Next Article in Journal
Quiet Zone Enhancement for a Target Location Using an Improved Virtual Sensing Algorithm
Previous Article in Journal
Microstrip Patch Antenna Assisted Compact Dual Band Planar Crossover
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study of the Contributions of Donor and Acceptor Photoexcitations to Open Circuit Voltage in Bulk Heterojunction Organic Solar Cells

School of Engineering and Information Technology, Charles Darwin University, Darwin, NT 0909, Australia
*
Author to whom correspondence should be addressed.
Electronics 2017, 6(4), 75; https://doi.org/10.3390/electronics6040075
Submission received: 1 September 2017 / Revised: 28 September 2017 / Accepted: 28 September 2017 / Published: 3 October 2017

Abstract

:
One of the key parameters in determining the power conversion efficiency (PCE) of bulk heterojunction (BHJ) organic solar cells (OSCs) is the open circuit voltage ( V O C ) . The processes of exciting the donor and acceptor materials individually in a BHJ OSC are investigated and are found to produce two different expressions for V O C s . Using the contributions of electron and hole quasi-Fermi levels and charge carrier concentrations, the two different V O C expressions are derived as functions of the energetics of the donor and acceptor materials and the photo-generated charge carrier concentrations, and calculated for a set of donor-acceptor blends. The simultaneous excitation of both the donor and acceptor materials is also considered and the corresponding V O C , which is different from the above two, is derived. The V O C calculated from the photoexcitation of the donor is found to be somewhat comparable with that obtained from the photoexcitation of the acceptor in most combinations of the donor and acceptor materials considered here. It is also found that the V O C calculated from the simultaneous excitations of donor and acceptor in BHJ OSCs is also comparable with the other two V O C s . All three V O C s thus derived produce similar results and agree reasonably well with the measured values. All three V O C s depend linearly on the concentration of the photoexcited charge carriers and hence incident light intensity, which agrees with experimental results. The outcomes of this study are expected to help in finding materials that may produce higher V O C and hence enhanced PCE in BHJ OSCs.

1. Introduction

Organic solar cells (OSCs) based on bulk heterojunction (BHJ) structure are promising candidates for the generation of clean, affordable electricity owing to their low fabrication cost compared to inorganic solar cells [1,2]. Recent research efforts into the development of new donor-acceptor blends have led to a significant improvement in the power conversion efficiency (PCE) of OSCs, above 11% [3,4,5,6], which is gradually moving OSCs closer to commercialization. However, poor material properties such as low charge carrier mobilities, low dielectric constants, and poor band offsets have led to the low PCE of OSCs relative to their silicon counterparts [1]. Hence, optimization of these parameters is essential for improving the performance of OSCs. Considerable attention has recently been focused on exploring strategies for the further optimization of the PCE, guided by a thorough understanding of the fundamental mechanisms that govern the photovoltaic performance of OSCs [7]. The following four processes distinguish BHJ OSCs from their inorganic counterparts: (i) photon absorption and exciton formation; (ii) exciton diffusion to the donor-acceptor (D-A) interface; (iii) exciton dissociation into free charge carriers at the D-A interface; and (iv) charge transport and collection at their respective electrodes [8,9]. These processes should be efficient enough to achieve an enhanced PCE of BHJ OSCs.
One of the important parameters of BHJ OSCs is the open-circuit voltage V O C , which is the voltage at which the net current flowing through the device is zero. It is established that in BHJ OSCs, V O C depends on the energetic difference between the lowest unoccupied molecular orbital (LUMO) of the acceptor material ( E LUMO A ) and the highest occupied molecular orbital (HOMO) of the donor material ( E HOMO D ) , called the donor-acceptor effective energy gap ( E D A ) [10,11,12,13,14]. Until now, the most explored approach to increasing V O C of BHJ OSCs has been to increase the energy of the LUMO of the fullerene acceptor material or decrease the energy of the HOMO of the donor material, or both [10,11,12,13,14,15]. Scharber and co-workers [10] empirically defined V O C as:
q V O C = E LUMO A E HOMO D Δ E loss
where q is the elementary charge and Δ E loss is an empirical value denoting the energy losses occurring in transporting the charge carriers to the electrodes [16,17,18].
The validity of Equation (1) has been established for a number of donor-acceptor (fullerene) blends with a voltage loss Δ E loss ranging from 0.3−0.6 eV [19,20]. The question that may be asked is: what are the possible causes of the voltage loss Δ E loss and how can it be reduced? The origin of Δ E loss is still being debated, with several possible mechanisms such as bimolecular recombination, coulombic interactions, energetic disorder, etc., [12,16,17,18,21,22]. An important loss mechanism in BHJ OSCs is the bimolecular recombination of free charge carriers with its rate R given by [23,24]:
R = γ n p
where γ = γ p r e q ε o ε r ( μ e + μ h ) is the bimolecular recombination coefficient (m3s−1): γ p r e is the dimensionless reduction prefactor, ε 0 is the permittivity of free space, ε r is the dielectric constant, and μ e ( μ h ) is the electron (hole) mobility (m2/Vs) in the active organic layer.
By minimizing the nonradiative bimolecular recombination in BHJ OSCs [25], one can minimize the voltage loss and eventually increase V O C . Recent studies [26,27] have shown that V O C can be determined more accurately by measuring the bimolecular recombination rate. These studies have also considered the influence of light intensity on V O C by varying the carrier concentration due to photoexcitation. Lange et al. [28] studied the dependence of V O C on the charge carrier concentration in polymer:fullerene blends using bias amplified charge extraction (BACE) measurements. In addition, as stated above, the energy loss also occurs due to the energetic disorder, which reduces the V O C . A few recent studies have shown the effect of energetic disorder on V O C using a Gaussian density of states (DOS) [29,30,31]. Garcia et al. [29] developed a model to predict the V O C values by introducing DOS distributions of the HOMO in donor and the LUMO in acceptor materials. In their study, the energetic disorder is modeled by the Gaussian DOS with the mean energy equal to E LUMO A and standard deviation σ e for the acceptor material and corresponding E HOMO D and σ h for the donor material. In other studies [16,25,32,33,34,35,36], the role of energetics with respect to charge carrier dynamics and coulombic interactions have been analyzed. Credgington and Durrant [32] studied the dependence of V O C on the generation and recombination dynamics; they studied the dependence of V O C on the charge carrier concentration and carrier recombination lifetime. Vandewal et al. [25,33,34,35] derived a relationship between V O C and the energy of the charge transfer (CT) exciton state E C T by incorporating the energetic disorders. However, other works [16,36] have derived different expressions relating V O C and E C T in OSCs which exclude the influence of energetic disorder.
Despite numerous attempts to study V O C and the associated energetic loss mechanisms in BHJ OSCs, there appears to be few attempts being made to understand whether it depends on the concentration of photoexcited charge carriers in the donor and acceptor materials. This may be due to the fact that as the two materials, donor and acceptor, are blended together in a BHJ structure, they cannot be separately excited experimentally. However, in order to understand the contributions of photoexcitation of donor and acceptor materials to the performance of a BHJ OSC, one may seek to assess the performance considering the two cases separately: (1) when absorption and photoexcitation occur only in the donor and (2) when these occur only in the acceptor. Such a study can only be carried out theoretically, as presented here. When an exciton is excited in the donor in a BHJ OSC, its electron can be transferred from the donor LUMO to the acceptor LUMO at the D-A interface because the acceptor LUMO lies at a lower energy (see Figure 1a). In this process, an exciton becomes a CT exciton as its electron is transferred to the acceptor and its hole remains in the donor, as shown in Figure 1c. Likewise, when an exciton is excited in the acceptor, its hole can be transferred from the acceptor HOMO to the donor HOMO, being at a lower hole energy (see Figure 1b), which also creates a CT exciton with an electron in the acceptor LUMO and a hole in the donor HOMO, as shown in Figure 1c.
This has prompted several researchers to investigate the above two processes in more detail [37,38,39,40,41,42,43,44,45,46]. Recently, two different external quantum efficiencies (EQE) due to different donor and acceptor absorption bands have been reported [37]. The difference in the absorption bands can be attributed to the difference in the energy levels involved. The times of transfer of electrons from the donor LUMO to the acceptor LUMO and that of holes from the HOMO of the acceptor to the HOMO of the donor have been measured [38,39] and have been found to be different; the hole transfer was found to be faster than the electron transfer. This may be expected to lead to generating different concentrations n and p of electrons and holes. Different charge generations by exciting the donor and acceptor individually have also been reported in some donor-acceptor combinations [40], and different internal quantum efficiencies (IQE) have also been observed [41]. Some studies have demonstrated that BHJ OSCs based on thienothiophene-substituted diketopyrrolopyrrole (DPP) polymers (PDPP2TT-T) possess an EQE of 0.63–0.78 within the wavelength range of acceptor-fullerene absorption (400–650 nm), whereas a lower EQE of 0.35–0.50 is achieved within the range of donor-polymer absorption (650–900 nm) [41,42]. Such a difference may not be attributed to a difference in absorption of photons in the different materials, but may be due to the difference in charge generation. In another study, Baulin et al. [43] compared the dynamics of CT states generated through electron-transfer and hole-transfer exciton-dissociation pathways in three BHJ OSCs, and they found that the dynamics of CT state recombination are very similar for the two charge-generation processes, implying that the nature of the generated CT states is independent of the pathways of CT formation. They did, however, observe that the generation of CT states and free charge carriers greatly depends on the actual LUMO energy offset Δ E LUMO = E LUMO D E LUMO A (electron transfer) and the HOMO energy offset Δ E HOMO = E HOMO D E HOMO A (hole transfer) of the blended system [44]. These energy offsets are independent of the pathway of the formation of CT excitons. Singh et al. [45,46] recently studied the contribution of exciton generation in the donor and acceptor materials and its dissociation to the performance of BHJ OSCs. They found that the absorption and dissociation rates as well as exciton diffusion lengths are comparable whether excitons are generated in the donor or acceptor.
In this paper, we investigate in detail the difference in the open circuit voltage from the photoexcited donor and photoexcited acceptor separately in the following seven different polymer:fullerene blends: poly[2-methoxy-5-(3′,7′-dimethyloctyloxy)-1,4-phenylenevinylene] (MDMO-PPV), regioregular poly(3-hexylthiophene) (P3HT), poly[N-9′-heptadecanyl-2,7-carbazole-alt-5,5-(4′,7′-di-2-thienyl-2′,1′,3′-benzothiadiazole)] (PCDTBT), poly[2-(3,7-dimethoxyoctyloxy)-5-methoxy-1,4-phenylene vinylene] (OC1C10-PPV), poly[4,8-bis[(2-ethylhexyl)oxy]benzo[1,2-b:4,5-b′]dithiophene-2,6-diyl][3-fluoro-2-[(2-ethylhexyl)carbonyl]thieno[3,4-b]thiophenediyl] (PTB7), mono-DPP, and bis-DPP blended with different derivatives of methano-fullerene[6,6]-phenyl C61,-butyric acid methyl ester (PCBM). We derived two different expressions for V O C in BHJ OSCs. The first expression, which is a result of photoexcitation of the donor, depends primarily on the energetic difference between the hole quasi-Fermi level ( E F h ) and the HOMO of the donor material ( E F h E HOMO D ) and the electron concentration n . The second expression, which is a result of photoexcitation of the acceptor, depends primarily on the energetic difference between the LUMO and the electron quasi-Fermi level ( E F e ) of the acceptor material ( E LUMO A E F e ) and the hole concentration p . Thus, the two open circuit voltages calculated in each selected D-A blend are presented here and their influences on the performance of BHJ OSCs are discussed. Following this, the simultaneous photoexcitation of both the donor and acceptor materials is also considered and the corresponding V O C , which is different from the above two, is derived, and the calculated V O C s are compared with their corresponding measured values and a reasonably good agreement is found between them.

