Next Article in Journal
Theoretical Analysis of Terahertz Dielectric–Loaded Graphene Waveguide
Next Article in Special Issue
Electrochemical Investigation of Phenethylammonium Bismuth Iodide as Anode in Aqueous Zn2+ Electrolytes
Previous Article in Journal
Nanoporous Gold Monolith for High Loading of Unmodified Doxorubicin and Sustained Co-Release of Doxorubicin-Rapamycin
Previous Article in Special Issue
Inorganic Materials by Atomic Layer Deposition for Perovskite Solar Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Characterising Exciton Generation in Bulk-Heterojunction Organic Solar Cells

by
Kiran Sreedhar Ram
1,
Hooman Mehdizadeh-Rad
1,2,
David Ompong
1,2,
Daniel Dodzi Yao Setsoafia
1 and
Jai Singh
1,2,*
1
College of Engineering, IT and Environment, Purple 12, Charles Darwin University, Darwin, NT 0909, Australia
2
Energy and Resources Institute, Charles Darwin University, Darwin, NT 0909, Australia
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(1), 209; https://doi.org/10.3390/nano11010209
Submission received: 17 December 2020 / Revised: 12 January 2021 / Accepted: 12 January 2021 / Published: 15 January 2021
(This article belongs to the Special Issue Perovskite Nanostructures: From Material Design to Applications)

Abstract

:
In this paper, characterisation of exciton generation is carried out in three bulk-heterojunction organic solar cells (BHJ OSCs)—OSC1: an inverted non-fullerene (NF) BHJ OSC; OSC2: a conventional NF BHJ OSC; and OSC3: a conventional fullerene BHJ OSC. It is found that the overlap of the regions of strong constructive interference of incident and reflected electric fields of electromagnetic waves and those of high photon absorption within the active layer depends on the active layer thickness. An optimal thickness of the active layer can thus be obtained at which this overlap is maximum. We have simulated the rates of total exciton generation and position dependent exciton generation within the active layer as a function of the thicknesses of all the layers in all three OSCs and optimised their structures. Based on our simulated results, the inverted NF BHJ OSC1 is found to have better short circuit current density which may lead to better photovoltaic performance than the other two. It is expected that the results of this paper may provide guidance in fabricating highly efficient and cost effective BHJ OSCs.

1. Introduction

High efficiency, stability, low cost and short energy pay back times are the key issues to be considered in the development and commercialization of any type of solar cell to convert solar energy into electricity, which is currently dominated by the silicon solar cells [1,2,3]. The silicon solar cells are highly efficient and stable [4,5] but their fabrication technology is very expensive because it is high temperature based. As a result, much research efforts have currently been devoted to the development of organic solar cells (OSCs) which can be fabricated through the low temperature based chemical process. Among the organic solar cells, the bulk heterojunction (BHJ) OSCs reaching the power conversion efficiency up to 18% [6] have the potential to replace silicon solar cells [7]. The fabrication methods for OSCs are compatible with printing techniques and roll-to-roll processes which are suitable for mass production [8,9]. Currently, research efforts are focused on fabricating highly efficient and stable BHJ OSCs to make them commercially viable and competitive to inorganic crystalline silicon solar cells [10]. As organic solids degrade due to long exposure to sunshine, BHJ OSCs are not only poor in efficiency but also in stability [11]. This opens enormous research opportunities for enhancing the poor efficiency and stability of OSCs which at present are the stumbling blocks towards commercializing this environmentally friendly technology.
The basic structure of a BHJ OSC consists of an active layer composed of two types of blended organic semiconductor materials, a donor and an acceptor [12]. The photogeneration of excitons can occur in both the materials, but when excitons reach a donor-acceptor interface (D-A), the charge transfer (CT) excitons are formed such that the donor donates electrons and accepts holes whereas the acceptor donates holes and accepts electrons [13]. Therefore, donor materials should be efficient hole acceptors and acceptor materials should be efficient electron acceptors [14]. Such an active layer is sandwiched between two electrodes of different work functions that provide an internal electric field in BHJ OSCs to transport the free electrons and holes, generated from the dissociation of CT excitons within the active layer, to their respective electrodes [15].
As organic semiconductors have a low dielectric constant, about 3–4 [16,17,18,19], the electron and hole pairs excited by the absorption of photons by the donor and acceptor in the active layer form Frenkel excitons instantly, which need to be dissociated efficiently into free electrons and holes for the operation of OSCs. Therefore, the research in improving the power conversion efficiency (PCE) focuses on high absorption and fast dissociation of excitons. The dissociation of Frenkel excitons occurs at a D-A interface, which is usually not far away in a blended donor-acceptor active layer. When a Frenkel exciton excited in the donor reaches a D-A interface, the electron gets transferred from the donor’s lowest unoccupied molecular orbital (LUMO) to the acceptor’s LUMO, being at a lower energy, and thus forms a CT exciton [20,21] by releasing the excess energy in the form of molecular vibrations. Likewise, when a Frenkel exciton excited in the acceptor reaches a D-A interface, the hole gets transferred from acceptor highest occupied molecular orbital (HOMO) to donor HOMO, being at a lower energy, and forms a CT exciton in the form of molecular vibrations [13]. If this excess energy is equal or larger than the exciton binding energy and impacts back on the CT exciton, the latter gets dissociated into a free electron and hole which are transported to their respective electrodes by the built-in electric field due to work function difference between the electrodes. After the dissociation, some of the free electrons and holes may recombine before reaching their electrodes and do not contribute to current generation. The recombination losses in OSCs can occur through three major carrier loss pathways, which are the bimolecular, bulk trap assisted, and surface trap assisted pathways [22]. The remaining charge carriers reach their respective electrodes and contribute to the external current. In view of the above, it is clear that the exciton generation plays a crucial role in PCE of BHJ OSCs. It is therefore very desirable to study the profile of exciton generation in OSCs with a view to enhance their photovoltaic performance.
BHJ OSCs with an active layer based on fullerene acceptor have currently dominated the research activities in organic photovoltaic [23] because of their excellent charge transport properties [24]. Recently, Mehdizadeh Rad et al. [25] applied the optical transfer matrix method (OTMM) to profile the exciton generation rate as a function of the thickness of the active layer for the fullerene acceptor (PCBM) based BHJ OSC in conventional as well inverted configurations. The normalized modulus squared of the electric field and its constructive interference points (CIPs) are analysed using the contour plots. Assuming equal mobility for electrons and holes, and analysing the positions of CIPs within the active layer, they found that for the efficient extraction of charge carriers, excitons should be generated halfway within the active layer; about equal distance from the two electrodes. This is important in the design of organic solar cells to minimize the creation of space charge in the device. These results also conclude that for the same active layer thickness, an inverted fullerene BHJ OSC structure can generate larger short circuit current density than a conventional fullerene BHJ OSC structure [25].
However, fullerene acceptor materials have some disadvantages which include limited chemical and energetic tunability, narrow range of absorption spectra, and unstable morphology, thereby limiting the overall PCE and stability of the device thus fabricated [26,27]. Therefore, the research focus has moved to the use of non-fullerene (NF) acceptors [28,29,30,31] in BHJ OSCs. Recently, Liu et al. [6] fabricated a NF acceptor based BHJ OSC with a PCE of 18.22%, which is the highest to date for a single junction BHJ OSC device. Unlike the fullerene OSCs, the NF devices are much easier to process using solution-based chemistry [32]. It is also possible to increase the device V o c while maintaining a high J s c which is not possible with the fullerene based OSC [33]. This is achieved because of the easy tunability of bandgap energy of the NF acceptor materials and the low binding energy of the CT excitons which reduces the energy loss at the CT state to dissociate the excitons into free charge carriers [33]. To the best of the authors’ knowledge, little theoretical work has been done in understanding the characteristics of exciton generation in conventional and inverted NF acceptor based BHJ OSCs.
Many research activities have been carried out to study the influence of electron transport layer (ETL) and hole transport layer (HTL) in BHJ OSCs [34,35,36]. The work of Rasool et al. [37] aimed at improving the stability of BHJ OSCs by studying the effect of direct contact of the ETL materials on the active layer. Research by Yang et al. [38] focused on improving the built-in potential and reducing the series resistance of BHJ OSCs by optimising the work function of ETL and forming ohmic contacts with the active layer acceptor materials. A report by Lee et al. [39] focused on optimising the thickness of ETL [39]. Most of the researches on HTL are focused on the fabrication of organic and inorganic HTL materials with high transparency and low resistance, as exemplified by the work of Sivakumar et al. [40]. The focus of Chien et al. [41] on HTL was in understanding the inhomogeneity and its effect on the device performance. However, there has not been much focus in understanding the interference effects which the incident light undergoes when passing through ETL and HTL and the effect it will have on the photon absorption and exciton generation in the active layer.
In view of the above, in this paper, we studied the interference of the incident and reflected electric fields of the electromagnetic radiation and exciton generation rate in the active layer of a conventional fullerene acceptor based BHJ OSC, and two NF acceptor based BHJ OSCs, of inverted and conventional structures, as a function of the thickness of the active layer. In addition, the exciton generation rate was also studied as a function of the thickness of other layers in all three OSCs.
The paper is organized as follows. In Section 1, a detailed introduction is presented, followed by Section 2 which describes the theories that are used to study the interference of incident and reflected electric fields and exciton generation rate in BHJ OSCs, including the study of recombination losses which affect the short circuit current density. In Section 3, the results and discussions are presented, and finally, in Section 4 the conclusions are drawn.

