, no blocking interaction happened, their TD arms still glided up on their {111} glide planes. However, at point A, the motion of TD arm was blocked. We believed that TD blocking processes observed from TEM images explain for better crystal quality in the In0.3Ga0.7As epifilms grown using 8- and 10-SG buffer layers. Figure 5(c) shows plan-view TEM image of In0.3Ga0.7As on 8-SG sample, only one TD appeared indicated by the circle marked in the figure, indicating that a TD density of about 1 × 106 cm−2 in the In0.3Ga0.7As epilayer has been obtained.

(a)–(c) Cross-sectional bright-field TEM images of In0.3Ga0.7As film grown on GaAs substrate using 6-, 8-, and 10-SG InxGa1−xAs buffer layers, respectively

(Color online). (a)–(c) AFM images of In0.3Ga0.7As epifilms grown on GaAs substrate using 6-, 8-, and 10-SG InxGa1−xAs buffer layers, respectively

(a) Plan-view TEM image observed at the interface between two InxGa1−xAs buffer layers in 8-SG buffer structure. (b) Plan-view TEM image of In0.3Ga0.7As epilayer on GaAs substrate using 6-SG, and (c) 8-SG buffer layer
After optimizing the thickness and composition profile of InxGa1−xAs buffer layer for growing high crystal quality 1-eV subcell, we continued growing a “quasi-structure” of 2-junction 1- and 0.75-eV subcell on GaAs substrate. The structure includes 3 μm In0.3Ga0.7As on 8-SG InxGa1−xAs (x=0.05–0.3) followed by 0.8 μm 6-SG InxGa1−xAs (x=0.35–0.5) buffer layer, concluded by the growth of final 1.5 μm In0.52Ga0.48As epilayer. Interestingly, the crystal quality of 1-eV In0.3Ga0.7As was not affected by the growth of subsequent InGaAs buffer layer and 0.75-eV In0.52Ga0.48As epilayer. As shown in Figure 6(a), no TD was observed in the In0.52Ga0.48As and In0.3G0.7aAs epilayers. The plan-view TEM image (Figure 6(b)) of the In0.52Ga0.48As epilayer shows only three TDs (appearing near stars masked in the figure) on a large area of 200 μm2, which corresponds to a TD density of 1.5 × 106 cm−2 in In0.52Ga0.48As. The growth of high indium composition also caused an increase in RMS roughness of the film to 2.4 nm (Figure 6(c)).

(Color online). (a) Cross-sectional TEM image, (b) plan-view TEM image, (c) AFM image, and (d) (115) ω–2θ scan XRD of In0.3Ga0.7As and In0.52Ga0.48As epilayers grown on GaAs substrate by optimizing the InxGa1−xAs buffer layers
Figure 6(d) represents the asymmetric scan for (115) reflection obtained from 1.5 and 3 μm In0.52Ga0.48As and In0.3Ga0.7As epilayers on GaAs substrate, respectively. Asymmetric rocking curve scans on (115) reflection of substrate were recorded in ω–2θ scans to determine indium composition and relaxation degree of InxGa1−xAs alloy on GaAs substrate. The indium composition (x) in InxGa1−xAs epilayer is obtained using Vegard’s law (Bowen and Tanner 2005)
where as = 5.6533 Å, aInAs = 6.0583 Å are lattice constants of GaAs substrate and InAs, respectively. af is bulk equivalent or unstrained lattice constant defined by
where ν is Poisson ratio, a// and a⊥ are the in-plane and out-of-plane lattice parameters of InxGa1−xAs epilayers, respectively, calculated from Bragg’s law reflection position by
where θB is Bragg angle, φ is the angle between the diffraction plane and the sample surface, h, k, and l are Miller indices of reflection plane, and λ is the wavelength of the X-rays (λ = 1.5406 Å). Degree of relaxation R is defined by
Bragg angles were determined from peak position of In0.3Ga0.7As and In0.52Ga0.48As epilayers, and the in-plane and out-of-plane lattice parameters were calculated to respective values a||=5.7625 (±0.001) a⊥=5.7836 (±0.001) Å for In0.3Ga0.7As, and a|| = 5.8651 (±0.001) and a⊥=5.8705 (±0.001) Å for In0.52Ga0.48As. From [2] and [6], the indium composition and relaxation degree of the epilayers were defined as ×=0.295±0.05 and R=91% for In0.3Ga0.7As and ×=0.516±0.05 and R=97% for In0.52Ga0.48As, respectively.