2. Theoretical Formalism

Under no illumination condition of equilibrium, assuming the condition of the open circuit voltage in a BHJ OSC, the electron and hole current densities are assumed to be independently zero, and under this condition the quasi-Fermi levels in the donor and acceptor materials are iso-energetic at energy E F throughout the device. Once such a BHJ OSC is illuminated, charge carriers are continuously generated in the active layer. As the concentration of charge carriers increases, the quasi-Fermi level E F splits into the hole and electron quasi-Fermi levels located, respectively, at energies E F h in the donor material and at E F e in the acceptor, as shown in Figure 2.
Under the condition of illumination, V O C in any BHJ OSC is given by [47]:
q V O C = E F e E F h
The Fermi energies E F e and E F h may be expected to be different for excitations of the donor and acceptor individually. Although the difference may not be expected to be significant, this may lead to two different values of V O C according to Equation (3).

2.1. Photoexcitation of the Donor

We consider the photoexcitation of only the donor material here, where an excited exciton at the D-A interface forms a CT exciton by transferring its electron to the acceptor’s LUMO, being at a lower energy [48]. The electron concentration ( n ) in the LUMO of the fullerene-acceptor material and the hole concentration ( p ) in the HOMO of the polymer-donor material are, respectively, given by the Boltzmann distribution as [49]:
n = N C exp [ ( E F e E LUMO A ) / k B T ]
p = N V exp [ ( E HOMO D E F h ) / k B T ]
where N C ( N V ) (m−3) is the effective density of states for the LUMO (HOMO) of the acceptor (donor) material, T is the temperature, and k B is the Boltzmann constant.
Under the conditions of V O C and bimolecular recombination, the generation rate of photoexcited electron and hole pairs G is equal to the bimolecular recombination rate given in Equation (2), which gives:
G = R = γ n p
Assuming a non-optimized BHJ OSC with unbalanced charge carrier mobilities ( μ e μ h ) , we assume that n p [39,40], and then divide Equation (6) by n 2 to obtain:
p n = G γ n 2
Dividing Equation (5) by Equation (4) and using it in Equation (7) gives:
G γ n 2 = N V N C exp [ 1 k B T ( E HOMO D + E LUMO A E F h E F e ) ]
Substituting E LUMO A = E HOMO D + E D A into Equation (8) and using Equation (3), we get:
k B T ln N C G N V γ n 2 = ( 2 E HOMO D 2 E F h q V O C + E D A )
Assuming N C = N V [38,41], and rearranging Equation (9), we obtain V O C related to the excitation of donor only as:
V O C D = 1 q ( E D A A + 2 k B T ln n )
where:
A = 2 ( E F h E HOMO D ) + k B T ln ( G / γ )

2.2. Photoexcitation of the Acceptor

Next, we consider the photoexcitation of only the acceptor material, where the excited exciton at the D-A interface forms a CT exciton by transferring its hole to the donor’s HOMO, being at a lower energy [48]. This gives a different open circuit voltage, which is derived as follows.
Here, we divide both sides of Equation (6) by p 2 to obtain:
n p = G γ p 2
Dividing Equation (4) by Equation (5) and using it in Equation (12) gives:
G γ p 2 = N C N V exp [ 1 k B T ( E F e E HOMO D E LUMO A + E F h ) ]
Substituting E HOMO D = E LUMO A E D A into Equation (13) and using Equation (3), we get:
k B T ln N C G N V γ p 2 = ( 2 E F e 2 E LUMO A q V O C + E D A )
Assuming N C = N V [38,41] and rearranging Equation (14), we obtain V O C for this case as:
V O C A = 1 q ( E D A B + 2 k B T ln p )
where:
B = 2 ( E LUMO A E F e ) + k B T ln ( G / γ )
We have thus derived two different expressions for the V O C ; the first one in Equation (10) is obtained when the donor material is excited and the second in Equation (15) is obtained when the acceptor is excited.