2. Theory

To apply the optical transfer matrix method (OTMM) to study the optical properties, a BHJ OSC can be considered as a multilayer stack as shown in Figure 1. As the thickness of each layer is comparable with the wavelength of incident light, the light absorption and exciton generation are affected by the interference effects [25]. For this work, we coded the OTMM program and used it for simulation on MATLAB version 2020a Update 1 (9.8.0.1359463), USA.
The total exciton generation rate in the active layer of the BHJ OSC shown in Figure 1 can be written as [25,42,43]:
G ˙ j = x = 0 L j G ˙ j ( x ) d x
where G ˙ j ( x ) is the exciton generation rate at distance x from the top interface of the active layer of thickness L j , and it can be given by [25,42,43]:
G ˙ j ( x ) = λ = λ 1 λ 2 G ˙ j ( x , λ ) d λ
where λ 1 and λ 2 are the lower and upper wavelength integration limits of solar irradiance,   λ is the wavelength of incident solar irradiance and G ˙ j ( x , λ ) is wavelength and position dependent exciton generation rate given by [25,42,43]:
G ˙ j ( x , λ ) = λ h c Q j ( x , λ )
where h is Planck’s constant, c is the speed of light, and Q j ( x , λ ) is the time average of the energy absorbed per second in the active layer j at a position x at normal incidence and it is given by [25,42,43]:
Q j ( x , λ ) = α j I 0 T i n t | t j + | 2 n j n 0 { e α j x + ρ j 2 e α j ( 2 L j x ) + 2 ρ j e α j L j cos [ 4 π n j ( L j x ) λ δ j ] }
where α j = 4 π k j / λ is the absorption coefficient, k j is the extinction coefficient of the active layer blend, I 0 is the intensity of incident solar irradiance at the cell surface, T i n t is the internal transmittance of the substrate, t j + is the internal transfer coefficient which relates the incident plane wave to the internal electric field propagating from the front end to rear end of the cell, n j and n 0 are the refractive indices of the active layer and substrate, respectively, and ρ j and δ j are the magnitude and angle of complex reflection coefficient, respectively. The wavelength dependent electric field of the electromagnetic (EM) waves in the active layer j at a distance x from the top interface is given by [25]:
E j ( x , λ ) = 2 Q j ( x , λ ) c ε 0 α j n j
where, ε 0 is the permittivity of free space.
Next step is to calculate the short circuit current density, J s c , from the total exciton generation rate G ˙ j after considering the recombination losses. The excitons generated in the blend have to reach the D-A interface to undergo dissociation into free electrons and holes [21]. Usually in BHJ OSCs, the exciton diffusion length is less than 10 nm, hence the probability that the excitons diffuse to an interface before recombination is nearly unity [25,44,45,46]. As stated above, the three main recombination loss rates that affect the J s c in a BHJ OSC are the rates of bimolecular recombination loss, G ˙ j b m r e c , bulk trap assisted recombination loss, G ˙ j b u l k r e c , and surface trap assisted recombination loss, G ˙ j s u r f r e c , which are respectively given by [22]:
G ˙ j b m r e c = e L j ε 0 ε r ξ ( μ n + μ p ) n 2
G ˙ j b u l k r e c = e L j ε 0 ε r μ c N t . b u l k n
and
G ˙ j s u r f r e c = e ε 0 ε r μ s N t . s u r f n e x p { e ( V b V c o r ) k T }
where n is the charge carrier density, ε r is the dielectric constant, ξ is the reduction factor, μ n and μ p are the mobility of electrons and holes, respectively, V c o r = V I t r a n s R a , is the corrected voltage, V is the applied voltage, I t r a n s = e { G ˙ j ( G ˙ j b m r e c + G ˙ j b u l k r e c ) } is the transport current density, R a = ρ L j is the series resistance of the active layer, ρ is the resistivity of the active layer blend, N t . b u l k and N t . s u r f are the densities of traps in the bulk and surfaces, respectively, of the active layer, V b is the built-in voltage, μ c is the mobility of slower moving charge carrier in the bulk, and μ s is the mobility of the minority charge carrier on the surfaces of the active layer. The net rate of the collected charge carriers reaching their respective electrodes, G ˙ j c o l , can be given by:
G ˙ j c o l = G ˙ j ( G ˙ j b m r e c + G ˙ j b u l k r e c + G ˙ j s u r f r e c )
where the rates of recombination losses in Equations (6)–(8) have been subtracted from the rate in Equation (1). Using Equation (9), J s c can be obtained as:
J s c = e G ˙ j c o l

3. Results and Discussions

Simulations of the position dependent exciton generation rate G ˙ j ( x ) within the active layer j of the following three BHJ OSC1, OSC2, and OSC3 have been carried out. The schematic structures of OSC1, OSC2, and OSC3 are shown in Figure 2a–c, respectively, and the details of their structures are as follows: OSC1: an inverted BHJ OSC with a non-fullerene acceptor of the structure: Glass/indium tin oxide (ITO) (150 nm)/zinc oxide (ZnO) (30 nm)/Poly[(2,6-(4,8-bis(5-(2-ethylhexylthio)-4-fluorothiophen-2-yl)-benzo [1,2-b:4,5-b′]dithiophene))-alt-(5,5-(1′,3′-di-2-thienyl-5′,7′-bis(2-ethylhexyl)benzo[1’,2′-c:4′,5′-c’]dithiophene-4,8-dione)]: C94H78F4N4O2S4 (PBDBTSF:IT4F)/molybdenum trioxide (MoO3)(10 nm)/Aluminium (Al) (100 nm) [47,48] (Figure 2a); OSC2: a conventional non-fullerene BHJ OSC of the structure: Glass/ITO(150 nm)/poly(3,4-ethylenedioxythiophene):polystyrenesulfonate (PEDOT:PSS) (30 nm)/PBDB-T-SF:IT-4F/Poly(9,9-bis(3′-(N,N-dimethyl)-N-ethylammoinium-propyl-2,7-fluorene)-alt-2,7-(9,9-dioctylfluorene))dibromide (PFN-Br) (5 nm)/Al (100 nm) [49] (Figure 2b); and OSC3: a conventional fullerene BHJ OSC of the structure: Glass/ITO (180 nm)/PEDOT:PSS (45 nm)/poly(3-hexylthiophene):[6,6]-phenyl C61-butyric acid methyl ester (P3HT:PCBM)/Lithium Fluoride (LiF) (1 nm)/Al (100 nm) [50,51] (Figure 2c). It may be noted that the chosen thicknesses of the layers, other than the active layer, in the above three OSCs are the optimal thicknesses obtained from experiments cited above.

3.1. Test of Simulation

Before presenting the details of our simulation, it is desirable to check the accuracy of our simulation by comparing the simulated J s c with some experimental results. For this purpose, we calculated J s c of OSC1, OSC2, and OSC3 using Equations (1)–(10), which require the input parameters given in Table 1 and experimental values of the refractive index n l and extinction coefficient k l (where l = 1, 2, j, 4, and 5) for PBDBTSF:IT4F and PFN-Br from [49,52], for the Glass, ITO, P3HT:PCBM, PEDOT:PSS, LiF and Al from [50,51,53,54,55], and for ZnO and MoO3 from [56]. For completeness, the experimental values of nj and kj of the two active layer materials PBDBTSF:IT4F and P3HT:PCBM are plotted in Figure 3 as a function of the wavelength of the incident light.
The simulated J s c for OSC1, OSC2, and OSC3, thus obtained, are plotted as a function of the active layer thickness L j in Figure 4 along with the experimental results from [47] for OSC1, from [49] for OSC2, and from [50,51] for OSC3. According to Figure 4, the simulated results of J s c agree reasonably well with the experimental ones and hence prove that our simulation can be regarded to be reasonably accurate. However, it may be pointed out here that for OSC2, we have the experimental result only for one thickness (see Figure 4).

