Abstract
One of the most significant features of lead-halide perovskites is their ability to have comparably slow recombination despite the fact that these materials are mostly processed from solution at room temperature. The slow recombination allows achieving high open-circuit voltages when the lead-halide perovskite layers are used in solar cells. This perspective discusses the state of the art of our understanding and of experimental data with regard to recombination and open-circuit voltages in lead-halide perovskites. A special focus is put onto open questions that the community has to tackle to design future photovoltaic and optoelectronic devices based on lead-halide perovskites and other semiconductors with similar properties.
This article is part of a discussion meeting issue ‘Energy materials for a low carbon future’.
Keywords: carrier lifetime, surface recombination, built-in voltage, selective contacts, multiphonon recombination
1. Introduction
During most of the history of photovoltaic power conversion, there has been a clear correlation between the amount of effort and energy that was put into the preparation of a light absorbing semiconductor and the electronic quality of said semiconductor. The longest charge-carrier lifetimes were correlated with the lowest defect densities and therefore usually the highest crystallinity of the bulk material and the interfaces. If we look at the top of the solar cell efficiency charts [1], this is still true given that epitaxially grown GaAs and monocrystalline Si solar cells [2,3] have the highest efficiencies of all single junction solar cells as shown in figure 1. What has changed is the ranking among the thin-film technologies that were dominated by Cu(In,Ga)Se2 [7–9] and CdTe at least from the perspective of record efficiency on the lab scale. Since 2018, the non-stabilized efficiency records for single junction thin-film solar cells are held by lead-halide perovskite solar cells [5], despite the fact that these are solution-processed semiconductors and do not even possess an optimal band gap for single junction solar cells.
The key requirement for achieving efficient photovoltaic power conversion is to combine high absorption coefficients with long charge-carrier lifetime [10,11]. An obvious problem for assigning a simple figure of merit to this insight is that the absorption coefficient is a function of energy, so the magnitude and shape of the absorption coefficient matters. For instance, the high absorption coefficient of Si above its direct band gap at 3.4 eV is clearly much less relevant for photovoltaics than its low absorption coefficient close to the indirect band gap at 1.12 eV. However, when studying only direct semiconductors, there are relatively few differences in the shape and sharpness of the absorption between the most important photovoltaic materials [12]. Thus, it is the charge-carrier lifetime that differs between the different technologies and that makes the difference between high and low power conversion efficiencies [13]. In the following, we will, therefore, discuss the concept of the charge-carrier lifetime in more detail. In halide perovskites just as in other direct semiconductors [14], charge-carrier lifetimes are often studied using the method of transient photoluminescence [15–19]. We, therefore, show some exemplary transient photoluminescence data on halide perovskite layers on a hole transport layer and discuss how to interpret and analyse the data in terms of charge-carrier lifetimes and surface recombination velocities. Subsequently, I will discuss the step from high carrier lifetimes to high open-circuit voltages and completed solar cells. In order to rationalize the results, non-radiative recombination and the theory of multiphonon recombination is briefly reviewed before the key open questions in the context of recombination and high open-circuit voltages in lead-halide perovskites are discussed.
2. Charge-carrier lifetimes
The charge-carrier lifetimes of semiconductor films, layer stacks or devices are both highly relevant for their functionality as well as subject to a variety of more or less intuitive definitions that complicate an understanding of their meaning. The charge-carrier lifetime describes the dependence of the electron or hole concentration as a function of time after the generation of electrons and holes (e.g. by a laser pulse) has stopped. In a doped semiconductor, at low excitation densities, one type of carrier will be in excess of the other. This situation is termed low-level injection. For instance, in a p-type semiconductor, the electrons would be the minority carriers while the hole concentration p would be constant as a function of time t and given by the density NA of ionized acceptors. Clearly, in this situation, it only makes sense to define an electron lifetime, while the hole lifetime would be infinite because it does not decay at all with time. By contrast, if the excitation density is high enough and the doping density low enough that excess charge carriers of both types exceed the concentration of carriers due to ionized dopants, the sample is in high-level injection and there will be a finite lifetime for both types of carriers.
In addition to the aspect of injection level, lifetimes can be associated with different recombination mechanisms, such as radiative recombination (with rate Rrad), Shockley–Read–Hall (SRH) recombination [20,21] via defects in the bulk (with rate RSRH), SRH recombination at surfaces and Auger recombination (with rate RAug). Including the three bulk recombination mechanisms, the electron density n as a function of time t and position x after a laser pulse is given by the differential equation
2.1 |
where p is the hole density, ni the intrinsic carrier concentration, krad is the radiative recombination coefficient, pr is the reabsorption probability taking into account the effect of photon recycling [22–25], τn,p are the SRH lifetimes for electrons and holes, and Cn,p are the Auger coefficients for electrons and holes. This differential equation has boundary conditions for the surfaces of the semiconductor that consider the loss of electrons via surface recombination. The boundary condition at the edges x = 0 and x = d, where d is the thickness, are given by
2.2 |
where the sign on the left-hand side is positive for x = 0 and negative for x = d. Here, Dn is the diffusion constant for electrons and Sn and Sp are the surface recombination velocities for electrons and holes. Note that equations analogous to equations (2.1) and (2.2) can be written down for the case of holes. Equations (2.1) and (2.2) imply that the measured decay may be influenced by four different recombination mechanisms (radiative, bulk defect, surface defect and Auger) and that in general each of the different mechanisms is a nonlinear function of n or p. Equations (2.1) and (2.2) alone do not immediately suggest that the concept of a lifetime is suitable to describe the decay of electrons or holes after a pulse. However, the concept of a lifetime becomes substantially more intuitive and meaningful in the case of low-level injection that is relevant for many inorganic semiconductors. In low-level injection, each of the recombination terms can be simplified such that it depends linearly on minority-carrier density. For instance, in a p-type semiconductor, equation (2.1) becomes
2.3 |
where the effective lifetime τeff is defined as the inverse of the square brackets in equation (2.3). Even the inclusion of surfaces in the concept of an effective lifetime is possible for low-level injection. For the case that the surface recombination velocities for electrons as minorities are the same at both surfaces, a surface lifetime can be determined as [26]
2.4 |
Here, the first term on the right-hand side describes the recombination at the surface itself and the second term describes the diffusion to the surface which may be limiting for large thicknesses d, low diffusion constants Dn and high surface recombination velocities Sn. The total effective lifetime in low-level injection would then follow from adding up the inverse lifetimes and then inverting the resulting sum again leading to
2.5 |
Note that τeff in equation (2.5) is a constant, carrier concentration independent value. This is due to the fact that even the two recombination mechanisms (radiative and Auger) that are in general nonlinear, become linear in minority carrier concentration in low-level injection. The doping density sets a natural upper limit to the effective lifetime, which can never exceed the value in the radiative limit that is given by . Because krad in direct semiconductors is typically on the order of 10−10 cm3 s−1 (e.g. 2 ×10−10 cm3 s−1 in GaAs [27,28], 0.6 × 10−10 cm3 s−1 to 2 × 10−10 cm3 s−1 in Cu(In,Ga)Se2 with different In to Ga ratios [29]), lifetimes of hundreds of ns are already a good value for direct semiconductors at typical doping concentrations of 1016 cm−3. Typical values reported for the radiative and bimolecular recombination coefficients in the literature for CH3NH3PbI3 are presented in table 1. By contrast, in indirect semiconductors such as Si, krad is typically substantially smaller (in Si krad = 4.7 × 10−15 cm3 s−1 [33]) allowing much longer lifetimes. For instance, in monocrystalline Si with good surface passivation, lifetimes of tens of milliseconds are possible [34]. Such lifetimes could (theoretically) be achieved in GaAs only for extremely low doping densities of order 1012 cm−3.