4 Conclusion
We found that surface morphology, crystal quality, and composition of InGaP sensitively depended on growth temperature, V/III, and III(In)/III(Ga) ratios due to high thermal decomposition temperature of PH3 source and due to the difference in bonding energy between In–P and Ga–P bonds. It is found that the bandgap of InGaP can be tuned not only by In composition but also by ordering parameter, which is dependent on growth conditions. However, the high crystal quality and smooth surface In0.5Ga0.5P for 1.9-eV top cell has been obtained by controlling growth conditions carefully. Besides, the effects of misfit gradient in metamorphic SG buffer layer have been investigated for the growth of 1- and 0.75-eV InGaAs-based subcells. By designing multiple discrete strained layers in metamorphic buffer structure, TDs were blocked effectively through annihilation reactions between TDs or TD blocked by MD mechanisms as predicted in theoretical models. The use of higher misfit gradient to design InxGa1−xAs metamorphic buffer results in less effective TD blocking, though the buffer layer thickness was reduced to about 600 nm. The buffer layer structure including 8-SG InGaAs with nonlinear indium gradient was found to be the optimal for blocking TDs more efficiently for growing 1-eV InGaAs on 1.4-eV GaAs middle cell, while the buffer layer thickness was maintained to be lower than 1 μm. The use of similar buffer structure to subsequently grow 0.75-eV InGaAs subcell was also impressive in positive achievements. TEM images confirmed that TDs as low as 1 × 106 cm−2 on 1- and 0.75-eV InGaAs active layers have been observed. We are going to combine results from this work and results from previous work to fabricate full 4-jucntion IMM solar cell structure in the next step. We also believe that metamorphic growth technique developed from this study can be applied for future 5- or 6-junction IMM solar cell.
Acknowledgement
This work was sponsored by the NCTU-UCB I-RiCE program, National Science Council, Taiwan, under Grant No. NSC 103-2911-I-009-302.
References
Bett, A. W. , S. P.Philipps, S.Essig, S.Heckelmann, R.Kellenbenz, V.Klinger, M.Niemeyer, D.Lackner, and F.Dimroth. 2013. “Overview about technology perspectives for high et1iciency solar cells for space and terrestrial applications.” 28th European Photovoltaic Solar Energy Conference, Iap.I.1.Suche in Google Scholar
Bowen, D. K. , and B. K.Tanner. 2005. High Resolution X-Ray Diffractometry and Topography. London: Taylor & Francis.Suche in Google Scholar
Freund, L. B. 1990. “A Criterion for Arrest of a Threading Dislocation in a Strained Epitaxial Layer Due To an Interface Misfit Dislocation in Its Path.” Journal of Applied Physics68:2073.Suche in Google Scholar
Galiev, G. B. , I. S.Vasilevskii, E. A.Klimov, S. S.Pushkarev, A. N.Klochkov, P. P.Maltsev, M.Yu, A.Presniakov, I. N.Trunkin, and A. L.Vasiliev. 2014. “Effect of (1 0 0) GaAS Substrate Misorientation on Electrophysical Parameters, Structural Properties and Surface Morphology of Metamorphic HEMT Nanoheterostructures InGaAs/InAlAs.” Journal of Crystal Growth392:11–19.Suche in Google Scholar