2.3. Simultaneous Photoexcitation of the Donor and Acceptor

The theoretical developments presented in Section 2.1 and Section 2.2, where only the donor or acceptor is excited, are mostly applicable in layered structures of donor and acceptor organic solar cells. However, in a BHJ OSC where the donor and acceptor are blended together in the active layer, both the donor and acceptor may be excited simultaneously. In this case, we obtain a different V O C as follows.
Multiplying Equations (4) and (5), we obtain:
n p = N C N V exp [ ( E F e E F h + E HOMO D E LUMO A ) / k B T ]
Substituting E D A = E HOMO D E LUMO A and using Equation (3) in Equation (17) we thus get:
V O C D A = 1 q [ E D A C + 2 k B T ln ( n p ) ]
where:
C = k B T ln N C N V
It may be noted that, unlike V O C s obtained in Equations (10) and (15), which depend on the electron ( n ) and hole ( p ) concentrations, respectively, Equation (18) depends on both n and p . The dependence of V O C in Equation (18) on n p is presented mainly to compare it with the results derived in Equations (10) and (15).

3. Results

For calculating the V O C using Equations (10), (15), and (18), we require some material-dependent input parameters. For this, as stated above, we chose seven donor-acceptor blends with known input parameters available in the literature, as listed in Table 1.
The values of γ for PCDTBT:PC71BM [58], mono-DPP:PC71BM [61], and bis-DPP:PC71BM [61] are taken from the literature, and those of the remaining four blends, MDMO-PPV:PC61BM P3HT:PC60BM, OC1C10-PPV:PC61BM, and PTB7:PC71BM, are calculated using the bimolecular recombination coefficient equation described in the text. In addition, we need the values of ( E F h E HOMO D ) and ( E LUMO A E F e ) , as well as the carrier concentrations n and p in each blend to calculate V O C from Equations (10), (15), and (18), which are also listed in Table 2 and Table 3. However, it may be noted that both n and p are only available for PTB7:PC71BM [62] as given in Table 2 and Table 3. In mono-DPP:PC71BM and bis-DPP:PC71BM, only the electron concentration n is known in the literature [63]. Consequently, to find the hole concentration p in these two blends, we used Equation (7) and calculated p ; these n and p thus obtained are listed in Table 2 and Table 3. We first used the n and p values determined above for PTB7:PC71BM, mono-DPP:PC71BM, and bis-DPP:PC71BM in Equations (4) and (5), and calculated their energetic distances ( E LUMO A E F e ) and ( E F h E HOMO D ) , respectively, at 300 K by assuming the effective density of states N C = N V = 1 × 1025 m−3 in each blend.
For the other four blends considered here, MDMO-PPV:PC61BM, P3HT:PC60BM, PCDTBT:PC71BM, and OC1C10-PPV:PC61BM, the n and p concentrations are not known. As the n and p concentrations in all the three materials are of the order of 1022 m−3 (see Table 2 and Table 3), we used the average values of n , which is 1.8 × 1022 m−3, and p , which is 0.95 × 1022 m−3, for PTB7:PC71BM, mono-DPP:PC71BM, and bis-DPP:PC71BM, and calculated the energetic distances ( E F h E HOMO D ) and ( E LUMO A E F e ) for the other four blends (see Table 2 and Table 3). That is to say, using Equations (4) and (5) with N C = N V = 2.5 × 1025 m−3 known for MDMO-PPV:PC61BM [17,57], N C = N V = 2 × 1026 m−3 for P3HT:PC60BM [1], N C = N V = 5 × 1025 m−3 for PCDTBT:PC71BM [52], and N C = N V = 2.5 × 1025 m−3 for OC1CI0-PPV:PC61BM [1,59], in addition to the corresponding average charge carrier concentrations n = 1.8 × 1022 m−3 and p = 0.95 × 1022 m−3, we calculated ( E LUMO A E F e ) and ( E F h E HOMO D ) , respectively, in the other four blends.
Using these parameters in Equations (10) and (15), the V O C is calculated in all seven blends for the photoexcitation of the donor and acceptor as given in Table 2 and Table 3, respectively. The calculated values of V O C from Equation (10) when the donor is excited in most selected materials are somewhat comparable with those calculated from Equation (15) when the acceptor is excited, as also shown in Figure 3. In addition, considering the case of simultaneous photoexcitation of donor and acceptor materials, the open circuit voltage in each blend is calculated from Equation (18) as given in Table 4. All the three V O C s obtained from the asynchronous and simultaneous photoexcitations of donor and acceptor are found to be comparable and agree well with the corresponding measured values [51,54,59,61,64,65].
The open circuit voltages V O C s calculated from Equations (10) and (15) are plotted as a function of the effective band gap E D A / q as shown in Figure 3, which illustrates that V O C calculated from Equation (10) is slightly higher than that from Equation (15).
Finally, as stated previously, in a BHJ OSC the donor and acceptor materials are blended together. Consequently, both the donor and acceptor are excited simultaneously, resulting in one open circuit voltage as derived in Equation (18), which is used to calculate the V O C D A for simultaneous photoexcitations of both the donor and acceptor in each of the selected blends, as listed in Table 4. The measured values of open circuit voltage V O C meas . are also given in Table 4 for comparison.