Table 1.
reference | kext (cm3 s−1) | krad (cm3 s−1) | knon (cm3 s−1) |
---|---|---|---|
Staub et al. [23] | 4.78 × 10−11 | 8.7 × 10−10 | neglected |
Staub et al. [23,24] | 4.78 × 10−11 | 8.4 × 10−11 | 4.4 × 10−11 |
Crothers et al. [31] | 1.4 × 10−10 to 2 × 10−11 | 6.8 × 10−10 | 0 |
Richter et al. [32] | 8.1 × 10−11 | 7.1 × 10−11 | 7.2 × 10−11 |
Richter et al. [32] | 7.9 × 10−11 | 1.8 × 10−10 | 5.6 × 10−11 |
For lead-iodide perovskite films, layer stacks and devices, transient measurements such as transient photoluminescence are typically done in high-level injection due to the low doping density of the materials. This implies that both radiative and Auger recombination will lead to terms in the recombination rate that are quadratic or cubic in charge-carrier density thereby leading to non-exponential decays [23]. Figure 2 shows a photoluminescence decay of a methylammonium-lead-iodide (MAPI) film deposited on a glass/ITO substrate coated with a thin layer of PTAA (poly(triarylamin)) that is used as a hole contact in solar cells and light emitting diodes [35]. For details on the simulation, see the electronic supplementary material accompanying this article. We note that the decay is faster for short times followed by a nearly exponential decay at longer times. The fast features at short times can be explained by radiative recombination. At longer times, the long decay can either be due to bulk recombination or surface recombination at the MAPI/PTAA interface. In addition, there a range of additional effects such as lateral diffusion away from the illuminated spot, charge transfer to the contact layers, charge accumulation in the contact layers [36] and lateral transport on the contact layers that may affect the shape of the transients. While the two effects cannot be distinguished unambiguously, it is possible to perform numerical simulations [35,36] to include the different effects mentioned above and to estimate minimum values for SRH lifetime in the bulk and upper limits for the surface recombination velocity. In the present case, figure 2 suggests that S < 1 cm s−1 and τbulk > 5 µs best describe the experimental decay. To put the values of surface recombination velocities into perspective, a brief word on typical values for inorganic semiconductors used for photovoltaics. In crystalline Si wafers, good surface passivation leads to surface recombination velocities just below 10 cm s−1 [34]. In the case of AlGaAs/GaAS or GaInP/GaAs interfaces, surface recombination velocities on the order of 1–15 cm s−1 have been reported already in the 1980s [37,38]. In CdS/CdTe heterostructures surface recombination, velocities are typically on the order of 105 cm s−1 [39] while high-quality MgCdTe/CdTe heterostructures have surface recombination velocities on the order of 1 cm s−1 [40].
3. Realizing high open-circuit voltages in solar cells
We have shown in the previous chapter that long bulk lifetimes and low surface recombination velocities S < 10 cm s−1 are possible in perovskite layers and at perovskite/PTAA interfaces. In order to actually benefit from these promising values in a complete device, it is necessary to achieve low surface recombination velocities at both interfaces in the device and to maintain a good compromise between low surface recombination velocities, low series resistances of the used contact layers and a beneficial electric field distribution inside the absorber layer that optimizes charge collection and minimizes recombination at the maximum power point [41–45].
Figure 3 shows how for a constant thickness (d = 300 nm) and a perfect defect free bulk, the open-circuit voltage, the voltage at the maximum power point and the power conversion efficiency of a MAPI solar cell depends on surface recombination and series resistance. The simulation includes the effects of Auger recombination using an Auger coefficient C = 5.4 × 10−28 cm6 s−1 [46] and photon recycling as described in more detail in [47–49]. For details on the simulation, see the electronic supplementary material accompanying this article. We note that for MAPI efficiencies greater than 25% should be possible but require series resistances below 5 Ωcm2 and surface recombination velocities below 100 cm s−1 at both contacts. The maximum possible open-circuit voltage is 1.32 eV for MAPI with a band gap Eg = 1.6 eV and it decreases for S > 100 cm s−1, consistent with recent reports by Wang et al. [50]. Note that the exact dependence of Voc on S also depends on the minority carrier concentration at the two contacts which are determined in turn by the differences between contact workfunction and conduction band (at the cathode) or valence band (at the anode) of the perovskite. Here, we use an offset of 0.1 eV on each side which leads to a built-in voltage Vbi = 1.4 eV (band gap minus twice the offset of 0.1 eV). The effect of the (area-related) series resistance Rs on power conversion efficiency η and the voltage Vmpp at the maximum power point is quite substantial and relatively linear with a doubling of Rs leading to double the loss in Vmpp relative to the optimum value for Rs = 0 which is consistent with analytical equations [51] describing the effect of Rs on Vmpp. Resistive effects in perovskite solar cells may originate from external contact layers such as the ITO [52] but also from electron and hole contact layers, especially if these are made up of undoped organic semiconductors such as PTAA and PCBM or C60 as is commonly done in p–i–n-type device stacks. The low permittivity of these electron and hole contact materials combined with the high permittivity [53] of the perovskite absorber layer and the electric field screening by ion movement may also lead to low electric fields in the perovskite absorber that may even change sign and impede majority carrier extraction as illustrated in [54,55]. In these cases, the good passivation provided by at least some organic contact materials may compromise efficient charge extraction [43,44].
It is worth comparing the simulated open-circuit voltages shown in figure 3b with experimentally achieved values of the open-circuit voltage Voc. Figure 4a shows Jsc as a function of Voc for a range of highly efficient lead-halide perovskite solar cells reported in the literature (table 2 for a full set of values and references). The solid line defines the respective values of Jsc and Voc in the Shockley–Queisser (SQ) model. Figure 4a shows that typical values for most of the high-efficiency lead-halide perovskite cells are Voc ≈ 1.15 V and Jsc ≈ 22 mA cm−2. Some are optimized for higher open-circuit voltages Voc > 1.2 V [15,35] and some have slightly lower band gaps allowing higher photocurrents around Jsc ≈ 25 mA cm−2 [5,58].