Garcia, I. , I. R.Stolle, C.Algora, W.Stolz, and K.Volz. 2008. “Influence of GaInP Ordering on the Electronic Quality of Concentrator Solar Cells.” Journal of Crystal Growth310:5209–13.Suche in Google Scholar
Geisz, J. F. , D. J.Friedman, J. S.Ward, A.Duda, W. J.Olavarria, T. E.Moriarty, J. T.Kiehl, M. J.Romero, A. G.Norman, and K. M.Jones. 2008. “40.8% Efficient Inverted Triple-Junction Solar Cell with Two Independently Metamorphic Junctions.” Applied Physics Letters93:123505.Suche in Google Scholar
Geisz, J. F. , M. W.Sarah Kurtz, J. S.Wanlass, W. A.Duda, D. J.Friedman, J. M.Olson, W. E.McMahon, T. E.Moriarty, and J. T.Kiehl. 2007. “High-Efficiency GaInP/GaAS/InGaAs Triple-Junction Solar Cells Grown Inverted with a Metamorphic Bottom Junction.” Applied Physics Letters91:023502.Suche in Google Scholar
Goldman, R. S. , K. L.Kavanagh, H. H.Wieder, S. N.Ehrlich, and R. M.Feenstra. 1998. “Effects of GaAS Substrate Misorientation on Strain Relaxation in InxGa1xAs Films and Multilayers.” Journal of Applied Physics83:5137.Suche in Google Scholar
Green, M. A. , K.Emery, Y.Hishikawa, W.Warta, and E. D.Dunlop. 2013. Progress in Photovoltaics21:827.Suche in Google Scholar
Hageman, P. R. , A.van Geelen, W.Gabriëlse, G. J.Bauhuis, and L. J.Giling. 1992. “Optical and Electrical Quality of InGaP Grown on GaAS with Low Pressure Metalorganic Chemical Vapour Deposition.” Journal of Crystal Growth125:336.Suche in Google Scholar
King, R. R. , D.Bushari, D.Larrabee, X. Q.Liu, E.Rehder, K.Edmondson, H.Cotal, R. K.Jones, J. H.Ermer, C. M.Fetzer, et al. 2012. Progress in Photovoltaics20:801.Suche in Google Scholar
King, R. R. , D. C.Law, K. M.Edmondson, C. M.Fetzer, G. S.Kinsey, H.Yoon, D. D.Krut, J. H.Ermer, R. A.Sherif, and N. H.Karam. 2007. “Advances in High-Efficiency Iii-V Multijunction Solar Cells.” Advances in OptoElectronics2007:1–8. doi:10.1155/2007/29523Suche in Google Scholar
Law, D. C. , R. R.King, H.Yoon, M. J.Archer, A.Boca, C. M.Fetzer, S.Mesropian, T.Isshiki, M.Haddad, K. M.Edmondson, et al. 2010. “Future Technology Pathways of Terrestrial Iii–V Multijunction Solar Cells for Concentrator Photovoltaic Systems.” Solar Energy Materials and Solar Cells94:1314–18.Suche in Google Scholar
Lee, K. E. , and E. A.Fitzgerald. 2010. “High-Quality Metamorphic Compositionally Graded InGaAs Buffers.” Journal of Crystal Growth312:250–7.Suche in Google Scholar
Leite, M. S. , R. L.Woo, J. N.Munday, W. D.Hong, S.Mesropian, D. C.Law, and H. A.Atwater2013. “Towards an Optimized All Lattice-Matched InAlAs/InGaAsP/InGaAs Multijunction Solar Cell with Efficiency >50%.” Applied Physics Letters102:033901.Suche in Google Scholar
Matthews, J. W. , and A. E.Blakeslee. 1974. “Defects in Epitaxial Multilayers: I. Misfit Dislocations.” Journal CrystalGrowth27:118.Suche in Google Scholar