4. Discussions

A comprehensive study of the open circuit voltage in BHJ OSCs has been carried out. Three different expressions for V O C are derived; the first in Equation (10) is obtained when the donor material in a BHJ OSC is excited, the second in Equation (15) is obtained when the acceptor material is excited, and the third in Equation (18) is obtained when both the donor and acceptor materials are excited simultaneously. As stated previously, when the photoexcitation occurs in the donor, an electron is transferred from the LUMO of the donor to the LUMO of the acceptor, and conversely, when photoexcitation occurs in the acceptor, a hole is transferred from the HOMO of the acceptor to the HOMO of the donor [48]. The photoexcitation of the donor and acceptor individually produces comparable V O C values from Equations (10) and (15), as shown in Figure 3. Moreover, according to Figure 3, the linear correlation between V O C and E D A / q agrees well with previous studies [20,30,32,33,64]. The third open circuit voltage V O C D A calculated in Equation (18) from the simultaneous excitations of both the donor and acceptor, as given in Table 4, is also comparable with the other two V O C s . All three V O C s thus derived produce comparable results and agree reasonably well with the experimental values (see Table 2, Table 3 and Table 4). However, the calculated V O C values in Table 2, Table 3 and Table 4 are obtained from slightly different energetic parameters; as a result, we obtained three different sets of V O C s for the photoexcitation of the donor, photoexcitation of the acceptor, and simultaneous photoexcitation of both. However, experimentally, there is only one set of measured values of V O C to compare the theoretical results with. As a result, the calculated range of V O C for the materials considered here is 0.50–0.96 V, which is slightly different from the range of experimental values (0.52–0.85 V). Thus, as stated above, this discrepancy may be attributed to the different energies ( E F h E HOMO D ) / q used in the calculation of V O C from Equations (10), (15), and (18).
In addition, the two V O C s in Equations (10) and (15) depend linearly on ln n and ln p , respectively, and this agrees well with previously measured results [28,63]. This also holds for V O C derived from Equation (18), which depends on ln ( n p ) . By setting n = p in Equations (10), (15), and (18), we found that all three V O C s increase linearly with ln n and have the same slope equal to k B T q / log ( exp ( 1 ) ) 2.3 × 2 k B T / q 119   mV . This implies that the materials with higher photo-generated charge carrier concentration will have higher V O C in comparison with materials with lower carrier concentration.
It is also worth noting that the dependence of V O C on the incident light intensity has been studied [17,24]. Substituting n p = G γ I a b s = I o [ 1 exp ( α t ) ] into Equation (18), where I a b s is the absorbed light intensity, I 0 is the incident light intensity, α is the absorption coefficient, and t is the thickness of the active layer, we get:
V O C D A = 1 q [ E D A + k B T ln I 0 + constant ]
The V O C in Equation (20) depends linearly on ln I 0 with a slope of k B T q , which agrees perfectly well with the plotted V O C as a function of ln I 0 in [17,24].
Orlowski et al. [66] used the energetics of quasi-Fermi levels to study V O C in a heterojunction solar cell. They found that for a particular donor (p-type)-acceptor (n-type) heterojunction, the electron concentration ( n ) in the conduction band (LUMO) of the acceptor and the hole concentration ( p ) in the valence band (HOMO) of the donor, coupled with their respective quasi-Fermi levels, can create open circuit voltages like two independent cells. This agrees very well with our concept of deriving two different V O C s in Equations (10) and (15), although both the expressions produce similar results.
Furthermore, we find that the V O C D lies 0.34–0.49 V below the E D A / q in column 5 of Table 2 when the donor is excited, comparable to 0.34–0.48 V below the E D A / q in column 5 of Table 3 when the acceptor is excited. This implies that both the donor and acceptor photoexcitations can contribute to the achievable open circuit voltage as well as the photovoltaic performance of BHJ OSCs; this agrees with our earlier results [45]. The voltage offsets ( E D A q V O C j ) (where j = D , A o r D A ) of the various blends listed in Table 2, Table 3 and Table 4, ranging from 0.34–0.50 V, agree well with the experimental values measured at room temperature in the range of 0.25–0.48 V for small molecule OSCs [67], 0.32 V for polymer:fullerene solar cells of indacenoedithiophene (IDT) polymer [68], 0.34–0.44 V for a range of donor-acceptor blends [69], 0.38 V for OC1C10-PPV:PCBM solar cells [12], and 0.30–0.60 V for BHJ OSCs [20]. Other low voltage offsets ( E D A q V O C j ) that have been reported recently are 0.25 V for an evaporated bilayer OSC [70], 0.23 V and 0.26 V for diketopyrrolopyrrole-thieno[2,3-f]benzofuran (DTD):PC60BM and DTD:naphthalene diimide acceptor-polymer (N2200) systems, respectively [71], and 0.34–0.40 V for BHJ OSCs [72], with which our calculated values also somewhat agree in the range.
Overall, it may be deduced from our results that separate and simultaneous photoexcitations of the donor and acceptor produce comparable V O C s , and the highest loss in V O C is found to be contributed by the energetics of the donor and acceptor materials. Therefore, to achieve a high V O C , it is necessary to choose higher efficiency polymer-donor and acceptor materials, as well as control the donor and acceptor material energetics, which may lead to improvements in the PCEs of BHJ OSCs [32].

5. Conclusions

In summary, we have derived two different expressions for the calculation of the V O C in BHJ OSCs; one is used when the exciton is excited in the donor material and the other is employed when the exciton is excited in the acceptor material. We have also derived an expression for V O C when both the donor and acceptor materials are excited simultaneously, and its calculated values agree better with their measured values. All three V O C s obtained from separate and simultaneous excitations are found to be comparable, implying that the contributions of both donor and acceptor photoexcitations are comparable in BHJ OSCs. Also, it is found that materials that generate higher photo-generated charge carrier concentrations and have lower energetic distances of ( E F h E HOMO D ) or ( E LUMO A E F e ) may lead to higher V O C . Invention and application of such materials in the fabrication of BHJ OSCs may lead to higher V O C s and hence enhanced PCEs.