3.2. Electric Field and Exciton Generation Rate Distributions in OSC1

In this section, we present the simulated distributions of the normalised modulus of electric field component | E j ( x , λ ) | (Equation (5)) of the electromagnetic (EM) radiation and the subsequent exciton generation rate G ˙ j ( x , λ ) (Equation (3)) within the active layer of OSC1 using the OTMM method described in Section 2. All input parameters given in Table 1 and n l and k l values for PBDBTSF:IT4F from [49,52], for the Glass, ITO and Al from [50,51,53,54,55], and for ZnO and MoO3 from [56] are used in this simulation. Both | E j ( x , λ ) | from Equation (5) and G ˙ j ( x , λ ) from Equation (3) are calculated for three active layer thicknesses—35 nm, 95 nm, and 180 nm—of OSC1 and their contour plots as a function of position in the active layer and wavelength of the solar radiation are shown in Figure 5 in six parts; three on the left-hand side present the electric field distribution in the active layer of three thicknesses—35, 95, and 180 nm—and the corresponding right hand side figures present the distribution of the exciton generation rate.
As a BHJ OSC is a multilayer stack, these brighter and darker regions of | E j ( x , λ ) | in the contour plots shown in Figure 5a,c,e are due to constructive interference (CI) and destructive interference (DI), respectively, of the forward travelling electric field, E j + , and reflected electric field, E j of EM waves inside the active layer, as shown in Figure 1. Likewise, in the contour plots of Figure 5b,d,f, the brighter spots represent regions of higher G ˙ j ( x , λ ) which are also the higher photon absorption regions (HPARs) and lower G ˙ j ( x , λ ) , or lower photon absorption regions (LPARs) are represented by darker spots.
In the active layer, higher generation rate G ˙ j ( x , λ ) , is expected to give higher Jsc and hence an active layer thickness that gives the widest brighter spot for G ˙ j ( x , λ ) may be regarded to exhibit the optimum performance. However, although all the figures shown in Figure 5 are of the same width, they correspond to different thicknesses of the active layer. That means, widths of the brighter spots in Figure 5a,b are much smaller than those in Figure 5c,d and largest in Figure 5e,f. The wider brighter spot in G ˙ j ( x , λ ) may occur if the electric field distribution is also wider in the active layer. However, the locations of brighter spots in G ˙ j ( x , λ ) occur where the wavelength dependent extinction coefficient ( k j ) is non-zero but such spots may not always coincide with the locations of CI of | E j ( x , λ ) | for all active layer thicknesses. In view of this, it becomes more important to monitor the brighter spots in G ˙ j ( x , λ ) whether these spots correlate or not with those in | E j ( x , λ ) | . However, an optimal thickness of the active layer can still be expected to be in which both G ˙ j ( x , λ ) and | E j ( x , λ ) | have maximum overlap of bright spots. With this view, it may be concluded that the active layer thickness of 95 nm with | E j ( x , λ ) | and G ˙ j ( x , λ ) shown in Figure 5c,d, respectively, is expected to give the optimal photovoltaic performance. This conclusion that the optimal thickness is 95 nm agrees very well with the experimental results of Zhao et al. [47], who found optimal PCE at the same thickness. However, if one compares Jsc shown in Figure 4, the active layer thicknesses of 180 nm has only slightly less Jsc (≈0.03 mA cm2 less) than at the optimal active layer thickness of 95 nm. This is because, as explained above, the amount of exciton generation depends on the thickness; with larger thickness more exciton generation rate and hence an active layer of thickness 180 nm is expected to have the widest brighter exciton generation spots leading to Jsc comparable with that of thickness 95 nm. However, at Lj = 180 nm the fill factor (FF) is less than that at Lj = 95 nm and hence the PCE reduces. In addition to lower PCE, an increased thickness of L j = 180 nm is expected to require nearly twice the amount of active layer material and therefore is uneconomical.