Mori, M. J. , and E. A.Fitzgerald2009. “Microstructure and Luminescent Properties of Novel InGaP Alloys on Relaxed GaAsP Substrates.” Journal of Applied Physics105:013107.Suche in Google Scholar
Murata, H. , I. H.Ho, and G. B.Stringfellow. 1997. “Effects of Growth Temperature and V/Iii Ratio on Surface Structure and Ordering in Ga0.5in0.5p.” Journal Crystal Growth170:219.Suche in Google Scholar
Nguyen, H. Q. , E. Y.Chang, H. W.Yu, K. L.Lin, and C. C.Chung. 2011. “High-Quality 1-eV In0.3ga0.7as on GaAS Substrate by Metalorganic Chemical Vapor Deposition for Inverted Metamorphic Solar Cell Application.” Apex4:075501.Suche in Google Scholar
Novak, J. , S.Hasenohrl, R.Kudela, M.Kucera, M. I.Alonso, and M.Garriga. 2002. “Effect of Strain and Ordering on the Band-Gap Energy of InGaP.” Materials Science and EngineeringB88:139–42.Suche in Google Scholar
Patel, P. , D.Aiken, A.Boca, B.Cho, D.Chumney, M. B.Clevenger, A.Cornfeld, N.Fatemi, Y.Lin, J.McCarty, et al. 2012. “Experimental Results from Performance Improvement and Radiation Hardening of Inverted Metamorphic Multijunction Solar Cells.” IEEE Journal of Photovoltaics2(3):377–81.Suche in Google Scholar
Quitoriano, N. J. , and E. A.Fitzgerald. 2007. “Relaxed, High-Quality InP on GaAS by Using InGaAs and InGaP Graded Buffers to Avoid Phase Separation.” Journal of Applied Physics102:033511.Suche in Google Scholar
Romanov, A. E. , W.Pompe, S.Mathis, G. E.Beltz, and J. S.Speck. 1999. “Threading Dislocation Reduction in Strained Layers.” Journal of Applied Physics85:182.Suche in Google Scholar
Seki, Y. , H.Watanabe, and J.Matsui. 1978. “Impurity Effect on Grown-in Dislocation Density of InP and GaAS Crystals.” Journal of Applied Physics49:822.Suche in Google Scholar
Sherif, R. A. , R. R.King, N. H.Karam, and D. R.Lillington. 2005. Proceedings of the 31st IEEE Photovoltaic Specialist Conference (PVSC), Lake Buena Vista, Florida, 2005.Suche in Google Scholar
Speck, J. S. , M. A.Brewer, G.Beltz, A. E.Romanov, and W.Pompe. 1996. “Scaling Laws for the Reduction of Threading Dislocation Densities in Homogeneous Buffer Layers.” Journal of Applied Physics80:3808.Suche in Google Scholar
Tatavarti, R. , G.Hillier, C.Youtsey, D.McCallum, G.Martin, A.Wibowo, R.Navaratnarajah, F.Tuminello, D.Hertkorn, M.Disabb, et al. 2009, “Lightweight, Low-Cost InGaP/GaAs Dual-Junction Solar Cells on 4” Epitaxial Liftoff (ELO) Wafers.” Proc. of SPIE Vol. 7407, 74070B. doi: 10.1117/12.827449.Suche in Google Scholar
Yu, H. W. , E. Y.Chang, H. Q.Nguyen, J. T.Chang, C. C.Chung, C. I.Kuo, Y. Y.Wong, and W. C.Wang. 2010. “Effect of Substrate Misorientation on the Material Properties of GaAS/Al0.3ga0.7as Tunnel Diodes.” Applied Physics Letters97:231903.Suche in Google Scholar
Zunger, A. , and S.Mahajan. 1994. Handbook on Semiconductors, vol. 3, Chap. 19, “Atomic Ordering and Phase Separation in Epitaxial III-V Alloys”, 1402–1507, edited by S. Mahajan. Amsterdam: Elsevier Science.Suche in Google Scholar
©2014 by De Gruyter
Artikel in diesem Heft
- Frontmatter
- III–V Multijunction Solar Cell Integration with Silicon: Present Status, Challenges and Future Outlook