Author Contributions

Both the authors have contributed to the publication of this paper. Douglas Yeboah carried out all the calculations. With the guidance from Jai Singh, the paper was initially drafted by Yeboah and then was finalized by them together.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Inche Ibrahim, M.L.; Ahmad, Z.; Sulaiman, K.; Muniandy, S.V. Combined influence of carrier mobility and dielectric constant on the performance of organic bulk heterojunction solar cells. AIP Adv. 2014, 4, 057133. [Google Scholar] [CrossRef]
  2. Zhong, Y.; Tada, A.; Izawa, S.; Hashimoto, K.; Tajima, K. Enhancement of V O C without loss of J S C in organic solar cells by modification of donor/acceptor interfaces. Adv. Energy Mater. 2014, 4, 1301332. [Google Scholar] [CrossRef]
  3. Chen, C.-C.; Chang, W.-H.; Yoshimura, K.; Ohya, K.; You, J.; Gao, J.; Hong, Z.; Yang, Y. An efficient triple-junction polymer solar cell having a power conversion efficiency exceeding 11%. Adv. Mater. 2014, 26, 5670–5677. [Google Scholar] [CrossRef] [PubMed]
  4. Zhao, J.; Li, Y.; Yang, G.; Jiang, K.; Lin, H.; Ade, H.; Ma, W.; Yan, H. Efficient organic solar cells processed from hydrocarbon solvents. Nat. Energy 2016, 1, 15027. [Google Scholar] [CrossRef]
  5. Zhang, S.; Ye, L.; Hou, J. Breaking the 10% efficiency barrier in organic photovoltaics: Morphology and device optimization of well-known PBDTTT polymers. Adv. Energy Mater. 2016, 6, 1502529. [Google Scholar] [CrossRef]
  6. Gan, Q.; Bartoli, F.J.; Kafafi, Z.H. Plasmonic-enhanced organic photovoltaics: Breaking the 10% efficiency barrier. Adv. Mater. 2013, 25, 2385–2396. [Google Scholar] [CrossRef] [PubMed]
  7. Deibel, C.; Wagenpfahl, A.; Dyakonov, V. Influence of charge carrier mobility on the performance of organic solar cells. Phys. Status Solidi 2008, 2, 175–177. [Google Scholar] [CrossRef]
  8. Narayan, R.M.; Singh, J. Study of the mechanism and rate of exciton dissociation at the donor-acceptor interface in bulk-heterojunction organic solar cells. J. Appl. Phys. 2013, 114. [Google Scholar] [CrossRef]
  9. Ompong, D.; Singh, J. Diffusion length and Langevin recombination of singlet and triplet excitons in organic heterojunction solar cells. ChemPhysChem 2015, 16, 1281–1285. [Google Scholar] [CrossRef] [PubMed]
  10. Scharber, M.C.; Muhlbacher, D.; Koppe, M.; Denk, P.; Waldauf, C.; Heeger, A.J.; Brabec, C.J. Design rules for donors in bulk-heterojunction solar cells—Towards 10% energy-conversion efficiency. Adv. Mater. 2006, 18, 789–794. [Google Scholar] [CrossRef]
  11. Kietzke, T.; Egbe, D.A.M.; Hörhold, H.-H.; Neher, D. Comparative study of M3EH−PPV-based bilayer photovoltaic devices. Macromolecules 2006, 39, 4018–4022. [Google Scholar] [CrossRef]
  12. Mihailetchi, V.D.; Blom, P.W.M.; Hummelen, J.C.; Rispens, M.T. Cathode dependence of the open-circuit voltage of polymer:fullerene bulk heterojunction solar cells. J. Appl. Phys. 2003, 94, 6849–6854. [Google Scholar] [CrossRef]
  13. Ross, R.B.; Cardona, C.M.; Guldi, D.M.; Sankaranarayanan, S.G.; Reese, M.O.; Kopidakis, N.; Peet, J.; Walker, B.; Bazan, G.C.; van Keuren, E.; et al. Endohedral fullerenes for organic photovoltaic devices. Nat. Mater. 2009, 8, 208–212. [Google Scholar] [CrossRef] [PubMed]
  14. Lenes, M.; Wetzelaer, G.; Kooistra, F.B.; Veenstra, S.C.; Hummelen, J.C.; Blom, P.W.M. Fullerene bisadducts for enhanced open-circuit voltages and efficiencies in polymer solar cells. Adv. Mater. 2008, 20, 2116–2119. [Google Scholar] [CrossRef]
  15. Fengling, Z.; Bijleveld, J.; Perzon, E.; Tvingstedt, K.; Barrau, S.; Inganäs, O.; Andersson, M.R. High photovoltage achieved in low band gap polymer solar cells by adjusting energy levels of a polymer with the LUMOs of fullerene derivatives. J. Mater. Chem. 2008, 18, 5468–5474. [Google Scholar] [CrossRef]
  16. Rand, B.; Burk, D.; Forrest, S. Offset energies at organic semiconductor heterojunctions and their influence on the open-circuit voltage of thin-film solar cells. Phys. Rev. B 2007, 75, 115327. [Google Scholar] [CrossRef]
  17. Koster, L.J.A.; Mihailetchi, V.D.; Ramaker, R.; Blom, P.W.M. Light intensity dependence of open-circuit voltage of polymer:fullerene solar cells. Appl. Phys. Lett. 2005, 86, 123509. [Google Scholar] [CrossRef]
  18. Riede, M.; Mueller, T.; Tress, W.; Schueppel, R.; Leo, K. Small-molecule solar cells-status and perspectives. Nanotechnology 2008, 19, 424001. [Google Scholar] [CrossRef] [PubMed]
  19. Brabec, C.J.; Cravino, A.; Meissner, D.; Sariciftci, N.S.; Fromherz, T.; Rispens, M.T.; Sanchez, L.; Hummelen, J.C. Origin of the open circuit voltage of plastic solar cells. Adv. Funct. Mater. 2001, 11, 374–380. [Google Scholar] [CrossRef]
  20. Yang, Q.-D.; Li, H.-W.; Cheng, Y.; Guan, Z.; Liu, T.; Ng, T.-W.; Lee, C.-S.; Tsang, S.-W. Probing the energy level alignment and the correlation with open-circuit voltage in solution-processed polymeric bulk heterojunction photovoltaic devices. ACS Appl. Mater. Interfaces 2016, 8, 7283–7290. [Google Scholar] [CrossRef] [PubMed]
  21. Cho, N.; Schlenker, C.W.; Knesting, K.M.; Koelsch, P.; Yip, H.-Y.; Ginger, D.S.; Jen, A.K.-Y. High dielectric constant side-chain polymers show reduced non-geminate recombination in heterojunction. Adv. Energy Mater. 2014, 4, 1301857. [Google Scholar] [CrossRef]
  22. Chen, S.; Tsang, S.W.; Lai, T.H.; Reynolds, J.R.; So, F. Dielectric effect on the photovoltage loss in organic photovoltaic cells. Adv. Mater. 2014, 26, 6125–6131. [Google Scholar] [CrossRef] [PubMed]
  23. Koster, L.J.A.; Mihailetchi, V.D.; Blom, P.W.M. Bimolecular recombination in polymer/fullerene bulk heterojunction solar cells. Appl. Phys. Lett. 2006, 88, 052104. [Google Scholar] [CrossRef]
  24. Cowan, S.R.C.; Roy, A.; Heeger, A.J. Recombination in polymer-fullerene bulk heterojunction solar cells. Phys. Rev. B 2010, 82, 245207. [Google Scholar] [CrossRef]
  25. Vandewal, K.; Tvingstedt, K.; Gadisa, A.; Inganas, O.; Manca, J.V. On the origin of the open-circuit voltage of polymer-fullerene solar cells. Nat. Mater. 2009, 8, 904–909. [Google Scholar] [CrossRef] [PubMed]
  26. Shuttle, C.G.; O’Regan, B.; Ballantyne, A.M.; Nelson, J.; Bradley, D.D.C.; Durrant, J.R. Bimolecular recombination losses in polythiophene: Fullerene solar cells. Phys. Rev. B 2008, 78, 113201. [Google Scholar] [CrossRef]