Table 2.
reference | Eg (eV) | η | Jsc(mAcm−2) | Voc (V) | FF | ΔVoc (V) | Jsc/Jsc,SQ |
---|---|---|---|---|---|---|---|
Jeon [5]* | 1.56 | 22.6 | 24.9 | 1.13 | 80.5 | 0.160 | 0.924 |
Luo [57]* | 1.59 | 20.9 | 21.8 | 1.18 | 81.7 | 0.140 | 0.840 |
Turren-Cruz [58] | 1.53 | 20.4 | 25.1 | 1.08 | 75.5 | 0.179 | 0.893 |
Stolterfoht [59]* | 1.62 | 19.6 | 22.0 | 1.15 | 77.6 | 0.197 | 0.886 |
Saidaminov [60] | 1.61 | 21.0 | 22.2 | 1.16 | 81.3 | 0.170 | 0.881 |
Zheng [61]* | 1.59 | 20.6 | 23.0 | 1.13 | 79.1 | 0.184 | 0.888 |
Wang [62] | 1.61 | 20.6 | 22.6 | 1.14 | 80.0 | 0.194 | 0.897 |
Saliba [63] | 1.63 | 21.8 | 22.8 | 1.18 | 81.0 | 0.173 | 0.924 |
Tan [64]* | 1.60 | 20.1 | 21.7 | 1.17 | 79.4 | 0.155 | 0.849 |
Yang [65]* | 1.55 | 22.1 | 24.9 | 1.10 | 80.3 | 0.173 | 0.913 |
Abdi-Jalebi [15] | 1.56 | 21.5 | 23.3 | 1.17 | 79.0 | 0.117 | 0.862 |
Abdi-Jalebi [15] | 1.78 | 17.5 | 18.0 | 1.23 | 79.0 | 0.263 | 0.890 |
Hou [66] | 1.61 | 21.2 | 22.6 | 1.17 | 80.0 | 0.164 | 0.899 |
Anaraki [67] | 1.62 | 20.5 | 22.5 | 1.18 | 77.0 | 0.163 | 0.906 |
Liu [35] | 1.60 | 20.7 | 20.5 | 1.26 | 80.1 | 0.065 | 0.803 |
However, figure 4a is not well suited to identify potential for improvement for a given device because the band gap varies slightly between the different perovskite compositions. Therefore, figure 4b provides a more direct relation between the experimental data and the values given by the SQ model by normalizing Jsc to Jsc,SQ and taking the difference between the open-circuit voltage in the SQ limit and the real open-circuit voltage. Figure 4b shows that many of the high-efficiency devices reach greater than 90% of the photocurrent in the SQ model which is a good value compared to many other technologies. However, values of approximately 95% are possible in crystalline Si and CdTe solar cells [68,69], i.e. there is still some room for improvement. The voltage loss in open circuit is typically greater than 120 mV for most devices suggesting that there is indeed still some room for improvement given that GaAs solar cells have voltage losses less than 40 mV in the best cases (e.g. Fig. 2b in [70]).
Among the solar cells presented in figure 4 and table 2, the device approaching the Voc,SQ most closely is the one labelled Liu [35]. It is based on a simple MAPI cell with a lead-acetate precursor with 8% lead-chloride and using ITO/PTAA and PCBM/BCP/Ag as hole and electron contact layers, respectively. Figure 5 shows the power density and current density curves of this MAPI solar cell (shown for the downscan at 0.1 V s−1) and compares it with three computed scenarios. Each scenario keeps the parameters of the previous case but increases first the FF, then Voc and finally Jsc to the ideal value that is defined by the SQ model. The differences between the curves then allow us to study the losses due to each mechanism. The FF of the experimental curve is 80%, but it may be up to 90% for this open-circuit voltage [71,72]. This corresponds to an increase by about 120 mV in the voltage at the maximum power point. Further increasing Voc by removing all non-radiative recombination would only yield another approximately 60 mV in Voc and Vmpp. A further substantial gain in power conversion efficiency would be possible if more complete light absorption was achieved. Finally, increasing Jsc to Jsc,SQ would yield an efficiency of approximately 30%, i.e. approximately 10% more than the experimental value and approximately 5% more than the value where Voc and FF are already taken as ideal. Thus, for this specific device, the open-circuit voltage is nearly fully optimized with little additional gains being expected. By contrast, the photocurrent is far from optimized, which sets optical means to trap light, processes to fabricate thicker layers, optimization of charge collection and minimization of parasitic absorption on the research agenda.
4. Multiphonon recombination
We learned in the previous sections that solution-processed lead-halide perovskites allow closely approaching the thermodynamic limit for the open-circuit voltage. This is possible thanks to very long bulk lifetimes of several microseconds and by achieving surface recombination velocities that are lower than those of the best passivation layers for crystalline Si solar cells [34]. Thus, the question remains how we can rationalize these long lifetimes. There have been various physical mechanisms suggested in the literature to explain slow recombination [73–79]. Some mechanisms, such as the existence of an indirect gap [80,81] serve to explain slow radiative recombination. However, radiative recombination in MAPI is not actually slow but corresponds to what is expected [23,82] based on the absorption coefficient of MAPI and the fairly low effective densities of states. In addition, slow radiative recombination would not help achieving higher efficiencies as detailed in [11,83]. Only slow non-radiative recombination would actually be helpful. In order to achieve slow non-radiative recombination, a semiconductor (bulk or interface) needs to have both a low density of defects and low transition rates between at least one of the bands and the defect. If the defect only interacts efficiently with one of the bands, recombination will be less likely to happen. Instead, the defects will trap and then later detrap charge carriers. While this mechanism slows down transport, it would not lead to reduced open-circuit voltages.
Thus, in high-level injection, where n ≈ p, defects with substantially asymmetric capture coefficients for electrons and holes might be less detrimental than those with symmetric capture coefficients. Asymmetry of capture coefficients may originate, for instance from the position of the defect and from its charge state. For instance, defects close to the conduction band are more likely to interact with electrons in the conduction band than with holes in the valence band. In addition, positively charged defects are more likely to capture electrons than neutral defects. Thus, it would be useful to be able to estimate the relevance of each of these effects on the value of the capture coefficients. These can be calculated for non-radiative processes using the theory for multiphonon recombination, which assumes that every downward transition between two states has to involve the emission of multiple phonons in order to dissipate the energy of the excited state [84,85]. Figure 6 shows the capture coefficients calculated using the theory discussed in [86] and originating from the work of Markvart [84,87] and Ridley [85,88]. The capture coefficients show a strong dependence on the position of the defect with shallow defects having one high and one extremely low capture coefficient while for deep defects the values of the capture coefficients intersect. When comparing the graph for the donor-like and the acceptor-like defect, we note that the effect of Coulomb attraction on the capture coefficients is fairly small.