3.3. Influence of Other Layers on Exciton Generation Rate

Here we study the effect of varying the thickness of electrodes and charge transport layers on G ˙ j and G ˙ j ( x ) in all three OSCs, which have layer 1 as the front electrode (FE) of thickness L1, layer 2 as the front charge transport layer (FCTL) of thickness L2, layer 4 as the rear charge transport layer (RCTL) of thickness L4, and layer 5 as the rear electrode (RE) of thickness L5. To the best of the authors’ knowledge, a theoretical study about the influence of thickness of layers other than the active layer on G ˙ j and G ˙ j ( x ) in NF BHJ OSCs has not been done yet. For OSC1, using n l and k l values for each layer as described above and other required parameters from Table 1, G ˙ j (Equation (1)) is calculated for two active layer thickness Lj = 95 nm and 180 nm by varying the thicknesses of other layers. G ˙ j ( x ) (Equation (2)) as a function of x is calculated only for the optimal active layer thicknesses Lj = 95 nm by varying other layer thicknesses. Thus, calculated G ˙ j is plotted as a function of the thickness of each of the four layers on the left column of Figure 6 and the corresponding contour plot G ˙ j ( x ) within the active layer of optimal thickness on the right column of Figure 6.
According to Figure 6a,c,e, G ˙ j exhibits an oscillatory behaviour when the thickness of the ITO, ZnO, and MoO3 layers are varied. These oscillations may be attributed to the interference effect of the forward and reflected electric fields of the electromagnetic waves. However, such oscillations do not occur when the thickness of Al layer (layer 5) is varied as shown in Figure 6g, which will be discussed later.
In Figure 6a, the maxima of G ˙ j are observed for the ITO thickness L1 = 148 nm and 19 nm corresponding to the active layer thickness Lj = 95 nm and 180 nm, respectively. From this, it is evident that for thicker active layer Lj, thinner ITO layer L1 is required to get maximum G ˙ j . Although G ˙ j is the highest for L1 = 19 nm at Lj = 180 nm, such an OSC may have very large resistive power loss due to the large series resistance offered by the thin ITO layer L1 [67]. Thus, the simulated optimal thickness of ITO is chosen to be L1 = 148 nm when the active layer thickness Lj = 95 nm, which is in good agreement with the experimental result [47]. Figure 6b shows the effect of varying L1 on G ˙ j ( x ) at different positions inside the active layer of the optimal thickness, Lj = 95 nm. According to Figure 6b, the most exciton generation rate G ˙ j ( x ) occurs within the first 60 nm of the optimal active layer thickness Lj = 95 nm with two maxima (brightest) spots at L1 = 50 nm and 148 nm, which are consistent with the maxima in G ˙ j shown in Figure 6a.
The effect of varying the thickness L2 (ZnO) on G ˙ j and G ˙ j ( x ) are shown in Figure 6c,d, respectively. It is found from Figure 6c that G ˙ j is high for 0 < L2 < 30 nm at the optimum active layer thickness Lj = 95 nm, but for the active layer thickness of Lj = 180 nm, G ˙ j is high for 120 < L2 < 140 nm. The FCTL (ZnO) helps in reducing the charge recombination and contact resistance between the active layer and FE (ITO). Therefore, on one hand, L2 should be non-zero to serve the above two purposes. On the other hand, all the layers above the active layer should be as thin as possible to maximise the transmission of light to the active layer. For the optimal active layer thickness of Lj = 95 nm, the L2 may be optimised to maximise photon absorption and least charge recombination loss. Hence, the optimum value of L2 may be considered to be 30 nm which is in good agreement with the experimental result [47]. According to Figure 6d, the high G ˙ j ( x ) occurs within the active layer for 0 < x < 40 nm and for 0 < L2 < 30, which is consistent with the corresponding G ˙ j in Figure 6c.
Figure 6e,f shows G ˙ j and G ˙ j ( x ) , respectively, as a function of the thickness L4 of layer 4 (MoO3). Here again, the oscillatory nature of G ˙ j occurs due to interference but the amplitude of oscillations is relatively larger in comparison with all other layers. This may be attributed to the fact that the unabsorbed light from the active layer and reflected from the Al cathode passes through this layer again before getting absorbed in the active layer, leading to a larger interference effect. The amplitudes of oscillations for the active layer thickness Lj = 95 nm are much larger than those for Lj = 180 nm. According to Figure 6e, the simulated optimal thickness of L4 (MoO3) is 10 nm at the optimal active layer thickness Lj = 95 nm, which is in good agreement with experimental L4 [47]. This is also in agreement with G ˙ j ( x ) , shown in Figure 6f, where the widest bright spot is found to occur at L4 = 10 nm.
In Figure 6g,h, G ˙ j and G ˙ j ( x ) are plotted, respectively, as a function of the thickness L5 of the cathode (Al) and as stated above no oscillations are found in G ˙ j (Figure 6g) for L5 > 55 nm. This is because all the unabsorbed light from the active layer gets reflected from the interface between MoO3 and Al layer back into the active layer without any interference effect occurring within the Al layer. Therefore, G ˙ j remains unaffected by any variation in the thickness of Al layer when L5 > 55 nm. For 0 < L5 < 55 nm, the reflection from Al starts to increase from zero, which increases the absorption in the active layer and hence G ˙ j starts increasing linearly from its non-zero value at L5 = 0 nm. Such a dependence of G ˙ j on L5 is observed for both the active layer thicknesses 95 and 180 nm (see Figure 6g). For L5 > 16 nm, the G ˙ j in the active layer of thickness Lj = 95 nm shown in Figure 6g is larger than that in Lj = 180 nm. As Jsc is proportional to the net G ˙ j c o l (see Equation (10)), this is consistent with the larger Jsc in Lj = 95 nm than in Lj = 180 nm as illustrated in Figure 4. However, for L5 < 16 nm, this trend reverses because the photon absorption and exciton generation in the active layer increase with the thickness and the absorption of reflected photons from Al becomes negligible at such a low thickness. In view of the discussion presented above, it may be concluded that L5 may be equal to 55 nm or larger for obtaining maximum G ˙ j in the active layer, which agrees very well with the L5 = 100 nm used experimentally [47]. The thickness L5 > 55 nm used in the fabrication [47] may serve the purpose of enhancing conductivity and thereby reducing the resistive power loss in Al layer.
We will now discuss the simulated results of OSC2 of the structure: Glass/ITO/PEDOT:PSS/PBDB-T-SF:IT4F/PFN-Br/Al. The simulated Jsc of OSC2 is shown in Figure 4 along with that of OSC1 for comparison. It is found that Jsc in OSC2 is maximum at the active layer thickness Lj = 100 nm, which agrees very well with the experimental value [49]. It is also evident from Figure 4 that Jsc of the conventional structure OSC2 is slightly lower than that of the inverted structure OSC1 which agrees with our previous simulation work [25], where it has been found that for a fullerene based BHJ OSC with P3HT:PCBM as the active layer, the inverted structure has a higher Jsc than the conventional structure. In this view, it may be concluded that the inverted structure has higher Jsc than conventional structure in both fullerene and non-fullerene acceptor based BHJ OSCs.
Further analysis of the performance of OSC2 has been carried out by simulating G ˙ j and G ˙ j ( x ) using the n l and k l values for all layers decribed above and other required parameters from Table 1. The variation of G ˙ j with L1, L4, and L5 for OSC2 is given in Figure S1 in the Supplementary Information. The variation of G ˙ j in OSC2, shown in Figure S1, with the thicknesses L1, L4, and L5 is similar to that in OSC1 shown in Figure 6, with the only difference being the magnitude of G ˙ j ; in OSC2 G ˙ j is slightly lower than that in OSC1. According to Figure S1, the optimal thickness L1 of layer 1, L4 of layer 4, and L5 of layer 5 are 148 nm, 5 nm, and 100 nm, respectively, for OSC2. However, as discussed below, the dependence of G ˙ j and G ˙ j ( x ) on the thickness L2 of PEDOT:PSS in OSC2 shown in Figure 7a,b is slightly different than that of L2 which is ZnO in OSC1 shown in Figure 6c,d.
G ˙ j plotted as a function of L2 for two active layer thicknesses L j = 100 nm and 180 nm in OSC2 is shown in Figure 7a and it decreases as L2 increases with relatively small interference effect compared to that in OSC1. According to Figure 7a, G ˙ j at the active layer thickness L j = 100 nm decreases from about 1.42 × 1021 m2s1 at L2 = 0 nm to about 1.18 × 1021 m2s1 at L2 = 300 nm and for L j = 180 nm, it first decreases from about 1.36 × 1021 m2s1 at L2 = 0 nm to 1.26 × 1021 m2s1 at L2 = 80 nm and then increases to a maximum of 1.28 × 1021 m2s1 at L2 = 148 nm and then finally decreases again to 1.15 × 1021 m2s1 at L2 = 300 nm. This means that the G ˙ j for L j = 180 nm in OSC2 is high for 0 nm < L2 < 30 nm unlike the case of OSC1. Although G ˙ j is maximum at L2 = 0 nm, which is not suitable, the optimal thickness L2 of layer 2 is chosen to be 30 nm, to be consistent with the fabricated thickness [49] and with that used above for OSC1. It may be emphasized that G ˙ j in Figure 7a at L2 = 148 nm is higher for the active layer thickness 180 nm than that for 100 nm but it is still lower than that at L2 = 30 nm for the active layer thickness of 100 nm and hence the optimal thicknesses of both layer 2 and active layer for OSC2 are chosen to be 30 nm and 100 nm, respectively.
For comparing G ˙ j in OSC1 and OSC2, we find that at the optimal active layer thickness Lj = 95 nm in OSC1, G ˙ j at L2 = 30 nm is 1.385 × 1021 m2s1 and in OSC2 it is 1.355 × 1021 m2s1 at the optimal active layer thickness Lj = 100 nm. The lower G ˙ j in OSC2 contributes to a lower Jsc which is evident from Figure 4. The contour plots of | E j ( x , λ ) | and G ˙ j ( x , λ ) for OSC2 with the thicknesses of all layers optimised are plotted in Figure S2a,b, which are apparently similar to those of OSC1 shown in Figure 5c,d. For a comparative analysis we plotted G ˙ j ( x ) as a function of the position x within the active layer for the optimal structures of both OSC1 and OSC2 in Figure 8. The G ˙ j ( x ) of OSC1 (see Figure 8) is found to be slightly higher than that of OSC2 at the position x = 0 nm and the difference reduces gradually towards the end of the active layer thickness Lj and hence the inverted OSC1 can be expected to show better performance compared to conventional OSC2. As the optimised thickness Lj of the active layer of OSC1 is 95 nm and that of OSC2 is 100 nm, the amount of active layer material needed for the inverted OSC1 is less than that of the conventional OSC2 which may lead to lower fabrication cost in the mass production of inverted structure compared to the conventional structure.
Finally, we discuss the simulation results of OSC3, which is a fullerene acceptor based BHJ OSC of the conventional structure. Jsc of OSC3 plotted in Figure 4, reveals the optimal active layer thickness of 80 nm, which is in reasonable agreement with the experimental thickness of 70 nm [50,51]. It is also very conclusive from Figure 4 that OSC3 produces relatively the lowest Jsc in comparison with OSC1 and OSC2, which may be attributed to the use of fullerene acceptor [60]. Following the procedures for OSC1 and OSC2, G ˙ j and G ˙ j ( x ) as a function of the thickness of layer 1, layer 2, layer 4, and layer 5 for OSC3 are also simulated using the n l and k l values for P3HT:PCBM from [50,51,53,54,55], for the Glass, ITO, PEDOT:PSS, LiF and Al from [50,51,53,54,55] and other required parameters from Table 1 in Equations (1) and (2), respectively. G ˙ j (Equation (1)) is calculated for two active layer thickness Lj = 80 nm and 180 nm as a function of the thicknesses of other layers whereas the corresponding contour plots of G ˙ j ( x ) (Equation (2)) is calculated as a function of x within the active layer only of the optimal thicknesses Lj = 80 nm. The variations in G ˙ j and G ˙ j ( x ) of OSC3 with respect to the thickness of layer 4 and layer 5 (see Figure S3) show similar trends as those in OSC1 (see Figure 6) and OSC2 (see Figure S1) and their simulated optimal thickness are obtained as 1 nm and 100 nm, respectively. The variations in G ˙ j and G ˙ j ( x ) with the thickness of layers 1 and 2 are shown in Figure 9a–d.
The variation of G ˙ j with L1 for the active layer thicknesses of Lj = 80 nm and 180 nm in OSC3, as shown in Figure 9a, exhibits larger oscillatory nature compared to OSC1 (Figure 6a) and OSC2 (Figure S1a). This may be attributed to the larger interference effect caused by the combination of fullerene-based accepter in the active layer and layer 1. In view of this, a small variation from the optimal thickness of the layer 1 may lead to a larger loss in G ˙ j , and hence in Jsc in the fullerene based OSC3 in comparison with that in non-fullerene based OSC1 and OSC2. From Figure 9a, the optimal thickness L1 of layer 1 is found to be 46 nm at the optimal active layer thickness Lj = 80 nm. However, the thickness of ITO used in the fabrication of fullerene based OSC3 is usually 150–180 nm [50,51,68], which corresponds to the second maximum in Figure 9a and it is lower than the first one at 46 nm. Although the optimal thickness of 46 nm determined here from the simulation agrees with 40 nm used in our earlier simulation work [25], a thicker ITO layer used in the fabrication might help reducing the resistive power loss in the series resistance of layer 1 because ITO sheet resistance decreases significantly with the increase in its thickness [61]. It may be noted that, the first and second maxima shown in Figure 9a at the active layer thickness of 80 nm, are also clearly visible as widest bright spots at L1 = 46 nm and 180 nm, respectively, in Figure 9b.
In Figure 9c, we plotted G ˙ j as a function of the thickness L2 of PEDOT:PSS in OSC3 for the two active layer thicknesses Lj = 80 nm and 180 nm and the contour plot of G ˙ j ( x ) as a function of the thickness L2 and position x within the active layer of optimal thickness L j = 80 nm is shown in Figure 9d. According to Figure 9c, the G ˙ j is maximum at L2 = 0 nm for both the active layer thicknesses and hence the maximum exciton generation rate may occur in the absence of layer 2. This is also evident from Figure 9d, where the widest bright spot appears at L2 = 0 nm. This trend of maximum G ˙ j at L2 = 0, shown in Figure 9c is similar to that found in OSC2 (see Figure 7a) and also in OSC1 for Lj = 95 nm but for Lj = 180 nm the maximum G ˙ j occurs at non-zero thickness L2 (see Figure 6c). As mentioned above, the layer 2 is used to improve the charge carrier extraction efficiency and hence it cannot be of zero thickness. In the experimentally fabricated OSC3 [50,51], a thickness of L2 = 45 nm was used, and according to Figure 9c, the simulated G ˙ j was found to be 6.73 × 1020 m−2s−1. Although a similar G ˙ j may be found for L2 = 100 nm (see Figure 9c), the thinner layer may be preferred for reducing the amount of material used. Figure 9c also reveals that G ˙ j for the active layer thickness of 180 nm is higher than that for 80 nm in the whole range of thickness L2 but the optimal thickness of 80 nm is preferred. This may be due to the fact that the fill factor of OSCs decreases almost linearly as the thickness of active layer increases [42,59,69], which is caused by the enhanced recombination [59,69]. The contour plots of | E j ( x , λ ) | and G ˙ j ( x , λ ) for OSC3 with the thicknesses of all layers optimised are plotted in Figure S4a,b.
The simulation presented in this paper is expected to help in optimising the layered structure of both non-fullerene and fullerene based BHJ OSCs and has revealed that the inverted NF BHJ OSC1 has relatively better Jsc than the conventional NF BHJ OSC2 and fullerene based OSC3. The optimised structures thus simulated for all three OSCs are given in Table 2.