  27. Maurano, A.; Hamilton, R.; Shuttle, C.G.; Ballantyne, A.M.; Nelson, J.; O’Regan, B.; Zhang, W.; McCulloch, I.; Azimi, H.; Morana, M.; et al. Recombination dynamics as a key determinant of open circuit voltage in organic bulk heterojunction solar cells: A comparison of four different donor polymers. Adv. Mater. 2010, 22, 4987–4992. [Google Scholar] [CrossRef] [PubMed]
  28. Lange, I.; Kniepert, J.; Pingel, P.; Dumsch, I.; Allard, S.; Janietz, S.; Scherf, U.; Neher, D. Correlation between the open-circuit voltage and the energetics of organic bulk heterojunction solar cells. J. Phys. Chem. Lett. 2013, 4, 3865–3871. [Google Scholar] [CrossRef]
  29. Garcia-Belmonte, G.; Boix, P.P.; Bisquert, J.; Lenes, M.; Bolink, H.J.; La Rosa, A.; Filippone, S.; Martín, N. Influence of the intermediate density-of-states occupancy on open-circuit voltage of bulk heterojunction solar cells with different fullerene acceptors. J. Phys. Chem. Lett. 2010, 1, 2566–2571. [Google Scholar] [CrossRef]
  30. Garcia-Belmonte, G.; Juan Bisquert, J. Open-circuit voltage limit caused by recombination through tail states in bulk heterojunction polymer-fullerene solar cells. Appl. Phys. Lett. 2010, 96, 113301. [Google Scholar] [CrossRef]
  31. Blakesley, J.C.; Greenham, N.C. Charge transfer at polymer-electrode interfaces: The effect of energetic disorder and thermal injection on band bending and open-circuit voltage. J. Appl. Phys. 2009, 106, 034507. [Google Scholar] [CrossRef]
  32. Credgington, D.; Durrant, J.R. Insights from transient optoelectronic analyses on the open-circuit voltage of organic solar cells. J. Phys. Chem. Lett. 2012, 3, 1465–1478. [Google Scholar] [CrossRef] [PubMed]
  33. Vandewal, K.; Gadisa, A.; Oosterbaan, W.D.; Bertho, S.; Banishoeib, F.; Van Severen, I.; Lutsen, L.; Cleij, T.J.; Vanderzande, D.; Manca, J.V. The relation between open-circuit voltage and the onset of photocurrent generation by charge-transfer absorption in polymer: Fullerene bulk heterojunction solar cells. Adv. Funct. Mater. 2008, 18, 2064–2070. [Google Scholar] [CrossRef]
  34. Vandewal, K.; Tvingstedt, K.; Gadisa, A.; Inganäs, O.; Manca, J.V. Relating the open-circuit voltage to interface molecular properties of donor:acceptor bulk heterojunction solar cells. Phys. Rev. B 2010, 81, 125204. [Google Scholar] [CrossRef]
  35. Piersimoni, F.; Chambon, S.; Vandewal, K.; Mens, R.; Boonen, T.; Gadisa, A.; Izquierdo, M.; Filippone, S.; Ruttens, B.; D’Haen, J.; et al. Influence of fullerene ordering on the energy of the charge-transfer state and open-circuit voltage in polymer:fullerene solar cells. J. Phys. Chem. C 2011, 115, 10873–10880. [Google Scholar] [CrossRef]
  36. Giebink, N.C.; Wiederrecht, G.P.; Wasielewski, M.R.; Forrest, S.R. Ideal diode equation for organic heterojunctions. I. Derivation and application. Phys. Rev. B 2010, 82, 155305. [Google Scholar] [CrossRef]
  37. Hendriks, K.H.; Wijpkema, A.S.G.; van Franeker, J.J.; Wienk, M.M.; Janssen, R.A.J. Dichotomous role of exciting the donor or the acceptor on charge generation in organic solar cells. J. Am. Chem. Soc. 2016, 138, 10026–10031. [Google Scholar] [CrossRef] [PubMed]
  38. Cook, S.; Katoh, R.; Furube, A. Ultrafast studies of charge generation in PCBM:P3HT blend films following excitation of the fullerene PCBM. J. Phys. Chem. C 2009, 113, 2547–2552. [Google Scholar] [CrossRef]
  39. Bakulin, A.A.; Hummelen, J.C.; Pshenichnikov, M.S.; van Loosdrecht, P.H.M. Ultrafast hole-transfer dynamics in polymer/PCBM bulk heterojunctions. Adv. Funct. Mater. 2010, 20, 1653–1660. [Google Scholar] [CrossRef]
  40. Ren, G.; Schlenker, C.W.; Ahmed, E.; Subramaniyan, S.; Olthof, S.; Kahn, A.; Ginger, D.S.; Jenekhe, S.A. Photoinduced hole transfer becomes suppressed with diminished driving force in polymer-fullerene solar cells while electron transfer remains active. Adv. Funct. Mater. 2013, 23, 1238–1249. [Google Scholar] [CrossRef]
  41. Armin, A.; Kassal, I.; Shaw, P.E.; Hambsch, M.; Stolterfoht, M.; Lyons, D.M.; Li, J.; Shi, Z.; Burn, P.L.; Meredith, P.J. Spectral dependence of the internal quantum efficiency of organic solar cells: Effect of charge generation pathways. Am. Chem. Soc. 2014, 136, 11465–11472. [Google Scholar] [CrossRef] [PubMed]
  42. Li, W.; Hendriks, K.H.; Furlan, A.; Roelofs, W.S.C.; Wienk, M.M.; Janssen, R.A.J.J. Efficient tandem and triple-junction polymer solar cells. Am. Chem. Soc. 2013, 135, 18942–18948. [Google Scholar] [CrossRef] [PubMed]
  43. Bakulin, A.A.; Dimitrov, S.D.; Rao, A.; Philip, C.Y.; Chow, P.C.Y.; Christian, B.; Nielsen, C.B.; Bob, C.; Schroeder, B.C.; Iain McCulloch, J.; et al. Charge-transfer state dynamics following hole and electron transfer in organic photovoltaic devices. J. Phys. Chem. Lett. 2013, 4, 209–215. [Google Scholar] [CrossRef] [PubMed]
  44. Yeboah, D.; Singh, J. Dependence of exciton diffusion length and diffusion coefficient on photophysical parameters in bulk heterojunction organic solar cells. J. Electron. Mater. 2017. [Google Scholar] [CrossRef]
  45. Narayan, M.; Singh, J. Photovoltaic contribution of photo-generated excitons in acceptor material of organic solar cells. J. Mater. Sci. Mater. Electron. 2017, 28, 7070–7076. [Google Scholar] [CrossRef]
  46. Singh, J.; Narayan, M.; Ompong, D. Comparative contributions of singlet and triplet excitons in the performance of organic devices. Phys. Status Solidi C 2016, 13, 77–80. [Google Scholar] [CrossRef]
  47. Gregg, B.A. Excitonic Solar Cells. J. Phys. Chem. B 2003, 107, 4688–4698. [Google Scholar] [CrossRef]
  48. Singh, J.; Narayan, M.; Ompong, D.; Zhu, F. Dissociation of charge transfer excitons at the donor-acceptor interface in bulk heterojunction organic solar cells. J. Mater. Sci. Mater. Electron. 2017, 28, 7095–7099. [Google Scholar] [CrossRef]
  49. Wagenpfahl, A.; Rauh, D.; Binder, M.; Deibel, C.; Dyakonov, V. S-shaped current-voltage characteristics of organic solar devices. Phys. Rev. B 2010, 82, 115306. [Google Scholar] [CrossRef]
  50. Koster, L.J.A.; van Strien, W.J.; Beek, W.J.E.; Blom, P.W.M. Device operation of conjugated polymer/zinc oxide bulk heterojunction solar cells. Adv. Funct. Mater. 2007, 17, 1297–1302. [Google Scholar] [CrossRef]
  51. Qia, B.; Wang, J. Open-circuit voltage in organic solar cells. J. Mater. Chem. 2012, 22, 24315–24325. [Google Scholar] [CrossRef]