In order to calculate the SRH lifetimes introduced in equation (2.1) from the capture coefficients in figure 7, we need to specify a trap density. The SRH lifetimes are given by , where NT is the trap density. If we choose NT = 1015 cm−3, we obtain the lifetimes shown in figure 7a using the capture coefficients kn,p shown in figure 6. From the SRH lifetimes, we can then calculate the effective lifetime τeff, which in the limit of high-level injection assumed here (n ≈ p) is just given as τeff = τn + τp.
The Coulomb attraction is calculated in [86,89] by using the Sommerfeld factor to modify the capture cross sections of transitions with attractive Coulomb interaction. The Sommerfeld factor for attractive transitions is given by Markvart [89]
4.1 |
where meff/m0 is the effective mass relative to the electron rest mass, RH ≈ 13.6 eV is the Rydberg energy, kT is the thermal energy and ϵr is the low-frequency dielectric permittivity. Figure 7b illustrates the result of equation (4.1) for different values of meff/m0 and εr. It becomes clear that for the case of low effective masses and high permittivities, the Sommerfeld factor does not deviate much from unity.
Thus, the main contribution to the capture coefficients in the framework of multiphonon recombination using the parameters of MAPI as input would be the position of the trap. Midgap traps would still yield reasonably high recombination coefficients in the framework of multiphonon recombination theory. However, already small shifts of the trap position relative to midgap would strongly reduce the efficiency of recombination. Density functional theory calculations in the bulk [90] and the surface [91] of MAPI have shown that most defects in MAPI are shallow with the iodine interstitial being the only intrinsic deep point defect. However, even the iodine interstitial is likely not a midgap defect. This finding in combination with figure 7a suggests that intrinsic (i.e. undoped) MAPI could tolerate quite high densities of intrinsic point defects given that they are not likely to have symmetric and high capture coefficients.
5. Open questions
While the experimental results with respect to slow recombination and high open-circuit voltages are certainly impressive and while we have some understanding and rationale for explaining the experimental results, there remain a range of open questions that should be tackled in the future.
Recently, it has been shown that the model of the harmonic oscillator for the potential energy of electrons and holes is insufficient to explain the magnitude and temperature dependence of the charge-carrier mobility in lead-halide perovskites [92]. Instead, the anharmonic fluctuations of the lead-halide bonds affect the overlap of wavefunctions and thereby charge transport. In a similar way also the multiphonon recombination models that rely on harmonic oscillators to calculate transition rates between band-like states and defect states likely have to be modified to the special case of lead-halide perovskites.
In addition, the models for recombination generally only describe bulk recombination but not interface recombination. At an interface, defects are more likely to form than in the bulk. Therefore for most inorganic solar cell materials, interface recombination is even more important than bulk recombination. Recently, it has been argued [93–95] that the high energy of vibrational modes in organic semiconductors leads to fast recombination, especially for lower band gaps. This concept applied to interfaces between perovskite absorbers and organic electron and hole contact layers would suggest that organic interface materials might accelerate recombination because it is easier to dissipate the energy into the organic material. However, this does not seem to be the case, given that perovskite/organic interfaces have extremely low surface recombination velocities [35,59,96]. Thus, a better fundamental understanding of interface recombination is needed to design interfaces rather than find suitable ones by trial and error.
The low surface recombination velocities of organic electron and hole transport layers might bring about a fundamental compromise between recombination, resistive losses and optimization of the electric field distribution in the device [43,44]. As shown in figure 3, a low value of S might benefit Voc but if it comes with a high value of Rs, the efficiency might be reduced. The same holds true if the electric field distribution is such that it makes charge extraction more difficult, because most of the electric field drops over the electron and hole transport layers [43,44]. Thus, we need a better understanding of resistive and charge collection losses at interfaces and in electron and hole transport layers. While it is generally easy to measure lumped series resistances in solar cells [97,98], it is more difficult to disentangle [99,100] the influence of different layers on the series resistance. Thus, methods are needed that measure how the internal quasi-Fermi level splitting in the perovskite absorber is coupled to the external quasi-Fermi level splitting at the contacts at the maximum power point.
Finally, an open question is the practical influence of the built-in voltage Vbi on solar cell performance. Charge-carrier diffusion lengths are generally reported to be high enough to achieve efficient charge collection via diffusion within absorbers with thicknesses of less than a micron. However, the built-in voltage still affects the concentration of minority carriers at the interfaces to electron and hole contact materials. Only a high Vbi reduces the minority carrier concentration sufficiently to suppress recombination at these interfaces. Thus, for a given value of S, there is a minimum value for Vbi that the cells need to have in order to not have additional losses at the interfaces. Figure 8 shows how high Vbi has to be for different values of S in order to achieve optimum efficiency. The simulations shown in figure 8 are done using the same method as used for figure 3, i.e. they include radiative recombination, photon recycling and Auger recombination. In the simulation, lower built-in voltages lead to increased surface recombination while higher built-in voltages lead to increased Auger recombination due to the higher concentration of majority carriers close to the contacts. Thus, there is always an optimum Vbi for every value of S.
Figure 8 shows that while Vbi can be smaller than Eg/q, it should be ‘high enough’ (which is defined differently depending on the values of S as shown in figure 8) in order to reduce losses. However, it is often not clear whether the values of Vbi leading to optimum power conversion efficiency in figure 8 are consistent with the expected workfunction difference between the contacts. For instance, for the high Voc device discussed in [35], the electron and hole contact materials (PTAA and PCBM) are fairly intrinsic and the highly conductive materials ITO and Ag should not have a sufficiently high workfunction difference to enable efficient suppression of interfacial recombination. Nevertheless, the cells are in practice able to suppress recombination quite efficiently. This suggests that we lack knowledge on what the band diagram of the cell looks like and how the workfunctions and energy levels of the contact and absorber layers enable such low interfacial recombination rates (even if we do understand the low recombination velocities).
Supplementary Material
Data accessibility
This article has no additional data.
Competing interests
I declare I have no competing interests.
Funding
The author acknowledges support from the Impuls- und Vernetzungsfonds der Helmholtz Gemeinschaft via the project PEROSEED and the Bavarian Ministry of Economic Affairs and Media, Energy and Technology for the joint projects in the framework of the Helmholtz Institute Erlangen-Nürnberg.
References
- 1.Green MA, Hishikawa Y, Dunlop ED, Levi DH, Hohl-Ebinger J, Ho-Baillie AWY. 2018. Solar cell efficiency tables (Version 52). Prog. Photovolt: Res. Appl. 26, 427–436. ( 10.1002/pip.3040) [DOI] [Google Scholar]
- 2.Yoshikawa K, et al. 2017. Silicon heterojunction solar cell with interdigitated back contacts for a photoconversion efficiency over 26%. Nat. Energy 2, 17032 ( 10.1038/nenergy.2017.32) [DOI] [Google Scholar]
- 3.Yoshikawa K, et al. 2017. Exceeding conversion efficiency of 26% by heterojunction interdigitated back contact solar cell with thin film Si technology. Sol. Energy Mater. Sol. Cells 173, 37–42. ( 10.1016/j.solmat.2017.06.024) [DOI] [Google Scholar]
- 4.Best Research Cell Efficiences. 2018. https://www.nrel.gov/pv/assets/pdfs/pv-efficiency-chart.20181221.pdf .