4. Conclusions

Characterisation of the exciton generation rates has been carried out in three different BHJ OSCs: (i) an inverted non-fullerene acceptor based OSC1; (ii) a conventional non-fullerene acceptor based OSC2; and (iii) a conventional fullerene acceptor based OSC3. It was found that the regions of overlap of the constructive interference of forward travelling and reflected light waves and those of the high photon absorption regions depend on the thickness of active layer which is used to optimise the thickness of the active layer with the maximum overlap. The rates of total exciton generation G ˙ j and position dependent exciton generation G ˙ j ( x ) in the active layer have been optimised with respect to the thicknesses of each layer in the device structure. The optimal simulated designs thus obtained for the three OSCs are given in Table 2. It is expected that the results of this paper may help in fabricating highly efficient and cost-effective BHJ OSCs.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/11/1/209/s1. Figure S1: For OSC2, G ˙ j is plotted for two active layer thicknesses L j = 100 nm and 180 nm as function of the thickness of (a) L1 of layer 1 (ITO), (b) L4 of layer 4 (PFN-Br), (c) L5 of layer 5 (Al), Figure S2: (a) The contour plot of | E j ( x , λ ) | from Equation (5) in the active layer of OSC2 for L j   = 100 nm. The brighter spots represent regions of the constructive interference of electric field of EM waves within the active layer and the darker spots represent regions of destructive interference. (b) The corresponding G ˙ j ( x , λ ) contour plots from Equation (3), showing the brighter spots as positions of higher G ˙ j ( x , λ ) and darker spots that of lower G ˙ j ( x , λ ) , Figure S3: For OSC3, G ˙ j is plotted for two active layer thicknesses L j = 80 nm and 180 nm, as function of the thickness of (a) L4 of layer 4 (PFNBr) and (b) L5 of layer 5 (Al), Figure S4: (a) The contour plot of | E j ( x , λ ) | from Equation (5) in the active layer of OSC3 for L j = 80 nm. The brighter spots represent regions of the constructive interference of electric field of EM waves within the active layer and the darker spots represent regions of destructive interference. (b) The corresponding G ˙ j ( x , λ ) contour plots from Equation (3), showing the brighter spots as positions of higher G ˙ j ( x , λ ) and darker spots that of lower G ˙ j ( x , λ ) .

Author Contributions

K.S.R. carried out the details of simulation work under the supervision of the corresponding author J.S.; H.M.-R. and D.O. provided insight in MATLAB software and D.D.Y.S. helped with the literature review. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this article and supplementary information.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Abdulrazzaq, O.A.; Saini, V.; Bourdo, S.; Dervishi, E.; Biris, A.S. Organic Solar Cells: A Review of Materials, Limitations, and Possibilities for Improvement. Part. Sci. Technol. 2013, 31, 427–442. [Google Scholar] [CrossRef]
  2. Platts, S.P.G. BP Statistical Review of World Energy; 1 St James’s Square London SW1Y 4PD Natural Gas: London, UK, 2019; p. 64. [Google Scholar]
  3. Ram, K.S.; Singh, J. Over 20% Efficient and Stable Non-Fullerene-Based Ternary Bulk-Heterojunction Organic Solar Cell with WS 2 Hole-Transport Layer and Graded Refractive Index Antireflection Coating. Adv. Theory Simul. 2020, 3, 2000047. [Google Scholar] [CrossRef]
  4. Blakers, A.; Zin, N.; McIntosh, K.R.; Fong, K. High Efficiency Silicon Solar Cells. Energy Procedia 2013, 33, 1–10. [Google Scholar] [CrossRef] [Green Version]
  5. Bhattacharya, S.; John, S. Beyond 30% Conversion Efficiency in Silicon Solar Cells: A Numerical Demonstration. Sci. Rep. 2019, 9, 12482–12515. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, Q.; Jiang, Y.; Jin, K.; Qin, J.; Xu, J.; Li, W.; Ding, L. 18% Efficiency organic solar cells. Sci. Bull. 2020, 65, 272–275. [Google Scholar] [CrossRef] [Green Version]
  7. Dennler, G.; Sariciftci, N.S. Flexible Conjugated Polymer-Based Plastic Solar Cells: From Basics to Applications. Proc. IEEE 2005, 93, 1429–1439. [Google Scholar] [CrossRef]
  8. Hoth, C.N.; Schilinsky, P.; Choulis, S.A.; Brabec, C.J. Printing Highly Efficient Organic Solar Cells. Nano Lett. 2008, 8, 2806–2813. [Google Scholar] [CrossRef] [Green Version]
  9. Sondergaard, R.R.; Hösel, M.; Angmo, D.; Larsen-Olsen, T.T.; Krebs, F.C. Roll-to-roll fabrication of polymer solar cells. Mater. Today 2012, 15, 36–49. [Google Scholar] [CrossRef] [Green Version]
  10. Ram, K.S.; Singh, J. Highly Efficient and Stable Solar Cells with Hybrid of Nanostructures and Bulk Heterojunction Organic Semiconductors. Adv. Theory Simul. 2019, 2, 1900030. [Google Scholar] [CrossRef]
  11. Karakawa, M.; Suzuki, K.; Kuwabara, T.; Taima, T.; Nagai, K.; Nakano, M.; Yamaguchi, T.; Takahashi, K. Factors contributing to degradation of organic photovoltaic cells. Org. Electron. 2020, 76, 105448. [Google Scholar] [CrossRef]
  12. Duan, L.; Uddin, A. Progress in Stability of Organic Solar Cells. Adv. Sci. 2020, 7, 1903259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. 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]
  14. Bernède, J.C. Organic photovoltaic cells: History, principle and techniques. J. Chil. Chem. Soc. 2008, 53, 1549–1564. [Google Scholar] [CrossRef] [Green Version]
  15. Vithanage, D.A.; Devižis, A.; Abramavicius, D.; Infahsaeng, Y.; MacKenzie, R.C.I.; Keivanidis, P.E.; Yartsev, A.; Hertel, D.; Nelson, J.; Sundström, V.; et al. Visualizing charge separation in bulk heterojunction organic solar cells. Nat. Commun. 2013, 4, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Liu, X.; Xie, B.; Duan, C.; Wang, Z.; Fan, B.; Zhang, K.; Lin, B.; Colberts, F.J.M.; Ma, W.; Janssen, R.A.J.; et al. A high dielectric constant non-fullerene acceptor for efficient bulk-heterojunction organic solar cells. J. Mater. Chem. A 2018, 6, 395–403. [Google Scholar] [CrossRef]
  17. Roncali, J. Molecular Bulk Heterojunctions: An Emerging Approach to Organic Solar Cells. Accounts Chem. Res. 2009, 42, 1719–1730. [Google Scholar] [CrossRef] [Green Version]
  18. Kippelen, B.; Brédas, J. Organic photovoltaics. Energy Environ. Sci. 2009, 2, 251–261. [Google Scholar] [CrossRef]
  19. Brédas, J.; Norton, J.E.; Cornil, J.; Coropceanu, V. Molecular Understanding of Organic Solar Cells: The Challenges. Accounts Chem. Res. 2009, 42, 1691–1699. [Google Scholar] [CrossRef]
  20. Ono, S.; Ohno, K. Origin of Charge Transfer Exciton Dissociation in Organic Solar Cells. In Excitons; Pyshkin, S., Ed.; Intech Open: London, UK, 2018. [Google Scholar]
  21. Narayan, M.R.; 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, 073510. [Google Scholar] [CrossRef] [Green Version]
  22. Vollbrecht, J.; Brus, V.V.; Ko, S.; Lee, J.; Karki, A.; Cao, D.X.; Cho, K.; Bazan, G.C.; Nguyen, T. Quantifying the Nongeminate Recombination Dynamics in Nonfullerene Bulk Heterojunction Organic Solar Cells. Adv. Energy Mater. 2019, 9, 1901438. [Google Scholar] [CrossRef]
  23. Mohajeri, A.; Omidvar, A. Fullerene-based materials for solar cell applications: Design of novel acceptors for efficient polymer solar cells—A DFT study. Phys. Chem. Chem. Phys. 2015, 17, 22367–22376. [Google Scholar] [CrossRef] [PubMed]
  24. Ganesamoorthy, R.; Sathiyan, G.; Sakthivel, P. Review: Fullerene based acceptors for efficient bulk heterojunction organic solar cell applications. Sol. Energy Mater. Sol. Cells 2017, 161, 102–148. [Google Scholar] [CrossRef]
  25. Rad, H.M.; Zhu, F.; Singh, J. Profiling exciton generation and recombination in conventional and inverted bulk heterojunction organic solar cells. J. Appl. Phys. 2018, 124, 083103. [Google Scholar] [CrossRef]
  26. Barreiro-Argüelles, D.; Ramos-Ortiz, G.; Maldonado, J.-L.; Pérez-Gutiérrez, E.; Romero-Borja, D.; Meneses-Nava, M.-A.; Nolasco, J.C. Stability study in organic solar cells based on PTB7:PC71BM and the scaling effect of the active layer. Sol. Energy 2018, 163, 510–518. [Google Scholar] [CrossRef]
  27. Mousavi, S.L.; Jamali-Sheini, F.; Sabaeian, M.; Yousefi, R. Enhanced solar cell performance of P3HT:PCBM by SnS nanoparticles. Sol. Energy 2020, 199, 872–884. [Google Scholar] [CrossRef]
  28. Li, C.; Yue, Q.; Wu, H.; Li, B.; Zhu, X.; Zhu, X. Small bandgap non-fullerene acceptor enables efficient PTB7-Th solar cell with near 0 eV HOMO offset. J. Energy Chem. 2021, 52, 60–66. [Google Scholar] [CrossRef]
  29. Berger, P.R.; Kim, M. Polymer solar cells: P3HT:PCBM and beyond. J. Renew. Sustain. Energy 2018, 10, 013508. [Google Scholar] [CrossRef]
  30. Gurney, R.S.; Lidzey, D.G.; Wang, T. A review of non-fullerene polymer solar cells: From device physics to morphology control. Rep. Prog. Phys. 2019, 82, 036601. [Google Scholar] [CrossRef]
  31. Xu, X.; Yu, L.; Yan, H.; Li, R.; Peng, Q. Highly efficient non-fullerene organic solar cells enabled by a delayed processing method using a non-halogenated solvent. Energy Environ. Sci. 2020, 13, 4381–4388. [Google Scholar] [CrossRef]
  32. Hou, J.; Inganäs, O.; Friend, R.H.; Gao, F. Organic solar cells based on non-fullerene acceptors. Nat. Mater. 2018, 17, 119–128. [Google Scholar] [CrossRef]
  33. Reshma, L.; Santhakumar, K. Non-fullerene organic solar cells with 7% efficiency and excellent air stability through morphological and interfacial engineering. Org. Electron. 2017, 47, 35–43. [Google Scholar] [CrossRef]
  34. Xu, H.; Yuan, F.; Zhou, D.; Liao, X.; Chen, L.; Chen, Y. Hole transport layers for organic solar cells: Recent progress and prospects. J. Mater. Chem. A 2020, 8, 11478–11492. [Google Scholar] [CrossRef]
  35. Agnihotri, P.; Sahu, S.; Tiwari, S. Recent advances & perspectives in electron transport layer of organic solar cells for efficient solar energy harvesting. In Proceedings of the 2017 International Conference on Energy, Communication, Data Analytics and Soft Computing (ICECDS), Chennai, India, 1–2 August 2017; pp. 1568–1573. [Google Scholar]
  36. Van Le, Q.; Choi, J.Y.; Kim, S.Y. Recent advances in the application of two-dimensional materials as charge transport layers in organic and perovskite solar cells. FlatChem 2017, 2, 54–66. [Google Scholar] [CrossRef]
  37. Rasool, S.; Van Doan, V.; Lee, H.K.; Lee, S.K.; Lee, J.-C.; Moon, S.-J.; So, W.W.; Song, C.E.; Shin, W.S.; Shafket, R. Enhanced photostability in polymer solar cells achieved with modified electron transport layer. Thin Solid Films 2019, 669, 42–48. [Google Scholar] [CrossRef]
  38. Yang, Z.; Zhang, T.; Li, J.; Xue, W.; Han, C.; Cheng, Y.; Qian, L.; Cao, W.; Yang, Y.; Chen, S. Multiple electron transporting layers and their excellent properties based on organic solar cell. Sci. Rep. 2017, 7, 9571. [Google Scholar] [CrossRef] [Green Version]
  39. Lee, K.H.; Kumar, B.; Park, H.-J.; Kim, S.-W. Optimization of an Electron Transport Layer to Enhance the Power Conversion Efficiency of Flexible Inverted Organic Solar Cells. Nanoscale Res. Lett. 2010, 5, 1908–1912. [Google Scholar] [CrossRef] [Green Version]
  40. Sivakumar, G.; Pratyusha, T.; Gupta, D.; Shen, W. Doping of Hole Transport Layer PEDOT: PSS with Pentacene for PCDTBT: PCBM Based Organic Solar Cells. Mater. Today Proc. 2017, 4, 6814–6819. [Google Scholar] [CrossRef]
  41. Chien, H.-T.; Pölzl, M.; Koller, G.; Challinger, S.; Fairbairn, C.; Baikie, I.; Kratzer, M.; Teichert, C.; Friedel, B. Effects of hole-transport layer homogeneity in organic solar cells—A multi-length scale study. Surf. Interfaces 2017, 6, 72–80. [Google Scholar] [CrossRef] [Green Version]
  42. Sievers, D.W.; Shrotriya, V.; Yang, Y. Modeling optical effects and thickness dependent current in polymer bulk-heterojunction solar cells. J. Appl. Phys. 2006, 100, 114509. [Google Scholar] [CrossRef] [Green Version]
  43. Pettersson, L.A.; Roman, L.S.; Inganäs, O. Modeling photocurrent action spectra of photovoltaic devices based on organic thin films. J. Appl. Phys. 1999, 86, 487–496. [Google Scholar] [CrossRef]
  44. Hoppe, H.; Sariciftci, N.S. Morphology of polymer/fullerene bulk heterojunction solar cells. J. Mater. Chem. 2006, 16, 45–61. [Google Scholar] [CrossRef]
  45. Zhu, F. Semitransparent organic solar cells. Front. Optoelectron. 2014, 7, 20–27. [Google Scholar] [CrossRef]
  46. 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]
  47. Zhao, W.; Li, S.; Yao, H.; Zhang, S.; Zhang, Y.; Yang, B.; Hou, J. Molecular Optimization Enables over 13% Efficiency in Organic Solar Cells. J. Am. Chem. Soc. 2017, 139, 7148–7151. [Google Scholar] [CrossRef] [PubMed]
  48. Li, S.; Ye, L.; Zhao, W.; Zhang, S.; Mukherjee, S.; Ade, H.; Hou, J. Energy-Level Modulation of Small-Molecule Electron Acceptors to Achieve over 12% Efficiency in Polymer Solar Cells. Adv. Mater. 2016, 28, 9423–9429. [Google Scholar] [CrossRef] [PubMed]
  49. Lin, Y.; Adilbekova, B.; Firdaus, Y.; Yengel, E.; Faber, H.; Sajjad, M.; Zheng, X.; Yarali, E.; Seitkhan, A.; Bakr, O.M.; et al. 17% Efficient Organic Solar Cells Based on Liquid Exfoliated WS 2 as a Replacement for PEDOT:PSS. Adv. Mater. 2019, 31, e1902965. [Google Scholar] [CrossRef]
  50. Monestier, F.; Simon, J.-J.; Torchio, P.; Escoubas, L.; Flory, F.; Bailly, S.; De Bettignies, R.; Guillerez, S.; Defranoux, C. Modeling the short-circuit current density of polymer solar cells based on P3HT:PCBM blend. Sol. Energy Mater. Sol. Cells 2007, 91, 405–410. [Google Scholar] [CrossRef]
  51. Ompong, D.; Narayan, M.; Singh, J. Optimization of photocurrent in bulk heterojunction organic solar cells using optical admittance analysis method. J. Mater. Sci. Mater. Electron. 2017, 28, 7100–7106. [Google Scholar] [CrossRef]
  52. Anthopoulos, Y.F.T.D.; KAUST Solar Center (KSC), Thuwal, Saudi Arabia. Personal communication, 2020.
  53. Sopra, S.A. Optical Data from Sopra SA. Available online: http://sspectra.com/sopra.html (accessed on 31 December 2020).
  54. Optical Constants. Available online: https://www.jawoollam.com/resources/ ellipsometry-tutorial/optical-constants (accessed on 31 December 2020).
  55. Zhu, F.; Hong Kong Baptist University, Kowloon Tong, Hong Kong. Personal communication, 2017.
  56. Stelling. Refractive Index Database. 2017. Available online: https://refractiveindex.info/ (accessed on 31 December 2020).
  57. Kumar, P.; Jain, S.C.; Kumar, V.; Chand, S.; Tandon, R.P. Effect of illumination on the space charge limited current in organic bulk heterojunction diodes. Appl. Phys. A 2009, 94, 281–286. [Google Scholar] [CrossRef]
  58. Zhang, X.; Zhang, D.; Zhou, Q.; Wang, R.; Zhou, J.; Wang, J.; Zhou, H.; Zhang, Y. Fluorination with an enlarged dielectric constant prompts charge separation and reduces bimolecular recombination in non-fullerene organic solar cells with a high fill factor and efficiency > 13%. Nano Energy 2019, 56, 494–501. [Google Scholar] [CrossRef]
  59. Nam, Y.M.; Huh, J.; Jo, W.H. Optimization of thickness and morphology of active layer for high performance of bulk-heterojunction organic solar cells. Sol. Energy Mater. Sol. Cells 2010, 94, 1118–1124. [Google Scholar] [CrossRef]
  60. Ram, K.S.; Ompong, D.; Rad, H.M.; Setsoafia, D.D.Y.; Singh, J. An Alternative Approach to Simulate the Power Conversion Efficiency of Bulk Heterojunction Organic Solar Cells. Phys. Status solidi 2020, 2000597. [Google Scholar] [CrossRef]
  61. Kim, D.-H.; Park, M.-R.; Lee, H.-J.; Lee, G.-H. Thickness dependence of electrical properties of ITO film deposited on a plastic substrate by RF magnetron sputtering. Appl. Surf. Sci. 2006, 253, 409–411. [Google Scholar] [CrossRef]
  62. Schiefer, S.; Zimmermann, B.; Würfel, U. Determination of the intrinsic and the injection dependent charge carrier density in organic solar cells using the Suns-VOC method. J. Appl. Phys. 2014, 115, 044506. [Google Scholar] [CrossRef]
  63. Jeckelmann, B.; Piquemal, F. The Elementary Charge for the Definition and Realization of the Ampere. Ann. der Phys. 2019, 531, 1800389. [Google Scholar] [CrossRef] [Green Version]
  64. Steiner, R. History and progress on accurate measurements of the Planck constant. Rep. Prog. Phys. 2012, 76, 016101. [Google Scholar] [CrossRef]
  65. Porrat, D.; Bannister, P.R.; Fraser-Smith, A.C. Modal phenomena in the natural electromagnetic spectrum below 5 kHz. Radio Sci. 2001, 36, 499–506. [Google Scholar] [CrossRef]
  66. Pitre, L.; Plimmer, M.D.; Sparasci, F.; Himbert, M.E. Determinations of the Boltzmann constant. Comptes Rendus Phys. 2019, 20, 129–139. [Google Scholar] [CrossRef]
  67. Mei-Zhen, G.R.J.; De-Sheng, X.R.F.W. Thickness Dependence of Resistivity and Optical Reflectance of ITO Films. Chin. Phys. Lett. 2008, 25, 1380–1383. [Google Scholar] [CrossRef]
  68. Li, J.; Kim, S.; Edington, S.C.; Nedy, J.; Cho, S.; Lee, K.; Heeger, A.J.; Gupta, M.C.; Jr, J.T.Y. A study of stabilization of P3HT/PCBM organic solar cells by photochemical active TiOx layer. Sol. Energy Mater. Sol. Cells 2011, 95, 1123–1130. [Google Scholar] [CrossRef]
  69. Yuan, J.; Zhang, Y.; Zhou, L.; Zhang, G.; Yip, H.-L.; Lau, T.-K.; Lu, X.; Zhu, C.; Peng, H.; Johnson, P.A.; et al. Single-Junction Organic Solar Cell with over 15% Efficiency Using Fused-Ring Acceptor with Electron-Deficient Core. Joule 2019, 3, 1140–1151. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of a bulk-heterojunction organic solar cell (BHJ OSC) as a multilayer stack suitable for the optical transfer matrix method (OTMM). The electric field components of the incident and reflected electromagnetic waves in the lth layer are represented by E l + and E l , respectively, where l = 0 , 1 , j , m + 1 and j denotes the active layer.
Figure 1. Schematic diagram of a bulk-heterojunction organic solar cell (BHJ OSC) as a multilayer stack suitable for the optical transfer matrix method (OTMM). The electric field components of the incident and reflected electromagnetic waves in the lth layer are represented by E l + and E l , respectively, where l = 0 , 1 , j , m + 1 and j denotes the active layer.
Nanomaterials 11 00209 g001
Figure 2. Three BHJ OSC structures considered in the simulation: (a) OSC1—an inverted BHJ OSC with non-fullerene acceptor, (b) OSC2—a conventional BHJ OSC with non-fullerene acceptor, and (c) OSC3—a conventional BHJ OSC with fullerene acceptor. Layer 1 is the front electrode (FE), layer 2 is the front charge transport layer (FCTL), layer j is the active layer, layer 4 is the rear charge transport layer (RCTL), and layer 5 is the rear electrode (RE).
Figure 2. Three BHJ OSC structures considered in the simulation: (a) OSC1—an inverted BHJ OSC with non-fullerene acceptor, (b) OSC2—a conventional BHJ OSC with non-fullerene acceptor, and (c) OSC3—a conventional BHJ OSC with fullerene acceptor. Layer 1 is the front electrode (FE), layer 2 is the front charge transport layer (FCTL), layer j is the active layer, layer 4 is the rear charge transport layer (RCTL), and layer 5 is the rear electrode (RE).
Nanomaterials 11 00209 g002
Figure 3. The experimental values of the refractive index (nj) and extinction coefficient (kj) of the active layer materials P3HT:PCBM [51,55] and PBDBTSF:IT4F [49,52] plotted as a function of the wavelength of the incident light.
Figure 3. The experimental values of the refractive index (nj) and extinction coefficient (kj) of the active layer materials P3HT:PCBM [51,55] and PBDBTSF:IT4F [49,52] plotted as a function of the wavelength of the incident light.
Nanomaterials 11 00209 g003
Figure 4. Simulated Jsc of OSC1 of the structure Glass/ITO/ZnO/PBDB-T-SF:IT4F/MoO3/Al, OSC2 of the structure Glass/ITO/PEDOT:PSS/PBDB-T-SF:IT4F/PFN-Br/Al and OSC3 of the structure Glass/ITO/PEDOT:PSS/P3HT:PCBM/LiF/Al are plotted as a function of the thickness of the active layer along with their experimental results from [47,49,50,51], respectively.
Figure 4. Simulated Jsc of OSC1 of the structure Glass/ITO/ZnO/PBDB-T-SF:IT4F/MoO3/Al, OSC2 of the structure Glass/ITO/PEDOT:PSS/PBDB-T-SF:IT4F/PFN-Br/Al and OSC3 of the structure Glass/ITO/PEDOT:PSS/P3HT:PCBM/LiF/Al are plotted as a function of the thickness of the active layer along with their experimental results from [47,49,50,51], respectively.
Nanomaterials 11 00209 g004
Figure 5. Left column presents the contour plots of | E j ( x , λ ) | from Equation (5) in the active layer of OSC1 for three different thicknesses L j , (a) 35 nm, (c) 95 nm, and (e) 180 nm. The right column presents the corresponding G ˙ j ( x , λ ) contour plots from Equation (3) for the three thicknesses: (b) 35 nm, (d) 95 nm, and (f) 180 nm. The brighter spots on the left figures represent regions of the constructive interference of electric field of electromagnetic (EM) waves within the active layer and the darker spots represent regions of destructive interference. Likewise, the brighter spots on the right figures represent the positions of higher G ˙ j ( x , λ ) and darker spots that of lower G ˙ j ( x , λ ) .
Figure 5. Left column presents the contour plots of | E j ( x , λ ) | from Equation (5) in the active layer of OSC1 for three different thicknesses L j , (a) 35 nm, (c) 95 nm, and (e) 180 nm. The right column presents the corresponding G ˙ j ( x , λ ) contour plots from Equation (3) for the three thicknesses: (b) 35 nm, (d) 95 nm, and (f) 180 nm. The brighter spots on the left figures represent regions of the constructive interference of electric field of electromagnetic (EM) waves within the active layer and the darker spots represent regions of destructive interference. Likewise, the brighter spots on the right figures represent the positions of higher G ˙ j ( x , λ ) and darker spots that of lower G ˙ j ( x , λ ) .
Nanomaterials 11 00209 g005aNanomaterials 11 00209 g005b
Figure 6. (Left column) G ˙ j is plotted for two active layer thicknesses L j = 95 nm and 180 nm, as function of the thickness of (a) L1 of layer 1 (ITO), (c) L2 of layer 2 (ZnO), (e) L4 of layer 4 (MoO3), and (g) L5 of layer 5 (Al). (Right column) Contour plots of G ˙ j ( x ) as a function x within the active layer of optimal thickness L j = 95 nm, for other layers of thickness (b) L1 of layer 1 (ITO), (d) L2 of layer 2 (ZnO), (f) L4 of layer 4 (MoO3), and (h) L5 of layer 5 (Al).
Figure 6. (Left column) G ˙ j is plotted for two active layer thicknesses L j = 95 nm and 180 nm, as function of the thickness of (a) L1 of layer 1 (ITO), (c) L2 of layer 2 (ZnO), (e) L4 of layer 4 (MoO3), and (g) L5 of layer 5 (Al). (Right column) Contour plots of G ˙ j ( x ) as a function x within the active layer of optimal thickness L j = 95 nm, for other layers of thickness (b) L1 of layer 1 (ITO), (d) L2 of layer 2 (ZnO), (f) L4 of layer 4 (MoO3), and (h) L5 of layer 5 (Al).
Nanomaterials 11 00209 g006aNanomaterials 11 00209 g006b
Figure 7. (a) G ˙ j is plotted as function of the thickness L2 of layer 2 (PEDOT:PSS) in OSC2 for two active layer thicknesses L j = 100 nm and 180 nm. (b) Contour plots of G ˙ j ( x ) as a function of thickness L2 of layer 2 and position x within the active layer of OSC2 at the optimal thickness L j = 100 nm of the active layer.
Figure 7. (a) G ˙ j is plotted as function of the thickness L2 of layer 2 (PEDOT:PSS) in OSC2 for two active layer thicknesses L j = 100 nm and 180 nm. (b) Contour plots of G ˙ j ( x ) as a function of thickness L2 of layer 2 and position x within the active layer of OSC2 at the optimal thickness L j = 100 nm of the active layer.
Nanomaterials 11 00209 g007
Figure 8. G ˙ j ( x ) plotted as a function of the position x within the active layer of optimal thickness Lj = 95 nm and 100 nm for OSC1 and OSC2, respectively.
Figure 8. G ˙ j ( x ) plotted as a function of the position x within the active layer of optimal thickness Lj = 95 nm and 100 nm for OSC1 and OSC2, respectively.
Nanomaterials 11 00209 g008
Figure 9. (a) G ˙ j is plotted as function of the thickness L1 of layer 1 (ITO) for two active layer thicknesses L j = 80 nm and 180 nm, (b) contour plot of G ˙ j ( x ) as function of the thickness L1 of ITO and position x within the active layer of optimal thickness L j = 80 nm, (c) G ˙ j is plotted as function of the thickness L2 of layer 2 (PEDOT:PSS) for two active layer thicknesses L j = 80 nm and 180 nm, and (d) contour plot of G ˙ j ( x ) as function of the thickness L2 of PEDOT:PSS and position x within the active layer of optimal thickness L j = 80 nm.
Figure 9. (a) G ˙ j is plotted as function of the thickness L1 of layer 1 (ITO) for two active layer thicknesses L j = 80 nm and 180 nm, (b) contour plot of G ˙ j ( x ) as function of the thickness L1 of ITO and position x within the active layer of optimal thickness L j = 80 nm, (c) G ˙ j is plotted as function of the thickness L2 of layer 2 (PEDOT:PSS) for two active layer thicknesses L j = 80 nm and 180 nm, and (d) contour plot of G ˙ j ( x ) as function of the thickness L2 of PEDOT:PSS and position x within the active layer of optimal thickness L j = 80 nm.
Nanomaterials 11 00209 g009
Table 1. Input parameters required in the simulation.
Table 1. Input parameters required in the simulation.
P3HT:PCBMPBDB-T-SF:IT-4F
L j (nm)40–18040–180
ε r 3.0 [57] 3.40 [58]
μ n (m2V−1s−1)3 × 10−7 [59]6.27 × 10−9 [58]
μ p (m2V−1s−1)3 × 10−8 [59]2.26 × 10−8 [58]
ξ 1.3 × 10−3 [60]1.6 × 10−3 [58]
R C ( Ω cm 2 )3.0 [60]1.0 [60]
ρ ( Ω cm )4000 [60]1000 [60]
R s h e e t I T O ( Ω / sq )18.0 [61]13.7 [41,60]
L I T O (cm)0.5 [50]1 [47]
n3.2 × 1022 [62]
λ 1 (nm)300
λ 2 (nm)900
N t . b u l k ( m 3 )2.1 × 1019 [22]
N t . s u r f ( m 2 )6.3 × 1018 [22]
T (K)300 [25]
e (C)1.6 × 10−19 [63]
h (m2kgs−1)6.63 × 10−34 [64]
ε 0 (Fm−1)8.85 × 10−12 [65]
k (m2kgs−2K−1)1.38 × 10−23 [66]
Table 2. Optimised layer thickness for all three OSCs.
Table 2. Optimised layer thickness for all three OSCs.
OSC1OSC2OSC3
L1 (nm) 148 (ITO)148 (ITO)180 (ITO)
L2 (nm) 30 (ZnO)30 (PEDOT:PSS)45 (PEDOT:PSS)
Lj (nm)95 (PBDBTSF:IT4F)100 (PBDBTSF:IT4F)80 (P3HT:PCBM)
L4 (nm)10 (MoO3)5 (PFN-Br)1 (LiF)
L5 (nm)100 (Al)100 (Al)100 (Al)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sreedhar Ram, K.; Mehdizadeh-Rad, H.; Ompong, D.; Setsoafia, D.D.Y.; Singh, J. Characterising Exciton Generation in Bulk-Heterojunction Organic Solar Cells. Nanomaterials 2021, 11, 209. https://doi.org/10.3390/nano11010209

AMA Style

Sreedhar Ram K, Mehdizadeh-Rad H, Ompong D, Setsoafia DDY, Singh J. Characterising Exciton Generation in Bulk-Heterojunction Organic Solar Cells. Nanomaterials. 2021; 11(1):209. https://doi.org/10.3390/nano11010209

Chicago/Turabian Style

Sreedhar Ram, Kiran, Hooman Mehdizadeh-Rad, David Ompong, Daniel Dodzi Yao Setsoafia, and Jai Singh. 2021. "Characterising Exciton Generation in Bulk-Heterojunction Organic Solar Cells" Nanomaterials 11, no. 1: 209. https://doi.org/10.3390/nano11010209

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