- Monolithic Integration of Diluted-Nitride III–V-N Compounds on Silicon Substrates: Toward the III–V/Si Concentrated Photovoltaics
- Study of the Growth and Dislocation Blocking Mechanisms in InxGa1−xAs Buffer Layer for Growing High-Quality In0.5Ga0.5P, In0.3Ga0.7As, and In0.52Ga0.48As on Misoriented GaAs Substrate for Inverted Metamorphic Multijunction Solar Cell Application
- Organic, Flexible, Polymer Composites for High-Temperature Piezoelectric Applications
- Modeling of a Bridge-Shaped Nonlinear Piezoelectric Energy Harvester
- Enhanced Vibration Energy Harvesting Through Multilayer Textured Pb(Mg1/3Nb2/3)O3–PbZrO3–PbTiO3 Piezoelectric Ceramics
- Load-Tolerant, High-Efficiency Self-Powered Energy Harvesting Scheme Using a Nonlinear Approach
- Comparative Analysis of One-Dimensional and Two-Dimensional Cantilever Piezoelectric Energy Harvesters
- Modeling of Hybrid Piezoelectrodynamic Generators
- Opto-electrical Behavior of Pb(Zn1/3Nb2/3)O3–Pb0.97La0.03(Zr,Ti)O3 Transparent Ceramics with Varying Defect Structure
- Feasibility Study for Small Scaling Flywheel-Energy-Storage Systems in Energy Harvesting Systems
- Ca0.15Zr0.85O1.85 Thin Film for Application to MIM Capacitor on Organic Substrate
- Erratum to EHS 1 (1–2), 69–78 (2014), A High-Temperature Thermoelectric Generator Based on Oxides
- A Direct Entropic Approach to Uniform and Spatially Continuous Dynamical Models of Thermoelectric Devices
Artikel in diesem Heft
- Frontmatter
- III–V Multijunction Solar Cell Integration with Silicon: Present Status, Challenges and Future Outlook
- Monolithic Integration of Diluted-Nitride III–V-N Compounds on Silicon Substrates: Toward the III–V/Si Concentrated Photovoltaics
- Study of the Growth and Dislocation Blocking Mechanisms in InxGa1−xAs Buffer Layer for Growing High-Quality In0.5Ga0.5P, In0.3Ga0.7As, and In0.52Ga0.48As on Misoriented GaAs Substrate for Inverted Metamorphic Multijunction Solar Cell Application
- Organic, Flexible, Polymer Composites for High-Temperature Piezoelectric Applications
- Modeling of a Bridge-Shaped Nonlinear Piezoelectric Energy Harvester
- Enhanced Vibration Energy Harvesting Through Multilayer Textured Pb(Mg1/3Nb2/3)O3–PbZrO3–PbTiO3 Piezoelectric Ceramics
- Load-Tolerant, High-Efficiency Self-Powered Energy Harvesting Scheme Using a Nonlinear Approach
- Comparative Analysis of One-Dimensional and Two-Dimensional Cantilever Piezoelectric Energy Harvesters
- Modeling of Hybrid Piezoelectrodynamic Generators
- Opto-electrical Behavior of Pb(Zn1/3Nb2/3)O3–Pb0.97La0.03(Zr,Ti)O3 Transparent Ceramics with Varying Defect Structure
- Feasibility Study for Small Scaling Flywheel-Energy-Storage Systems in Energy Harvesting Systems
- Ca0.15Zr0.85O1.85 Thin Film for Application to MIM Capacitor on Organic Substrate
- Erratum to EHS 1 (1–2), 69–78 (2014), A High-Temperature Thermoelectric Generator Based on Oxides
- A Direct Entropic Approach to Uniform and Spatially Continuous Dynamical Models of Thermoelectric Devices