- 5.Jeon NJ, et al. 2018. A fluorene-terminated hole-transporting material for highly efficient and stable perovskite solar cells. Nat. Energy 3, 682–689. ( 10.1038/s41560-018-0200-6) [DOI] [Google Scholar]
- 6.Li H, Xiao Z, Ding L, Wang J. 2018. Thermostable single-junction organic solar cells with a power conversion efficiency of 14.62%. Sci. Bull. 63, 340–342. ( 10.1016/j.scib.2018.02.015) [DOI] [PubMed] [Google Scholar]
- 7.Jackson P, Wuerz R, Hariskos D, Lotter E, Witte W, Powalla M. 2016. Effects of heavy alkali elements in Cu(In,Ga)Se2 solar cells with efficiencies Up to 22.6%. Phys. Status Solidi RRL 10, 583–586. ( 10.1002/pssr.201600199) [DOI] [Google Scholar]
- 8.Friedlmeier TM, Jackson P, Bauer A, Hariskos D, Kiowski O, Wuerz R, Powalla M. 2015. Improved Photocurrent in Cu(In, Ga)Se-2 solar cells: from 20.8% to 21.7% efficiency with CdS Buffer and 21.0% Cd-free. IEEE J. Photov. 5, 1487–1491. ( 10.1109/JPHOTOV.2015.2458039) [DOI] [Google Scholar]
- 9.Jackson P, Hariskos D, Wuerz R, Kiowski O, Bauer A, Friedlmeier TM, Powalla M. 2015. Properties of Cu(In,Ga)Se2 solar cells with new record efficiencies Up to 21.7%. Phys Status Solidi-R 9, 28–31. ( 10.1002/pssr.201409520) [DOI] [Google Scholar]
- 10.Kirchartz T, Rau U. 2018. What makes a good solar cell? Adv. Energy Mater. 8, 1703385 ( 10.1002/aenm.201703385) [DOI] [Google Scholar]
- 11.Kirchartz T, Krückemeier L, Unger EL. 2018. Research update: recombination and open-circuit voltage in lead-halide perovskites. APL Mater. 6, 100702 ( 10.1063/1.5052164) [DOI] [Google Scholar]
- 12.De Wolf S, Holovsky J, Moon SJ, Löper P, Niesen B, Ledinsky M, Haug FJ, Yum JH, Ballif C. 2014. Organometallic halide perovskites: sharp optical absorption edge and its relation to photovoltaic performance. J. Phys. Chem. Lett. 5, 1035–1039. ( 10.1021/jz500279b) [DOI] [PubMed] [Google Scholar]
- 13.Blank B, Kirchartz T, Lany S, Rau U. 2017. Selection metric for photovoltaic materials screening based on detailed-balance analysis. Phys. Rev. Appl. 8, 024032 ( 10.1103/PhysRevApplied.8.024032) [DOI] [Google Scholar]
- 14.Ahrenkiel RK. 1993. Minority-carrier lifetime in III-V semiconductors. Minority Carriers in Iii-V Semicond. Phys. Appl. 39, 39–150. ( 10.1016/S0080-8784(08)62594-6) [DOI] [Google Scholar]
- 15.Abdi-Jalebi M, et al. 2018. Maximizing and stabilizing luminescence from halide perovskites with potassium passivation. Nature 555, 497–501. ( 10.1038/nature25989) [DOI] [PubMed] [Google Scholar]
- 16.deQuilettes DW, et al. 2016. Photo-induced halide redistribution in organic-inorganic perovskite films. Nat. Commun. 7, 11683 ( 10.1038/ncomms11683) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.deQuilettes DW, Vorpahl SM, Stranks SD, Nagaoka H, Eperon GE, Ziffer ME, Snaith HJ, Ginger DS. 2015. Impact of microstructure on local carrier lifetime in perovskite solar cells. Science 348, 683–686. ( 10.1126/science.aaa5333) [DOI] [PubMed] [Google Scholar]
- 18.Stranks SD, Burlakov VM, Leijtens T, Ball JM, Goriely A, Snaith HJ. 2014. Recombination kinetics in organic-inorganic perovskites: excitons, free charge, and subgap states. Phys. Rev. Appl. 2, 034007 ( 10.1103/PhysRevApplied.2.034007) [DOI] [Google Scholar]
- 19.Stranks SD, Eperon GE, Grancini G, Menelaou C, Alcocer MJP, Leijtens T, Herz LM, Petrozza A, Snaith HJ. 2013. Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber. Science 342, 341–344. ( 10.1126/science.1243982) [DOI] [PubMed] [Google Scholar]
- 20.Shockley W, Read WT. 1952. Statistics of the recombination of holes and electrons. Phys. Rev. 87, 835–842. ( 10.1103/PhysRev.87.835) [DOI] [Google Scholar]
- 21.Hall RN. 1952. Electron-hole recombination in germanium. Phys. Rev. 87, 387 ( 10.1103/PhysRev.87.387) [DOI] [Google Scholar]
- 22.Asbeck P. 1977. Self-absorption effects on radiative lifetime in GaAs-GaAlAs double heterostructures. J. Appl. Phys. 48, 820–822. ( 10.1063/1.323633) [DOI] [Google Scholar]
- 23.Staub F, Hempel H, Hebig JC, Mock J, Paetzold UW, Rau U, Unold T, Kirchartz T. 2016. Beyond bulk lifetimes: insights into lead halide perovskite films from time-resolved photoluminescence. Phys. Rev. Appl. 6, 044017 ( 10.1103/PhysRevApplied.6.044017) [DOI] [Google Scholar]
- 24.Staub F, Kirchartz T, Bittkau K, Rau U. 2017. Manipulating the net radiative recombination rate in lead halide perovskite films by modification of light outcoupling. J. Phys. Chem. Lett. 8, 5084–5090. ( 10.1021/acs.jpclett.7b02224) [DOI] [PubMed] [Google Scholar]
- 25.Rau U, Paetzold UW, Kirchartz T. 2014. Thermodynamics of light management in photovoltaic devices. Phys. Rev. B 90, 035211 ( 10.1103/PhysRevB.90.035211) [DOI] [Google Scholar]
- 26.Sproul AB. 1994. Dimensionless solution of the equation describing the effect of surface recombination on carrier decay in semiconductors. J. Appl. Phys. 76, 2851–2854. ( 10.1063/1.357521) [DOI] [Google Scholar]
- 27.Casey HC, Stern F. 1976. Concentration-dependent absorption and spontaneous emission of heavily doped GaAs. J. Appl. Phys. 47, 631–643. ( 10.1063/1.322626) [DOI] [Google Scholar]
- 28.Stern F. 1976. Calculated spectral dependence of gain in excited GaAs. J. Appl. Phys. 47, 5382–5386. ( 10.1063/1.322565) [DOI] [Google Scholar]
- 29.Werner JH, Mattheis J, Rau U. 2005. Efficiency limitations of polycrystalline thin film solar cells: case of Cu(In,Ga)Se2. Thin Solid Films 480–481, 399–409. ( 10.1016/j.tsf.2004.11.052) [DOI] [Google Scholar]