  52. Cowan, S.R.; Leong, W.L.; Banerji, N.; Dennler, G.; Heeger, A.J. Identifying a threshold impurity level for organic solar cells: Enhanced first-order recombination via well-defined PC84BM traps in organic bulk heterojunction solar cells. Adv. Funct. Mater. 2011, 21, 3083–3092. [Google Scholar] [CrossRef]
  53. Meijer, E.J.; de Leeuw, D.M.; Setayesh, S.; van veenendaal, E.; Huisman, B.-H.; Blom, P.W.M.; Hummelen, J.C.; Scherf, U.; Klapwijk, T.M. Solution-processed ambipolar organic field-effect transistors and inverters. Nat. Mater. 2003, 2, 678–682. [Google Scholar] [CrossRef]
  54. Ebenhoch, B.; Thomson, S.A.J.; Genevicius, K.; Juška, G.; Samuel, I.D.W. Charge carrier mobility of the organic photovoltaic materials PTB7 and PC71BM and its influence on device performance. Org. Electron. 2015, 22, 62–68. [Google Scholar] [CrossRef] [Green Version]
  55. Lin, J.D.A.; Liu, J.; Kim, C.; Tamayo, A.B.; Proctor, C.M.; Nguyen, T.-Q. Effect of structural variation on photovoltaic characteristics of phenyl substituted diketopyrrolopyrroles. RSC Adv. 2014, 4, 14101–14108. [Google Scholar] [CrossRef]
  56. Walker, B.; Tamayo, A.B.; Dang, X.; Zalar, P.; Seo, J.H.; Garcia, A.; Tantiwiwat, M.; Nguyen, T. Nanoscale phase separation and high photovoltaic efficiency in solution-processed, small-molecule bulk heterojunction solar cells. Adv. Funct. Mater. 2009, 19, 3063–3069. [Google Scholar] [CrossRef]
  57. Mandoc, M.M.; Koster, L.J.A.; Blom, P.W.M. Optimum charge carrier mobility inorganic solar cells. Appl. Phys. Lett. 2007, 90, 133504. [Google Scholar] [CrossRef]
  58. Philippa, B.; Stolterfoht, M.; Burn, P.L.; Juŝka, G.; Meredith, P.; Whitel, R.D.; Pivrikas, A. The impact of hot charge carrier mobility on photocurrent losses in polymer-based solar cells. Sci. Rep. 2014, 4, 5695. [Google Scholar] [CrossRef] [PubMed]
  59. Koster, L.J.A.; Smits, E.C.P.; Mihailetchi, V.D.; Blom, P.W.M. Device model for the operation of polymer/fullerene bulk heterojunction solar cells. Phys. Rev. B 2005, 72, 085205. [Google Scholar] [CrossRef]
  60. Würfel, U.; Neher, D.; Spies, A.; Albrecht, S. Impact of charge transport on current voltage characteristics and power-conversion efficiency of organic solar cells. Nat. Commun. 2011, 6, 6951. [Google Scholar] [CrossRef] [PubMed]
  61. Proctor, C.M.; Kim, C.; Neher, D.; Nguyen, T.Q. Nongeminate recombination and charge transport limitations in diketopyrrolopyrrole-based solution-processed small molecule solar cells. Adv. Funct. Mater. 2013, 23, 3584–3594. [Google Scholar] [CrossRef]
  62. Oosterhout, D.S.; Ferguson, A.J.; Larson, B.W.; Olson, D.C.; Kopidakis, N. Modeling the free carrier recombination kinetics in PTB7:PCBM organic photovoltaics. J. Phys. Chem. C 2016, 120, 24597–24604. [Google Scholar] [CrossRef]
  63. Collins, S.A.; Proctor, C.M.; Ran, N.A.; Nguyen, T.-Q. Understanding open-circuit voltage loss through the density of states in organic bulk heterojunction solar cells. Adv. Energy Mater. 2016, 6, 1501721. [Google Scholar] [CrossRef]
  64. Cravino, A. Origin of the open circuit voltage of donor-acceptor solar cells: Do polaronic energy levels play a role? Appl. Phys. Phys. Lett. 2007, 91, 243502. [Google Scholar] [CrossRef]
  65. Sun, Y.; Takacs, C.J.; Cowan, S.R.; Seo, J.H.; Gong, X.; Roy, A.; Heeger, A.J. Efficient, air-stable bulk heterojunction polymer solar cells using MoO(x) as the anode interfacial layer. Adv. Mater. 2011, 23, 2226–2230. [Google Scholar] [CrossRef] [PubMed]
  66. Orlowski, B.A.; Pieniazek, A.; Goscinski, K.; Kopalko, K. Quasi Fermi levels in semiconductor photovoltaic heterojunction. Acta Phys. Pol. A 2016, 129. [Google Scholar] [CrossRef]
  67. Tuladhar, S.M.; Azzouzi, M.; Delval, F.; Yao, J.; Guilbert, A.A.Y.; Kirchartz, T.; Montcada, N.F.; Dominguez, R.; Langa, F.; Palomares, E.; et al. Low open-circuit voltage loss in solution-processed small-molecule organic solar cells. ACS Energy Lett. 2016, 1, 302–308. [Google Scholar] [CrossRef]
  68. Baran, D.; Vezie, M.S.; Gasparini, N.; Deledalle, F.; Yao, J.; Schroeder, B.C.; Bronstein, H.; Ameri, T.; Kirchartz, T.; McCulloch, I.; et al. Role of polymer fractionation in energetic losses and charge carrier lifetimes of polymer: Fullerene solar cells. J. Phys. Chem. C 2015, 119, 19668–19673. [Google Scholar] [CrossRef]
  69. Yao, J.; Kirchartz, T.; Vezie, M.S.; Faist, M.A.; Gong, W.; He, Z.; Wu, H.; Troughton, J.; Watson, T.; Bryant, D.; et al. Quantifying losses in open-circuit voltage in solution-processable solar cells. Phys. Rev. Appl. 2015, 4, 014020. [Google Scholar] [CrossRef]
  70. Bartynski, A.N.; Gruber, M.; Das, S.; Rangan, S.; Mollinger, S.; Trinh, C.; Bradforth, S.E.; Vandewal, K.; Salleo, A.; Bartynski, R.A.; et al. Symmetry-breaking charge transfer in a zinc chlorodipyrrin acceptor for high open circuit voltage organic photovoltaics. J. Am. Chem. Soc. 2015, 137, 5397–5405. [Google Scholar] [CrossRef] [PubMed]
  71. Tang, Z.; Liu, B.; Melianas, A.; Bergqvist, J.; Tress, W.; Bao, Q.; Qian, D.; Inganäs, O.; Zhang, F. A new fullerene-free bulk-heterojunction system for efficient high-voltage and high-fill factor solution-processed organic photovoltaics. Adv. Mater. 2015, 27, 1900–1907. [Google Scholar] [CrossRef] [PubMed]
  72. Ran, N.A.; Kuik, M.; Love, J.A.; Proctor, C.M.; Nagao, I.; Bazan, G.C.; Nguyen, T.-Q. Understanding the charge-transfer state and singlet exciton emission from solution-processed small-molecule organic solar cells. Adv. Mater. 2014, 26, 7405–7412. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representations of (a) the photoexcitation of the donor and the transfer of an electron to the acceptor at the interface; (b) the photoexcitation of the acceptor and the transfer of a hole to the donor at the interface; and (c) the formation of charge transfer excitons (CT) through both (a,b) processes in bulk heterojunction (BHJ) organic solar cells (OSCs).
Figure 1. Schematic representations of (a) the photoexcitation of the donor and the transfer of an electron to the acceptor at the interface; (b) the photoexcitation of the acceptor and the transfer of a hole to the donor at the interface; and (c) the formation of charge transfer excitons (CT) through both (a,b) processes in bulk heterojunction (BHJ) organic solar cells (OSCs).