- 30.Staub F, Rau U, Kirchartz T. 2018. Statistics of the auger recombination of electrons and holes via defect levels in the band gap - application to lead-halide perovskites. ACS Omega 3, 8009–8016. ( 10.1021/acsomega.8b00962) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Crothers TW, Milot RL, Patel JB, Parrott ES, Schlipf J, Müller-Buschbaum P, Johnston MB, Herz LM. 2017. Photon reabsorption masks intrinsic bimolecular charge-carrier recombination in CH3NH3PbI3 perovskite. Nano Lett. 17, 5782–5789. ( 10.1021/acs.nanolett.7b02834) [DOI] [PubMed] [Google Scholar]
- 32.Richter JM, et al. 2016. Enhancing photoluminescence yields in lead halide perovskites by photon recycling and light out-coupling. Nat. Commun. 7, 13941 ( 10.1038/ncomms13941) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Trupke T, Green MA, Würfel P, Altermatt PP, Wang A, Zhao J, Corkish R. 2003. Temperature dependence of the radiative recombination coefficient of intrinsic crystalline silicon. J. Appl. Phys. 94, 4930–4937. ( 10.1063/1.1610231) [DOI] [Google Scholar]
- 34.Richter A, Glunz SW, Werner F, Schmidt J, Cuevas A. 2012. Improved quantitative description of auger recombination in crystalline silicon. Phys. Rev. B 86, 165202 ( 10.1103/PhysRevB.86.165202) [DOI] [Google Scholar]
- 35.Liu Z, et al. 2019. Open-circuit voltages exceeding 1.26 V in planar methylammonium lead iodide perovskite solar cells. ACS Energy Lett. 4, 110–117. ( 10.1021/acsenergylett.8b01906) [DOI] [Google Scholar]
- 36.Krogmeier B, Staub F, Grabowski D, Rau U, Kirchartz T. 2018. Quantitative analysis of the transient photoluminescence of CH3NH3PbI3/PC61BM heterojunctions by numerical simulations. Sustain. Energy Fuels 2, 1027–1034. ( 10.1039/C7SE00603A) [DOI] [Google Scholar]
- 37.Olson JM, Ahrenkiel RK, Dunlavy DJ, Keyes B, Kibbler AE. 1989. Ultralow recombination velocity at Ga0.5In0.5P/GaAs heterointerfaces. Appl. Phys. Lett. 55, 1208–1210. ( 10.1063/1.101656) [DOI] [Google Scholar]
- 38.Molenkamp LW, van Blik HFJ. 1988. Very low interface recombination velocity in (Al,Ga)As heterostructures grown by organometallic vapor phase epitaxy. J. Appl. Phys. 64, 4253–4256. ( 10.1063/1.341298) [DOI] [Google Scholar]
- 39.Kuciauskas D, Kanevce A, Burst JM, Duenow JN, Dhere R, Albin DS, Levi DH, Ahrenkiel RK. 2013. Minority carrier lifetime analysis in the bulk of thin-film absorbers using subbandgap (two-photon) excitation. IEEE J. Photov. 3, 1319–1324. ( 10.1109/JPHOTOV.2013.2270354) [DOI] [Google Scholar]
- 40.Zhao Y, et al. 2016. Monocrystalline CdTe solar cells with open-circuit voltage over 1 V and efficiency of 17%. Nat. Energy 1, 16067 ( 10.1038/nenergy.2016.67) [DOI] [Google Scholar]
- 41.Brendel R, Peibst R. 2016. Contact selectivity and efficiency in crystalline silicon photovoltaics. IEEE J. Photov. 6, 1413–1420. ( 10.1109/JPHOTOV.2016.2598267) [DOI] [Google Scholar]
- 42.Roe ET, Egelhofer KE, Lonergan MC. 2018. Limits of contact selectivity/recombination on the open-circuit voltage of a photovoltaic. ACS Appl. Energy Mater. 1, 1037–1046. ( 10.1021/acsaem.7b00179) [DOI] [Google Scholar]
- 43.Courtier NE, Cave JM, Foster JM, Walker AB, Richardson G. 2019. How transport layer properties affect perovskite solar cell performance: insights from a coupled charge transport/ion migration model. Energ. Environ. Sci. 12, 396–409. ( 10.1039/C8EE01576G) [DOI] [Google Scholar]
- 44.Tessler N, Vaynzof Y. 2018. Preventing hysteresis in perovskite solar cells by undoped charge blocking layers. ACS Appl. Energy Mater. 1, 676–683. ( 10.1021/acsaem.7b00176) [DOI] [Google Scholar]
- 45.Kirchartz T, Bisquert J, Mora-Sero I, Garcia-Belmonte G. 2015. Classification of solar cells according to mechanisms of charge separation and charge collection. Phys. Chem. Chem. Phys. 17, 4007–4014. ( 10.1039/C4CP05174B) [DOI] [PubMed] [Google Scholar]
- 46.Braly IL, deQuilettes DW, Pazos-Outon LM, Burke S, Ziffer ME, Ginger DS, Hillhouse HW. 2018. Hybrid perovskite films approaching the radiative limit with over 90% photoluminescence quantum efficiency. Nat. Photon. 12, 355–361. ( 10.1038/s41566-018-0154-z) [DOI] [Google Scholar]
- 47.Mattheis J, Werner JH, Rau U. 2008. Finite mobility effects on the radiative efficiency limit of Pn-junction solar cells. Phys. Rev. B 77, 085203 ( 10.1103/PhysRevB.77.085203) [DOI] [Google Scholar]
- 48.Mattheis J. 2008. Mobility and homogeneity effects on the power conversion efficiency of solar cells, p. 140 Stuttgart, Germany: Universität Stuttgart. [Google Scholar]
- 49.Kirchartz T, Staub F, Rau U. 2016. Impact of photon recycling on the open-circuit voltage of metal halide perovskite solar cells. ACS Energy Lett. 1, 731–739. ( 10.1021/acsenergylett.6b00223) [DOI] [Google Scholar]
- 50.Wang J, Fu W, Jariwala S, Sinha I, Jen AKY, Ginger DS. 2018. Reducing surface recombination velocities at the electrical contacts will improve perovskite photovoltaics. ACS Energy Lett. 4, 222–227. ( 10.1021/acsenergylett.8b02058) [DOI] [Google Scholar]
- 51.Taretto K, Soldera M, Troviano M. 2013. Accurate explicit equations for the fill factor of real solar cells - applications to thin-film solar cells. Prog. Photovolt: Res. Appl. 21, 1489–1498. ( 10.1002/pip.2235) [DOI] [Google Scholar]
- 52.Jacobs DA, Catchpole KR, Beck FJ, White TP. 2016. A re-evaluation of transparent conductor requirements for thin-film solar cells. J. Mater. Chem. A 4, 4490–4496. ( 10.1039/C6TA01670G) [DOI] [Google Scholar]
- 53.Sendner M, et al. 2016. Optical phonons in methylammonium lead halide perovskites and implications for charge transport. Mater. Horiz. 3, 613–620. ( 10.1039/C6MH00275G) [DOI] [Google Scholar]
- 54.Calado P, Telford AM, Bryant D, Li X, Nelson J, O'Regan BC, Barnes PRF. 2016. Evidence for ion migration in hybrid perovskite solar cells with minimal hysteresis. Nat. Commun. 7, 13831 ( 10.1038/ncomms13831) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Belisle RA, Nguyen WH, Bowring AR, Calado P, Li X, Irvine SJC, McGehee MD, Barnes PRF, O'Regan BC. 2017. Interpretation of inverted photocurrent transients in organic lead halide perovskite solar cells: proof of the field screening by mobile ions and determination of the space charge layer widths. Energ. Environ. Sci. 10, 192–204. ( 10.1039/C6EE02914K) [DOI] [Google Scholar]
- 56.Shockley W, Queisser HJ. 1961. Detailed balance limit of efficiency of Pn-junction solar cells. J. Appl. Phys. 32, 510–519. ( 10.1063/1.1736034) [DOI] [Google Scholar]
- 57.Luo D, et al. 2018. Enhanced photovoltage for inverted planar heterojunction perovskite solar cells. Science 360, 1442 ( 10.1126/science.aap9282) [DOI] [PubMed] [Google Scholar]
- 58.Turren-Cruz SH, Hagfeldt A, Saliba M. 2018. Methylammonium-free, high-performance, and stable perovskite solar cells on a planar architecture. Science 362, 449–453. ( 10.1126/science.aat3583) [DOI] [PubMed] [Google Scholar]
- 59.Stolterfoht M, et al. 2018. Visualization and suppression of interfacial recombination for high-efficiency large-area pin perovskite solar cells. Nat. Energy 3, 847–854. ( 10.1038/s41560-018-0219-8) [DOI] [Google Scholar]
- 60.Saidaminov MI, et al. 2018. Suppression of atomic vacancies via incorporation of isovalent small ions to increase the stability of halide perovskite solar cells in ambient air. Nat. Energy 3, 648–654. ( 10.1038/s41560-018-0192-2) [DOI] [Google Scholar]
- 61.Zheng X, Chen B, Dai J, Fang Y, Bai Y, Lin Y, Wei H, Zeng X, Huang J. 2017. Defect passivation in hybrid perovskite solar cells using quaternary ammonium halide anions and -cations. Nat. Energy 2, 17102 ( 10.1038/nenergy.2017.102) [DOI] [Google Scholar]
- 62.Wang Z, Lin Q, Chmiel FP, Sakai N, Herz LM, Snaith HJ. 2017. Efficient ambient-air-stable solar cells with 2D-3D heterostructured butylammonium-caesium-formamidinium lead halide perovskites. Nat. Energy 2, 17135 ( 10.1038/nenergy.2017.135) [DOI] [Google Scholar]
- 63.Saliba M, et al. 2016. Incorporation of rubidium cations into perovskite solar cells improves photovoltaic performance. Science 354, 206–209. ( 10.1126/science.aah5557) [DOI] [PubMed] [Google Scholar]
- 64.Tan H, et al. 2017. Efficient and stable solution-processed planar perovskite solar cells via contact passivation. Science, 355, 722–726. ( 10.1126/science.aai9081) [DOI] [PubMed] [Google Scholar]
- 65.Yang WS, et al. 2017. Iodide management in formamidinium-lead-halide-based perovskite layers for efficient solar cells. Science 356, 1376–1379. ( 10.1126/science.aan2301) [DOI] [PubMed] [Google Scholar]
- 66.Hou Y, et al. 2017. A generic interface to reduce the efficiency-stability-cost gap of perovskite solar cells. Science 358, 1192–1197. ( 10.1126/science.aao5561) [DOI] [PubMed] [Google Scholar]
- 67.Anaraki EH, et al. 2016. Highly efficient and stable planar perovskite solar cells by solution-processed tin oxide. Energ. Environ. Sci. 9, 3128–3134. ( 10.1039/C6EE02390H) [DOI] [Google Scholar]
- 68.Nayak PK, Cahen D. 2013. Updated assessment of possibilities and limits for solar cells. Adv. Mater. 26, 1622–1628. ( 10.1002/adma.201304620) [DOI] [PubMed] [Google Scholar]
- 69.Polman A, Knight M, Garnett EC, Ehrler B, Sinke WC. 2016. Photovoltaic materials: present efficiencies and future challenges. Science 352, aad4424 ( 10.1126/science.aad4424) [DOI] [PubMed] [Google Scholar]
- 70.Kirchartz T, Kaienburg P, Baran D. 2018. Figures of merit guiding research on organic solar cells. J. Phys. Chem. C 122, 5829–5843. ( 10.1021/acs.jpcc.8b01598) [DOI] [Google Scholar]
- 71.Green MA. 1981. Solar cell fill factors: general graph and empirical expressions. Solid State Electron 24, 788–789. ( 10.1016/0038-1101(81)90062-9) [DOI] [Google Scholar]
- 72.Green MA. 1982. Accuracy of analytical expressions for solar-cell fill factors. Solar Cells 7, 337–340. ( 10.1016/0379-6787(82)90057-6) [DOI] [Google Scholar]
- 73.Miyata K, Meggiolaro D, Trinh MT, Joshi PP, Mosconi E, Jones SC, De Angelis F, Zhu XY. 2017. Large polarons in lead halide perovskites. Sci Adv 3, e1701217 ( 10.1126/sciadv.1701217) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Emin D. 2018. Barrier to recombination of oppositely charged large polarons. J. Appl. Phys. 123, 055105 ( 10.1063/1.5019834) [DOI] [Google Scholar]
- 75.Tong CJ, Li L, Liu LM, Prezhdo OV. 2018. Long carrier lifetimes in PbI2-rich perovskites rationalized by ab initio nonadiabatic molecular dynamics. ACS Energy Lett. 3, 1868–1874. ( 10.1021/acsenergylett.8b00961) [DOI] [Google Scholar]
- 76.Long R, Prezhdo OV. 2015. Dopants control electron (hole recombination at perovskite (TiO2 interfaces: ab initio time-domain study. ACS Nano 9, 11143–11155. ( 10.1021/acsnano.5b05843) [DOI] [PubMed] [Google Scholar]
- 77.Azarhoosh P, McKechnie S, Frost JM, Walsh A, van Schilfgaarde M. 2016. Research update: relativistic origin of slow electron-hole recombination in hybrid halide perovskite solar cells. APL Mater. 4, 091501 ( 10.1063/1.4955028) [DOI] [Google Scholar]
- 78.Ambrosio F, Wiktor J, De Angelis F, Pasquarello A. 2018. Origin of low electron-hole recombination rate in metal halide perovskites. Energ. Environ. Sci., 11, 101–105. ( 10.1039/C7EE01981E) [DOI] [Google Scholar]
- 79.Meggiolaro D, Motti SG, Mosconi E, Barker AJ, Ball J, Andrea Riccardo Perini C, Deschler F, Petrozza A, De Angelis F. 2018. Iodine chemistry determines the defect tolerance of lead-halide perovskites. Energ. Environ. Sci. 11, 702–713. ( 10.1039/C8EE00124C) [DOI] [Google Scholar]
- 80.Hutter EM, Gelvez-Rueda MC, Osherov A, Bulovic V, Grozema FC, Stranks SD, Savenije TJ. 2017. Direct-indirect character of the bandgap in methylammonium lead iodide perovskite. Nat. Mater. 16, 115–120. ( 10.1038/nmat4765) [DOI] [PubMed] [Google Scholar]
- 81.Wang T, Daiber B, Frost JM, Mann SA, Garnett EC, Walsh A, Ehrler B. 2017. Indirect to direct bandgap transition in methylammonium lead halide perovskite. Energ. Environ. Sci. 10, 509–515. ( 10.1039/C6EE03474H) [DOI] [Google Scholar]
- 82.Zhang X, Shen JX, Wang W, Van de Walle CG. 2018. First-principles analysis of radiative recombination in lead-halide perovskites. ACS Energy Lett. 3, 2329–2334. ( 10.1021/acsenergylett.8b01297) [DOI] [Google Scholar]
- 83.Kirchartz T, Rau U. 2017. Decreasing radiative recombination coefficients via an indirect band gap in lead halide perovskites. J. Phys. Chem. Lett. 8, 1265–1271. ( 10.1021/acs.jpclett.7b00236) [DOI] [PubMed] [Google Scholar]
- 84.Markvart T. 2003. Multiphonon recombination. In Recombination in semiconductors (ed. Landsberg PT.), pp. 467–468. Cambridge, UK: Cambridge University Press. [Google Scholar]
- 85.Ridley BK. 2013. Quantum processes in semiconductors. Oxford, UK: Oxford University Press. [Google Scholar]
- 86.Kirchartz T, Markvart T, Rau U, Egger DA. 2018. Impact of small phonon energies on the charge-carrier lifetimes in metal-halide perovskites. J. Phys. Chem. Lett. 9, 939–946. ( 10.1021/acs.jpclett.7b03414) [DOI] [PubMed] [Google Scholar]
- 87.Markvart T. 1981. Semiclassical theory of non-radiative transitions. J. Phys. C: Solid State Phys. 14, L895–L899. ( 10.1088/0022-3719/14/29/006) [DOI] [Google Scholar]
- 88.Ridley BK. 1978. On the multiphonon capture rate in semiconductors. Solid State Electron 21, 1319–1323. ( 10.1016/0038-1101(78)90200-9) [DOI] [Google Scholar]
- 89.Markvart T. 2013. Multiphonon recombination. In Recombination in semiconductors (ed. Landsberg PT.), p. 475 Cambridge, UK: Cambridge University Press. [Google Scholar]
- 90.Du MH. 2015. Density functional calculations of native defects in CH3NH3PbI3: effects of spin-orbit coupling and self-interaction error. J. Phys. Chem. Lett. 6, 1461–1466. ( 10.1021/acs.jpclett.5b00199) [DOI] [PubMed] [Google Scholar]
- 91.Uratani H, Yamashita K. 2017. Charge carrier trapping at surface defects of perovskite solar cell absorbers: a first-principles study. J. Phys. Chem. Lett. 8, 742–746. ( 10.1021/acs.jpclett.7b00055) [DOI] [PubMed] [Google Scholar]
- 92.Mayers MZ, Tan LZ, Egger DA, Rappe AM, Reichman DR. 2018. How lattice and charge fluctuations control carrier dynamics in halide perovskites. Nano Lett. 18, 8041–8046. ( 10.1021/acs.nanolett.8b04276) [DOI] [PubMed] [Google Scholar]
- 93.Benduhn J, et al. 2017. Intrinsic non-radiative voltage losses in fullerene-based organic solar cells. Nat. Energy 2, 17053 ( 10.1038/nenergy.2017.53) [DOI] [Google Scholar]
- 94.Azzouzi M, Yan J, Kirchartz T, Liu K, Wang J, Wu H, Nelson J. 2018. Nonradiative energy losses in bulk-heterojunction organic photovoltaics. Phys. Rev. X 8, 031055 ( 10.1103/PhysRevX.8.031055) [DOI] [Google Scholar]
- 95.Liu X, Li Y, Ding K, Forrest S. 2019. Energy loss in organic photovoltaics: nonfullerene versus fullerene acceptors. Phys. Rev. Appl. 11, 024060 ( 10.1103/PhysRevApplied.11.024060) [DOI] [Google Scholar]
- 96.Stolterfoht M, et al. 2018. The perovskite/transport layer interfaces dominate non-radiative recombination in efficient perovskite solar cells (arXiv:1810.01333) (https://arxiv.org/abs/1810.01333) [Google Scholar]
- 97.Aberle AG, Wenham SR, Green MA. 1993. A new method for accurate measurements of the lumped series resistance of solar cells , pp. 133–139. [Google Scholar]
- 98.Pysch D, Mette A, Glunz SW. 2007. A review and comparison of different methods to determine the series resistance of solar cells. Sol. Energy Mater. Sol. Cells 91, 1698–1706. ( 10.1016/j.solmat.2007.05.026) [DOI] [Google Scholar]
- 99.Müller TCM, Pieters BE, Rau U, Kirchartz T. 2013. Analysis of the series resistance in pin-type thin-film silicon solar cells. J. Appl. Phys. 113, 134503 ( 10.1063/1.4798393) [DOI] [Google Scholar]
- 100.Helbig A, Kirchartz T, Schaeffler R, Werner JH, Rau U. 2010. Quantitative electroluminescence analysis of resistive losses in Cu(In, Ga)Se(2) thin-film modules. Sol. Energy Mater. Sol. Cells 94, 979–984. ( 10.1016/j.solmat.2010.01.028) [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
This article has no additional data.