Electronics 06 00075 g001
Figure 2. Schematic representation of the different energy levels, quasi-Fermi energy levels, and V O C in a BHJ OSC with a donor-acceptor blend under the open circuit condition.
Figure 2. Schematic representation of the different energy levels, quasi-Fermi energy levels, and V O C in a BHJ OSC with a donor-acceptor blend under the open circuit condition.
Electronics 06 00075 g002
Figure 3. Open circuit voltage V O C plotted as a function of effective band gap E D A / q for non-optimized P3HT:PC60BM with large photocarrier mobility mismatch ( μ e μ h ) , at energetic distances ( E F h E HOMO D ) / q = 0.26 V with n = 1.80 × 1022 m−3 and ( E LUMO A E F e ) / q = 0.24 V with p = 0.95 × 1022 m−3 in the donor and acceptor materials, respectively.
Figure 3. Open circuit voltage V O C plotted as a function of effective band gap E D A / q for non-optimized P3HT:PC60BM with large photocarrier mobility mismatch ( μ e μ h ) , at energetic distances ( E F h E HOMO D ) / q = 0.26 V with n = 1.80 × 1022 m−3 and ( E LUMO A E F e ) / q = 0.24 V with p = 0.95 × 1022 m−3 in the donor and acceptor materials, respectively.
Electronics 06 00075 g003
Table 1. List of input parameters required for calculating V O C from Equations (10), (15), and (18): E D A / q = ( E LUMO A E HOMO D ) / q , generation rate of bound polaron pairs ( G ) and bimolecular recombination coefficient ( γ ) , electron (hole) mobility μ e ( μ h ) , and dielectric constant ε r of each donor-acceptor (D-A) blend.
Table 1. List of input parameters required for calculating V O C from Equations (10), (15), and (18): E D A / q = ( E LUMO A E HOMO D ) / q , generation rate of bound polaron pairs ( G ) and bimolecular recombination coefficient ( γ ) , electron (hole) mobility μ e ( μ h ) , and dielectric constant ε r of each donor-acceptor (D-A) blend.
Active Layer Blend E D A q (V) G (1027m−3s−1) γ (10−17 m3s−1) μ e (m2V−1s−1) μ h (m2V−1s−1) ε r
MDMO-PPV:PC61BM1.30 12.7 85.732 × 10−7 82 × 10−8 83.4 8
P3HT:PC60BM1.00 26.25 27.081 × 10−7 21 × 10−8 23.4 2
PCDTBT:PC71BM1.20 31.0 90.12.9 × 10−7 93.0 × 10−9 93.4 9
OC1C10-PPV:PC61BM1.30 42.7 107.302.5 × 10−7 103.0 × 10−8 103.4 10
PTB7:PC71BM1.09 510 115.911 × 10−7 112.0 × 10−8 113.5 11
mono-DPP:PC71BM1.16 64.99 135.31 × 10−7 122.0 × 10−9 124.0 12
bis-DPP:PC71BM1.20 74.12 142.61.5 × 10−7 123.4 × 10−8 124.0 12
1 [50]; 2 [51]; 3 [52]; 4 [53]; 5 [54]; 6 [55]; 7 [56]; 8 [57]; 9 [58]; 10 [59]; 11 [60]; 12 [61].
Table 2. The calculated values of ( E F h E HOMO D ) / q and electron concentration n used in calculating the open circuit voltage V O C D from Equation (10), the calculated voltage offset ( E D A q V O C D ) for comparison with other known results, and the corresponding measured open circuit voltage V O C meas . .
Table 2. The calculated values of ( E F h E HOMO D ) / q and electron concentration n used in calculating the open circuit voltage V O C D from Equation (10), the calculated voltage offset ( E D A q V O C D ) for comparison with other known results, and the corresponding measured open circuit voltage V O C meas . .
Active Layer Blend ( E F h E HOMO D ) / q (V) n (m−3) V O C D (V) ( E D A q V O C D ) (V) V O C meas . (V)
MDMO-PPV:PC61BM0.201.80 × 1022 0.950.350.83 [64]
P3HT:PC60BM0.261.80 × 1022 0.510.490.63 [51]
PCDTBT:PC71BM0.221.80 × 1022 0.730.470.85 [65]
OC1C10-PPV:PC61BM0.201.80 × 1022 0.960.340.85 [59]
PTB7:PC71BM0.181.90 × 1022 0.750.340.75 [54]
mono-DPP:PC71BM0.202.50 × 1022 0.810.350.78 [61]
bis-DPP:PC71BM0.171.00 × 10220.850.350.52 [61]
Table 3. The calculated values of ( E LUMO A E F e ) / q and hole concentration p used in calculating open-circuit voltage V O C A from Equation (15), the calculated voltage offset ( E D A q V O C A ) , and the corresponding measured open circuit voltage V O C meas . .
Table 3. The calculated values of ( E LUMO A E F e ) / q and hole concentration p used in calculating open-circuit voltage V O C A from Equation (15), the calculated voltage offset ( E D A q V O C A ) , and the corresponding measured open circuit voltage V O C meas . .
Active Layer Blend ( E LUMO A E F e ) / q (V) p (m−3) V O C A (V) ( E D A q V O C A ) (V) V O C meas . (V)
MDMO-PPV:PCBM0.180.95 × 10220.930.370.83 [64]
P3HT:PC60BM0.240.95 × 10220.520.480.63 [51]
PCDTBT:PC71BM0.200.95 × 10220.730.470.85 [65]
OC1C10-PPV:PC61BM0.190.95 × 10220.940.360.85 [59]
PTB7:PC71BM0.160.89 × 10220.750.340.75 [54]
mono-DPP:PC71BM0.160.38 × 10220.790.370.78 [61]
bis-DPP:PC71BM0.181.58 × 10220.850.350.52 [61]
Table 4. The calculated values of open circuit voltage V O C D A from Equation (18) using the n and p values listed in Table 2 and Table 3, respectively, voltage offset ( E D A q V O C D A ) upon simultaneous excitation of both the donor and acceptor materials, and the corresponding measured open circuit voltage V O C meas . .
Table 4. The calculated values of open circuit voltage V O C D A from Equation (18) using the n and p values listed in Table 2 and Table 3, respectively, voltage offset ( E D A q V O C D A ) upon simultaneous excitation of both the donor and acceptor materials, and the corresponding measured open circuit voltage V O C meas . .
Active Layer Blend V O C D A (V) ( E D A q V O C D A ) (V) V O C meas . (V)
MDMO-PPV:PC61BM0.910.390.83 [64]
P3HT:PC60BM0.500.500.63 [51]
PCDTBT:PC71BM0.770.430.85 [65]
OC1C10-PPV:PC61BM0.910.390.85 [59]
PTB7:PC71BM0.740.350.75 [54]
mono-DPP:PC71BM0.790.370.78 [61]
bis-DPP:PC71BM0.850.350.52 [61]

Share and Cite

MDPI and ACS Style

Yeboah, D.; Singh, J. Study of the Contributions of Donor and Acceptor Photoexcitations to Open Circuit Voltage in Bulk Heterojunction Organic Solar Cells. Electronics 2017, 6, 75. https://doi.org/10.3390/electronics6040075

AMA Style

Yeboah D, Singh J. Study of the Contributions of Donor and Acceptor Photoexcitations to Open Circuit Voltage in Bulk Heterojunction Organic Solar Cells. Electronics. 2017; 6(4):75. https://doi.org/10.3390/electronics6040075

Chicago/Turabian Style

Yeboah, Douglas, and Jai Singh. 2017. "Study of the Contributions of Donor and Acceptor Photoexcitations to Open Circuit Voltage in Bulk Heterojunction Organic Solar Cells" Electronics 6, no. 4: 75. https://doi.org/10.3390/electronics6040075

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop