Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

High performance and low cost transparent electrodes based on ultrathin Cu layer

Open Access Open Access

Abstract

Transparent electrodes based on an ultrathin Cu layer, embedded between two dielectrics, are optimized by simulations and experiments. Different dielectrics are screened in transfer matrix simulations for maximizing the broad-band transmittance. Based on this, sputtered electrodes were developed with the Cu embedded between TiOX-coated glass or PET substrate and an Al-doped ZnO (AZO) top layer. It is found that, for ultrathin Cu layers, increased sputter power fosters island coalescence, leading to superior optical and electrical performance compared to previously reported Cu-based electrodes. Simulations showed that the electrode design optimized with air as ambient medium has to be adapted in the case of electrode implementation in a hybrid perovskite solar cell of inverted architecture.

© 2017 Optical Society of America

1. Introduction

Transparent electrodes (TEs) are materials that show a high transmittance while simultaneously providing a high conductivity. TEs are key components in modern optoelectronic applications, e.g. flat panel displays, photovoltaics or flexible electronics [1]. The most common materials for TEs are transparent conductive oxides (TCOs) like tin-doped indium oxide (ITO), aluminium-doped zinc oxide (AZO) or fluorine-doped tin oxide (FTO). Especially ITO is widely employed due to its high optical transmittance and low resistivity [2]. The main drawback of ITO is the scarcity of the rare earth element indium, and its resulting high costs. There has been a growing research interest in alternative materials such as carbon nanotubes, graphene, conductive polymer films, metal nanowires, metal meshes and ultrathin metal (UTM) films being reported as the most promising candidates [3–7]. TEs that are applicable to modern and flexible optoelectronic devices (like thin film photovoltaic foils) must fulfil a number of challenging requirements, such as low deposition temperatures and a high stability against mechanical stress. Thus, the brittle and barely flexible TCOs, which usually require high process temperatures, are not well suited for this kind of applications [6].

Composite transparent electrodes based on ultrathin metals (UTMs) consist of ultrathin Ag, Au, Cu or Al embedded between two dielectric materials and have been studied for low emissivity coatings in the last decades [8,9]. They exhibit promising features, such as widely adjustable optical and electrical properties through film thickness and material variations, ambient-temperature deposition, mechanical stability on flexible substrates due to metal ductility [3,6,10], low roughness [11] and temperature stability [12]. The electrical conductivity of these trilayer electrodes is mainly dependent on the thickness of the metal layer, which must be below 10-12 nm in order to avoid significant optical losses. The optical transmittance can be further controlled by the choice and thickness of the dielectric materials. It is possible to adjust the transmittance range of the TE according to the characteristics of the optoelectronic device, e.g. the luminosity function for flat screen displays. This adjustment is often performed by trial-and-error like variation of the dielectric layer thicknesses, although various simulation methods [6,13–15] allow the careful design of the layer structure and thicknesses beforehand. A challenging task in the deposition of ultrathin metal films is to avoid nanometer-sized, island-like patterns instead of continuous films. Thus, the percolation threshold must be surpassed to reach high transparency and small sheet resistance values [6].

In this paper, we present a comprehensive study of dielectric/Cu/dielectric electrodes, as Cu is a low-cost alternative to the more commonly used Ag or Au. A great variety of dielectric materials, covering a broad range of refractive indices (nλ = 550 from 1.5 to 2.6), have been considered in transfer matrix method (TMM) simulations, so as to maximise the electrode transmittance. TMM allows transmittance optimization for selected wavelength ranges, depending on the application requirements. Based on the simulation results, an optimum trilayer structure of TiOx/Cu/AZO has been identified, which contains only low-cost materials. The experimental optimization of this trilayer on glass yielded superior figure of merit as compared to Cu-based TEs in the literature. It was further seen that the sputter power for Cu deposition greatly influences the TE performance for ultrathin metallic films and to similar extent improves the agreement between the simulated and experimental optical spectra. The paper also investigates the coating of this electrode on flexible polyethylene terephthalate (PET) substrate and compares its performance to the one on glass, revealing strong differences between the two cases. Finally, the Cu-based TE performance was simulated and evaluated in a complete hybrid-perovskite solar cell with inverted architecture, demonstrating the importance of in-device electrode simulation and optimization as opposed to simulations with air as ambient medium.

2. Methods

2.1 Simulation method

Optical simulations based on the transfer matrix method (TMM) [16,17] were performed to calculate direct transmittance and specular reflectance of coherent thin film multilayer stacks on a substrate (glass and PET). The substrate was considered incoherently by adding following expression for a non-absorbing substrate to the simulations [18]:

T(α)=TTM(α)T0(α)1RTM(α)R0(α)
R(α)=R0(α)+T02(α)RTM(α)1RTM(α)R0(α),
where T(α) and R(α) are the transmittance and reflectance of the overall structure, including the substrate. TTM and RTM are the transmittance and reflectance of the thin films calculated by the TMM. T0 and R0 are the values of the transmittance and reflectance at the air/substrate interface calculated by applying Fresnel equations. Throughout this work the light is always incident from the substrate side. The TMM allows further for calculation of the position-dependent complex electric field Ej(x) in a certain layer j [17,19], where x is the position along the surface normal of the multilayer stack. The position dependent absorption Qj(x) in a certain layer j of a multilayer device, was calculated by the following equation [19]:
Qj(x)=2πcε0njkjλ|Ej(x)|2,
with c the speed of light, ε0 the vacuum permittivity, nj and kj real and imaginary part of the complex refractive index, respectively. The position dependent absorption Qj(x) was further integrated over the layer of interest j and normalized to the incident light intensity and therefore represents the fraction of light which was absorbed in layer j.

2.2 Thin film deposition

The films were deposited by DC magnetron sputtering (Leybold Univex 450C) on soda-lime glass substrates (Menzel Gläser) and PET (Dupont-Teijin Melinex ST506). The substrates were cleaned in ultrapure water and isopropanol ultrasonication baths and blown-dry using nitrogen. TiOX was deposited by reactive sputtering from a 4-in Ti target in Ar/O2 (80/20 mixture) atmosphere at 120 W sputter power, yielding a sputter rate of 0.015 nm/s. Cu was deposited from a 3-in Cu target in pure Ar atmosphere at sputter powers of 20, 40, 80 and 100 W, corresponding to rates of 0.23, 0.45, 0.90 and 1.09 nm/s. Al-doped ZnO (AZO) was sputtered from a 4-in ZnO target with 2 wt.% Al2O3 in pure Ar atmosphere at 80 W, resulting in a deposition rate of 0.43 nm/s. All materials were deposited at a gas pressure of 0.1 Pa. The target-substrate distance was 10 cm. When the Ti target is reactively sputtered, this relatively large target-substrate distance increases the probability of oxidation of metallic species in the Ar/O2 plasma, yielding more stoichiometric oxide films, but at the expense of lowering the sputtering rate. This configuration is also dictated by the construction characteristics of the sputter tool. The substrates were water-cooled at 25°C during the deposition process to allow for controlled conditions and further avoid heat induced deterioration of the flexible substrates during deposition. The sputter chambers have been evacuated to base pressures of 1.9 – 7.0 × 10−6 Pa. A profilometer (KLA-Tencor-Alpha-Step IQ) was used to determine the layer thicknesses by step-height analysis. Annealing was performed on a hot plate in air. The samples were put on the plate prior to the ramp-up process (5-20 min depending on the final temperature) and remained at a constant temperature for 60 min before being removed immediately. In the following, the thickness of a specific layer is denoted with a subscript. For example, Cu7.5 stands for a Cu layer of 7.5 nm thickness.

2.3 Characterization

The samples were optically characterized using a Bruker Vertex 70 Fourier transform infrared spectrometer (FTIR), additionally equipped with a visible light source. Direct transmittance was measured for normal incident light and referenced to the sample holder without the sample (air). Total transmittance (direct plus diffuse transmittance) was measured with a teflon (PTFE) integration sphere accessory for the Vertex 70. Reflectance was measured with the A513QA accessory of the spectrometer, at 13 deg angle of incidence and unpolarized light. The measured signal was referenced to a calibrated mirror (Ocean Optics, STAN-SSH-NIST) in the range from 330 nm to 1150 nm. A GaP and Si detector were used to record spectra in the range of 330-550 nm and 550-1150 nm, respectively. The sheet resistance of the samples was determined by four-point, in-line probe technique (Süss MicroTec probes connected to an Agilent 4156C semiconductor parameter analyzer). The surface morphology of the Cu films was studied by scanning electron microscopy (SEM) (Zeiss, SUPRA 40) with 5 kV acceleration voltage and in-lens detector. Atomic force microscopy (AFM) (Molecular Imaging, PicoPlus) measurements of PET electrodes were performed in tapping mode.

3. Results and discussion

3.1 Simulation-based optimization

The general structure of the investigated Cu trilayer TE is shown in Fig. 1(a). The Cu is embedded between two dielectric layers to reduce reflection losses and prevent the ultrathin metal (UTM) from oxidation. In the following, a systematic investigation will be presented, suggesting which dielectric materials would lead to highest transmittance in a certain wavelength range. Based on comprehensive reviews on transparent electrodes [3,5,6], the following dielectric materials were considered for the simulation (sorted by increasing refractive index at λ = 550 nm, nλ = 550): AZO [11], ITO [20], ZnO [21], NiO [22], MoO3 [23], TeO2 [24], Nb2O5 [25], ZnS [26] and TiOX [13]. Furthermore, SiO2 [27] and MgO [28] were added to cover a broad range of refractive indices (nλ = 550 from 1.48 to 2.53). The real and imaginary (where available) parts of the refractive index (n and k, respectively) were retrieved from the literature. The optical properties of Cu were taken from literature [29].

 figure: Fig. 1

Fig. 1 (a) Sketch of transparent electrode layer design. (b) Simulated average transmittance in the range from λ = 400 nm to 900 nm for various material combinations with optimized dielectric layer thicknesses. The materials are ordered with increasing refractive index value at λ = 550 nm.

Download Full Size | PDF

In this study, the transparent electrode was optimized for maximum average transmittance T400-900 in the wavelength range of 400-900 nm. The average transmittance in this range was calculated for each material combination using the TMM algorithm. The thicknesses of the dielectrics were varied between 0 nm and 100 nm to find the optimal layer thickness, while the Cu thickness was initially fixed to 7.5 nm. The highest transmittance values with optimized layer thicknesses are depicted in Fig. 1(b). In general, the best transmittance values are achieved for a bottom layer with high refractive index (nλ = 550 = 2.3–2.4) and a top coating of medium refractive index (nλ = 550 = 1.7 –2.0). The highest average transmittance is 0.845 for TeO2;40/Cu7.5/MgO70. Being an insulator, MgO is not suited as the top layer that connects the electrode with the functional materials of the device (e.g. the absorption layer in solar cells). There are many possibilities to select the optimal material combination, as the transmittance is higher than 0.82 if nMat 1, 550 > 2.0 and 1.7 < nMat 2, 550 < 2.1.

The ideal thicknesses of the dielectric layers were in the range of 25 nm to 75 nm (if SiO2 is excluded). Figure 2 presents the differences of the layer thicknesses Δd = dMat 1dMat 2. For symmetric electrodes, employing the same material in both layers, the optimal thicknesses are for both layers identical. For asymmetric electrodes, the material with higher reflective index must be thinner than the lower refractive index material. The thickness difference increases with the difference of the refractive indices.

 figure: Fig. 2

Fig. 2 Layer thickness difference Δd = dMat 1dMat 2 of various material combinations for maximum average transmittance in the spectral range from 400 nm to 900 nm.

Download Full Size | PDF

Based on the screening of dielectric materials and the respective applicability to optoelectronic devices, a designed trilayer structure, consisting of TiOX as the bottom layer and AZO as the top layer, was investigated in more detail. A schematic view of the electrode is presented in Fig. 3(a). Three different Cu mass-equivalent thicknesses dCu, namely 5, 7.5 and 10 nm, were further investigated. These are typical values for UTM electrodes to keep the absorption losses in the UTM low [3]. For each dCu, the thicknesses dTiOx and dAZO of the TiOx and AZO layer, respectively, were determined by applying the TMM algorithm. dTiOx and dAZO were varied from 0 nm to 100 nm and for each thickness combination, the average transmittance T400-900 was calculated. The results are shown in Fig. 3(b) for the Cu7.5 case, yielding a maximum transmittance of almost 0.83. The obtained optimal trilayers are TiOx;30/Cu5/AZO65, TiOx;31/Cu7.5/AZO60 and TiOx;32/Cu10/AZO58.

 figure: Fig. 3

Fig. 3 (a) Sketch of transparent electrode layer design used in this study. (b) Simulated average transmittance of a Cu7.5 electrode for different thicknesses of the TiOX and AZO layers in the wavelength range of 400-900 nm.

Download Full Size | PDF

3.2 Experimental optimization on glass substrate

The experimental and simulated transmittance and reflectance spectra of the designed transparent electrodes are presented in Fig. 4 for three different Cu thicknesses. Otherwise identical samples were prepared with two different sputter power values of 40 W and 100 W for the Cu deposition. The simulated transmittance of the Cu5 electrode is above 0.80 in a broad wavelength range from 450 to 900 nm. The transmittance around 700 nm rises slightly with increasing Cu thickness, but at the same time the transmittance drop around 530 nm becomes more significant for larger Cu thicknesses. This wavelength corresponds to an energy of 2.33 eV, where light absorption is expected from d-band electron transitions [30]. For the Cu10 trilayer, the shape of the experimental transmittance curve agrees well with the simulated one, although the measured transmittance is lower by about 0.03. The same applies for the Cu5 and Cu7.5 electrodes at wavelengths below 600 nm. Above 600 nm, the transmittance of the samples deposited at 40 W drops significantly in comparison to the simulation, which predicts high transmittance (>0.85) in this range for both electrode designs. Increasing the Cu sputter power to 100 W improves the overall transmittance for all electrodes, most remarkably for thinner Cu films. The influence of the sputter power on the electrode performance will be discussed in more detail later.

 figure: Fig. 4

Fig. 4 Simulated (Sim.) and experimental (Exp.) transmittance T (solid) and reflectance R (dashed) spectra for two different sputter powers (P) and different Cu thicknesses for (a) TiOX;30/Cu5/AZO65, (b) TiOX;31/Cu7.5 /AZO60 and (c) TiOX;32/Cu10/AZO58.

Download Full Size | PDF

The measured reflectance spectra fit well to the simulation results. The reflectance is less than 0.15 over the whole spectral range of the simulation and reaches minimum values of 0.04, proving that the layer thicknesses obtained by the TMM suppress reflection losses effectively. Total transmittance (direct plus scattered transmittance) was additionally measured for samples with 40 W and 100 W (data not shown here), showing good agreement (within 0.01) with direct transmittance measurements and therefore indicating that no scattering is present. The difference between simulation and experiment, especially for samples prepared at 40 W, must be therefore solely due to increased absorption losses in the ultrathin Cu layer.

The optical and electrical properties compared to commercial available ITO (703192-Sigma Aldrich, with app. thickness 120-160 nm) are summarized in Fig. 5. The average transmittance in the range of 400-900 nm is presented in Fig. 5(a) for the simulated and fabricated electrodes. The simulated average transmittance decreases with increasing Cu thickness. The average transmittance of the 100 W electrodes shows the same behaviour and reaches 0.82 for both Cu5 and Cu7.5 electrodes. The results for the 40 W electrodes differ significantly from the simulation due to the low transmittance above 600 nm, especially at low Cu thicknesses.

 figure: Fig. 5

Fig. 5 (a) Average transmittance T400-900, (b) sheet resistance RS and (c) Haake’s figure of merit for T400-900 as a function of Cu thickness for 40 W (green) and 100 W (orange) Cu sputter power. The measured values of commercial ITO are also depicted.

Download Full Size | PDF

The sheet resistance RS (as shown in Fig. 5(b)) of the TiOx;30/Cu5/AZO65 electrode drops significantly from 89 Ω/sq (ohms per square) to 21 Ω/sq when the sputter power is increased from 40 W to 100 W. A moderate decrease is observed for the TiOx;31/Cu7.5/AZO60 sample where the Rs decreases to 11 Ω/sq. For the TiOx;32/Cu10/AZO58 electrode the sheet resistance is not influenced by the Cu sputter power and remains at 6.5 Ω/sq.

To quantify the performance of the transparent electrode, the figure of merit (FoM) for transparent electrodes ϕ=T10/RS, where T is the transmittance and RS is the sheet resistance [31], was evaluated. Figure 5(c) presents the FoM for the 40 W and 100 W electrodes taking into account the average transmittance, T400-900, from 400 nm to 900 nm. The TE with Cu thickness 10 nm yields the best FoM400-900 with 1.45 × 10−2 Ω−1. Compared to other Cu-based trilayer electrodes, this is one of the best reported FoM values for average transmittance. Song et al. [32] reported a FoMavg. of 1.94 × 10−2 Ω−1 for AZO40/Cu8/AZO40, and Wu et al. [33] recently reported FoM400-800 of 1.46 × 10−2 Ω−1 for a BaSnO3;50/Cu9/BaSnO3;50 electrode. However, both groups referenced the transmittance to the glass substrate transmittance, Tglass, which naturally yields higher FoM values. The FoM when referenced to glass can be given by the relation: ϕglassref=(T/Tglass)10/RS, which is related to the FoM used in this work by: ϕ=Tglass10ϕglassref. Assuming a transmittance value of Tglass≈0.92 over a broad wavelength range, a correction factor of Tglass10=0.43 can be deduced. Considering this, the Cu7.5 and Cu10 electrodes presented in this paper perform superior to the Cu-based electrodes aforementioned. The commercial ITO sample has an average transmittance T400-900 of 0.86 and a sheet resistance of 11 Ω/sq resulting in a FoM400-900 of 1.95 × 10−2 Ω−1. Cu-based trilayer electrodes were measured right after deposition and after being stored for several months in ambient conditions. No changes in the optical and electrical performance have been observed. It is therefore concluded that the dielectric top layer efficiently protects the Cu layer from oxidation, as has been also pointed out for Cu- and Ag-based transparent electrodes in literature [12,34–36].

As shown before, the conductivity and transmittance are lower for Cu deposition power of 40 W than for 100 W. To investigate the reason for this, bilayers of TiOx;30/Cu5 were deposited at different Cu sputter powers of 20 W, 40 W, 80 W and 100 W. The morphology of the Cu layer was studied with SEM and the images are presented in Fig. 6.

 figure: Fig. 6

Fig. 6 SEM images of TiOx;30/Cu5 bilayer samples at different Cu sputter powers and their respective sheet resistance.

Download Full Size | PDF

At 20 W the Cu film is separated into small islands. With increasing sputter power the small islands merge into bigger clusters. The meander-like trenches at 80 W between the Cu islands reduce significantly with increasing power and transform to elongated voids at 100 W. In addition, the sheet resistance of the bilayer electrode decreases strongly. The film sputtered at 20 W is non-conductive. The sheet resistance drops to 3.8 kΩ/sq at 40 W, further to 137 Ω/sq at 80 W, down to 53 Ω/sq at 100 W. Bilayer films of TiOx;30/Cu7.5 (SEM data not shown) also show meander-like trenches for low sputter powers, which gradually disappear to form a continuous, defect-free film at 100 W. For the TiOx;30/Cu10 samples, a practically continuous Cu film is obtained even at 40 W power, with minor differences to the case of 100 W deposition. These observations regarding the Cu morphology are reflected by the differences in the corresponding optical spectra in Fig. 4. The importance of the choice of sputter power on film growth is pointed out in literature and is associated with strong crystallinity, faster growth, faster nucleation and improved electrical properties compared to Cu films, which were deposited at lower sputter powers [37]. Let us note that the presented sheet resistance values of the bilayers are influenced by oxidation of the unprotected Cu layer. When the AZO layer is deposited atop, the carrier transport will be eased by the filling of the voids in the Cu film, giving rise to the observed lower sheet resistance of the trilayer structures.

The performance of the TEs can be further increased by post-deposition annealing. Figure 7 presents the effects of thermal annealing for 1 h on the average transmittance and the sheet resistance of the TiOx;31/Cu7.5/AZO60 electrode. The average transmittance (400-900 nm) rises by ~0.02 when the electrodes are annealed between 150° C and 250° C. The sheet resistance, on the other hand, varies within ~1 Ω/sq. Annealing at 300° C reduced the performance to the as-deposited state. The electrode annealed at 250° C yielded a FoM400-900 of 1.5 × 10−2 Ω−1.

 figure: Fig. 7

Fig. 7 Average transmittance from 400 nm to 900 nm (black circles) and sheet resistance (red squares) as a function of annealing temperature for TiOx;31/Cu7.5/AZO60.

Download Full Size | PDF

The significance of the TiOx bottom layer goes beyond optical properties, as it offers an excellent adhesion of Cu to the substrate, as verified by adhesive-tape tests, and can be sputtered with low roughness on glass. The root-mean-square roughness of TiOx is 0.75 nm, as measured by AFM. It should be noted that for certain applications of these electrodes, e.g. in low emissivity coatings, high processing temperature steps are required, which dictate the use of additional layers to avoid agglomeration and diffusion of Cu into adjacent layers [8,9].

3.3 Experimental implementation of trilayer structure on flexible substrate

The trilayer structure was further applied to a flexible PET substrate. Simulations indicate, that even though PET has a slightly higher refractive index than glass (nPET≈1.65, nglass≈1.5), the optimal layer thicknesses of the dielectrics, as well as the transmittance spectra stays virtually unchanged. Figure 8(b) shows the direct and total (direct plus scatter) transmittance and specular reflectance spectra of the electrodes on glass and PET. The transmittance decreases significantly on PET and is in average about 0.2 lower than on glass. The reflectance is increased throughout the investigated wavelength range. Total transmittance measurements with an integration sphere showed that in the high wavelength regime about 0.07 of the direct transmittance loss is caused by light scattering, while at 600 nm no light scattering occurs. Taking into account scattering and reflection of the trilayer on PET and glass in the range 550-1150 nm shows that more than 0.15 of the transmittance loss is caused by increased absorption in the Cu layer. Furthermore, RS increased from 11 Ω/sq on glass to an average of 340 Ω/sq on PET and showed high variations over the electrodes surface (σRs = 170 Ω/sq).

 figure: Fig. 8

Fig. 8 (a) Sketch of trilayer on different substrate configurations. (b) Direct transmittance (solid), total transmittance (circles) and specular reflectance (dashed) spectra of trilayer structure TiOx;31/Cu7.5/AZO60 on glass, PET and PET/AZO10 substrate.

Download Full Size | PDF

Former studies [10,12] of AZO/Au/AZO electrodes showed only minor property changes when the substrate was changed from glass to PET. Thus, the effects of an extra buffer layer of AZO between the PET and the trilayer structure was investigated. Simulations showed that an additional 10 nm AZO buffer layer had very little influence on the the transmittance of the electrode (< 0.02). The experimental results of a PET/AZO10/TiOx;31/Cu7.5/AZO60 electrode, and the respective spectra are shown in Fig. 8(b). The transmittance increased significantly, no light scattering was observed and the specular reflectance decreased. The total average transmittance of the TE remains still 0.08 lower than on glass substrate. With a sheet resistance of 17 Ω/sq and a FoM400-900 of 3.1 × 10−3 Ω−1 the electrode performed inferior on PET than on glass, but comparable to the PET/AZO50/Au/AZO50 electrodes [10]. The FoM is also similar to recently reported PET/ZnO60/Cu8/ZnO60 and PET/ZnO60/Cu(N)6.5/ZnO60 electrodes [34,38], but not reaching the best performing PET/ZnO60/Cu(O = 5.0%)8/ZnO60 [38].

To elucidate the microscopic reason for the deterioration of the performance compared to glass, Fig. 9 presents AFM images (500 × 500 nm2) of PET coated with TiOx and AZO/TiOx. It has been shown, that the roughness of the transparent electrode can have a severe impact on the performance of the TE [13]. The roughness of the PET alone has a root mean square (RMS) value of 4.0 nm, while the roughness of TiOx and AZO/TiOx is significantly higher with 7.8 nm and 10.3 nm respectively. This is at first surprising as the layer structure PET/AZO10/TiOx;30 with higher RMS roughness value shows superior optical and electrical performance.

 figure: Fig. 9

Fig. 9 AFM images of (a) PET, (b) PET/TiOx;30 and (c) PET/AZO10/TiOx;30 including the RMS roughness, autocorrelation length (ACL) and kurtosis.

Download Full Size | PDF

Therefore, it is suggested that the RMS value as a quantity is not sufficient to explain the effect and additional quantities such as kurtosis and autocorrelation length (ACL) are needed. The kurtosis is a measure of the height distribution sharpness, with kurtosis > 3 caused by sharp peaks (spiky surface) and kurtosis < 3 by mild peaks (bumpy surface) [39]. The ACL is a measure of the average distance between surface irregularities. A higher ACL indicates a greater distance between grains in the profile. Even though the RMS roughness of the TiOx directly on PET is lower compared to the case with the additional AZO buffer layer, the total height scale covers a bigger range. This is an indication for sharp features, such as spikes or steep valleys. This is confirmed by the kurtosis value of 4.0 and seen as deep and steep trenches in the surface. Further, the TiOx directly on PET shows a smaller ACL, which indicates smaller grains compared to the case with the AZO buffer layer.

The TiOx grown on PET/AZO10 features a higher ACL and wider trenches, which are less steep and deep as indicated by the kurtosis value of 2.5. This promotes a smoother film growth of the Cu layer and explains the improved electrical and optical properties of the PET/AZO10/TiOx;31/Cu7.5/AZO60 electrodes compared to the case where no AZO buffer layer is present.

3.4 Simulation of TE implementation in a perovskite solar cell

In the above, the optical properties of the TE were optimized with air as ambient medium. We need to consider though, that for implementation in a device, all subsequent layers need to be taken into account, due to partial reflection at multiple interfaces and therefore complex interference effects. This means concretely that the dielectric thicknesses that maximize the transmittance, when air is the ambient medium, will not generally be the same when all layers in the device are considered. This also underlines the importance of optical simulations in the optimization process for the device of interest. Recently, works have been carried out for the optimization of ultrathin metal based electrodes in organic solar cells [40–43].

In this section we consider the implementation of the TE in a solar cell with methylammonium lead tri-iodide (MAPbI3) absorber and the so-called inverted architecture, shown in Fig. 10(a). The TiOx/Cu7.5/AZO electrode replaces the standardly-used ITO and resides on the glass substrate, followed by a PEDOT:PSS (poly(3,4-ethylenedioxythiophene):polystyrene sulfonic acid) hole transport layer. On top of this, the MAPbI3 is deposited, followed by a double electron transport layer PCBM (phenyl-C61-butyric acid methyl ester)/ZnO. Finally, an Al cathode is considered. This solar cell architecture (with ITO as anode) has been experimentally investigated in the literature [44].

 figure: Fig. 10

Fig. 10 (a) Sketch of the considered perovskite solar cell, (b) simulated absolute absorption in the perovskite layer of the cell in the wavelength range from 400 to 800 nm for a 7.5 nm Cu layer.

Download Full Size | PDF

Naturally, the absorption in the solar absorber is the important quantity and it is calculated with the TMM method as outlined in the methods section. Literature values were taken for the complex refractive indices of PEDOT:PSS [45], MAPbI3 [46], PCBM [45], ZnO [47] and Al [48]. The absorption of the incident light in the wavelength range from 400 to 800 nm in the perovskite layer (excluding the absorption in the other layers), as a function of the TiOx and AZO layer thicknesses is shown in Fig. 10(b). A maximum of 0.66 can be observed at dTiOx = 25 nm and dAZO = 62 nm for a Cu thickness of 7.5 nm. These thickness values are close to the optimized values obtained for air as ambient medium. This is the outcome of the considered device architecture. For instance, the simulations show that if the PEDOT:PSS is omitted, the AZO thickness for maximizing the absorption in the perovskite approaches zero.

The electric field intensity distribution within the active layer shows an exponential decay for wavelengths below 500 nm, where the extinction coefficient of the perovskite is highest. With increasing wavelength and consequently lower extinction coefficient, the intensity profile shows oscillations due to interference effects. Simulations showed that the field intensity distribution is similar compared to standard ITO electrodes with thickness 200 nm. Let us note that in an optimization process with the goal of maximizing absorption in the active layer, also variations in the thickness of the other layers, such as the PEDOT:PSS and the perovskite absorber, can be considered. This could modify the optimal dielectric thicknesses of the TE. However, the goal of the simulations presented here is to demonstrate that, for typically used device layer thicknesses, it is important to consider in-device simulations to optimize the electrode, as the results can be significantly different from the case when air is the ambient medium.

4. Summary

In conclusion, the introduced sputtered transparent electrode of TiOx/Cu/AZO combines the advantage of being composed of low-cost materials, and having high average transmittance of 0.80 and a minimum sheet resistance of 6.5 Ω/sq (for a Cu thickness of 10 nm) when deposited on glass substrate. The electrode is therefore competitive to the standard ITO. It was shown that high sputter power was key to avoid island-like growth of the Cu film and obtain superior optical and electrical properties. The electrode deposited on PET is far from reaching its performance on glass, due to the roughness induced by the TiOx deposition. The performance, though, can be partly restored by the introduction of an additional AZO buffer layer. The obtained electrode (for a Cu thickness of 7.5 nm) has an average transmittance of 0.74 and a sheet resistance of 17 Ω/sq. Finally, it was underlined that the optimization of the electrode design has to take into account the precise layer architecture of the device. For the particular case of an inverted hybrid perovskite absorber solar cell, it was shown that the electrode design optimized with air as ambient medium had to be adapted.

Funding

Austrian Climate and Energy Fund (853603, 848929).

References and links

1. C. Brabec, U. Scherf, and V. Dyakonov, Organic Photovoltaics: Materials, Device Physics, and Manufacturing Technologies (John Wiley & Sons, 2011).

2. H. Liu, V. Avrutin, N. Izyumskaya, U. Ozgur, and H. Morkoc, “Transparent conducting oxides for electrode applications in light emitting and absorbing devices,” Superlattices Microstruct. 48(5), 458–484 (2010). [CrossRef]  

3. C. Guillen and J. Herrero, “TCO/metal/TCO structures for energy and flexible electronics,” Thin Solid Films 520(1), 1–17 (2011). [CrossRef]  

4. K. Ellmer, “Past achievements and future challenges in the development of optically transparent electrodes,” Nat. Photonics 6(12), 809–817 (2012). [CrossRef]  

5. C. G. Granqvist, “Electrochromics for smart windows: Oxide-based thin films and devices,” Thin Solid Films 564, 1–38 (2014). [CrossRef]  

6. K. Zilberberg and T. Riedl, “Metal-nanostructures - a modern and powerful platform to create transparent electrodes for thin-film photovoltaics,” J. Mater. Chem. A Mater. Energy Sustain. 4(38), 14481–14508 (2016). [CrossRef]  

7. M. Girtan, “Comparison of ITO/metal/ITO and ZnO/metal/ZnO characteristics as transparent electrodes for third generation solar cells,” Sol. Energy Mater. Sol. Cells 100, 153–161 (2012). [CrossRef]  

8. J. Szczyrbowski, A. Dietrich, and K. Hartig, “Bendable silver-based low emissivity coating on glass,” Sol. Energy Mater. 19(1-2), 43–53 (1989). [CrossRef]  

9. J. Szczyrbowski, G. Brauer, M. Ruske, H. Schilling, and A. Zmelty, “New low emissivity coating based on TwinMag (R) sputtered TiO2 and Si3N4 layers,” Thin Solid Films 351(1-2), 254–259 (1999). [CrossRef]  

10. T. Dimopoulos, M. Bauch, R. A. Wibowo, N. Bansal, R. Hamid, M. Auer, M. Jäger, and E. J. W. List-Kratochvil, ““Properties of transparent and conductive Al:ZnO/Au/Al:ZnO multilayers on flexible PET substrates,” Mater. Sci. Eng. B-Adv. Funct,” Solid-State Mater. 200, 84–92 (2015). [CrossRef]  

11. T. Dimopoulos, G. Z. Radnoczi, B. Pécz, and H. Brückl, “Characterization of ZnO:Al/Au/ZnO:Al trilayers for high performance transparent conducting electrodes,” Thin Solid Films 519(4), 1470–1474 (2010). [CrossRef]  

12. T. Dimopoulos, G. Z. Radnoczi, Z. E. Horváth, and H. Brückl, “Increased thermal stability of Al-doped ZnO-based transparent conducting electrodes employing ultra-thin Au and Cu layers,” Thin Solid Films 520(16), 5222–5226 (2012). [CrossRef]  

13. M. Bauch and T. Dimopoulos, “Design of ultrathin metal-based transparent electrodes including the impact of interface roughness,” Mater. Des. 104, 37–42 (2016). [CrossRef]  

14. H. Köstlin and G. Frank, “Optimization of transparent heat mirrors based on a thin silver film between antireflection films,” Thin Solid Films 89(3), 287–293 (1982). [CrossRef]  

15. P. Grosse, R. Hertling, and T. Muggenburg, “Design of low emissivity systems based on a three-layer coating,” J. Non-Cryst. Solids 218, 38–43 (1997). [CrossRef]  

16. E. Centurioni, “Generalized matrix method for calculation of internal light energy flux in mixed coherent and incoherent multilayers,” Appl. Opt. 44(35), 7532–7539 (2005). [CrossRef]   [PubMed]  

17. L. Pettersson, L. Roman, and O. Inganas, “Modeling photocurrent action spectra of photovoltaic devices based on organic thin films,” J. Appl. Phys. 86(1), 487–496 (1999). [CrossRef]  

18. M. M. A.-G. Jafar, “Comprehensive formulations forthe total normal-incidence optical reflectance and transmittance of thin films laid on thick substrates,” Eur. Int. J. Sci. Technol. 2, 214–274 (2013).

19. P. Peumans, A. Yakimov, and S. Forrest, “Small molecular weight organic thin-film photodetectors and solar cells,” J. Appl. Phys. 93(7), 3693–3723 (2003). [CrossRef]  

20. R. J. Moerland and J. P. Hoogenboom, “Subnanometer-accuracy optical distance ruler based on fluorescence quenching by transparent conductors,” Optica 3(2), 112–117 (2016). [CrossRef]  

21. W. L. Bond, “Measurement of the refractive indices of several crystals,” J. Appl. Phys. 36(5), 1674–1677 (1965). [CrossRef]  

22. J.-W. Park, K. N. Choi, S. H. Baek, K. S. Chung, and H. Lee, “Optical properties of NiO thin films grown by using sputtering deposition and studied with spectroscopic ellipsometry,” J. Korean Phys. Soc. 52(6), 1868–1876 (2008). [CrossRef]  

23. L. Lajaunie, F. Boucher, R. Dessapt, and P. Moreau, “Strong anisotropic influence of local-field effects on the dielectric response of alpha-MoO3,” Phys. Rev. B 88(11), 115141 (2013). [CrossRef]  

24. N. Uchida, “Optical properties of single-crystal paratellurite (Te O 2),” Phys. Rev. B 4(10), 3736–3745 (1971). [CrossRef]  

25. L. Gao, F. Lemarchand, and M. Lequime, “Exploitation of multiple incidences spectrometric measurements for thin film reverse engineering,” Opt. Express 20(14), 15734–15751 (2012). [CrossRef]   [PubMed]  

26. M. Debenham, “Refractive indices of zinc sulfide in the 0.405-13-[mgr ]m wavelength range,” Appl. Opt. 23(14), 2238–2239 (1984). [CrossRef]   [PubMed]  

27. L. Gao, F. Lemarchand, and M. Lequime, “Refractive index determination of SiO2 layer in the UV/Vis/NIR range: spectrophotometric reverse engineering on single and bi-layer designs,” J. Eur. Opt. Soc.-Rapid Publ. 8, 1-8 (2013).

28. R. E. Stephens and I. H. Malitson, “Index of refraction of magnesium oxide,” J. Res. Natl. Bur. Stand. 49(4), 249–252 (1952). [CrossRef]  

29. P. B. Johnson and R. W. Christy, “Optical constants of the noble metals,” Phys. Rev. B 6(12), 4370–4379 (1972). [CrossRef]  

30. R. T. Poole, “The colour of the noble metals,” Phys. Educ. 18(6), 280–283 (1983). [CrossRef]  

31. G. Haacke, “New figure of merit for transparent conductors,” J. Appl. Phys. 47(9), 4086–4089 (1976). [CrossRef]  

32. S. Song, T. Yang, M. Lv, Y. Li, Y. Xin, L. Jiang, Z. Wu, and S. Han, “Effect of Cu layer thickness on the structural, optical and electrical properties of AZO/Cu/AZO tri-layer films,” Vacuum 85(1), 39–44 (2010). [CrossRef]  

33. M. Wu, S. Yu, L. He, and L. Yang, “Preparation and investigation of nano-thick BaSnO3/Cu/BaSnO3 multilayer structures for transparent electrodes,” Mater. Lett. 174, 201–203 (2016). [CrossRef]  

34. G. Zhao, S. M. Kim, S.-G. Lee, T.-S. Bae, C. Mun, S. Lee, H. Yu, G.-H. Lee, H.-S. Lee, M. Song, and J. Yun, “Bendable solar cells from stable, flexible, and transparent conducting electrodes fabricated using a nitrogen-doped ultrathin copper film,” Adv. Funct. Mater. 26(23), 4180–4191 (2016). [CrossRef]  

35. A. Behrendt, C. Friedenberger, T. Gahlmann, S. Trost, T. Becker, K. Zilberberg, A. Polywka, P. Görrn, and T. Riedl, “Highly robust transparent and conductive gas diffusion barriers based on tin oxide,” Adv. Mater. 27(39), 5961–5967 (2015). [CrossRef]   [PubMed]  

36. D. S. Ghosh, T. L. Chen, N. Formica, J. Hwang, I. Bruder, and V. Pruneri, “High figure-of-merit Ag/Al:ZnO nano-thick transparent electrodes for indium-free flexible photovoltaics,” Sol. Energy Mater. Sol. Cells 107, 338–343 (2012). [CrossRef]  

37. M.-T. Le, Y.-U. Sohn, J.-W. Lim, and G.-S. Choi, “Effect of sputtering power on the nucleation and growth of cu films deposited by magnetron sputtering,” Mater. Trans. 51(1), 116–120 (2010). [CrossRef]  

38. G. Zhao, W. Wang, T.-S. Bae, S.-G. Lee, C. Mun, S. Lee, H. Yu, G.-H. Lee, M. Song, and J. Yun, “Stable ultrathin partially oxidized copper film electrode for highly efficient flexible solar cells,” Nat. Commun. 6, 8830 (2015). [CrossRef]   [PubMed]  

39. B. R. Kumar and T. S. Rao, “AFM studies on surface morphology, topography and texture of nanostructured zinc aluminum oxide thin films,” Dig. J. Nano-Mater. Biostructures 7, 1881–1889 (2012).

40. A. Bou, P. Torchio, S. Vedraine, D. Barakel, B. Lucas, J.-C. Bernede, P.-Y. Thoulon, and M. Ricci, “Numerical optimization of multilayer electrodes without indium for use in organic solar cells,” Sol. Energy Mater. Sol. Cells 125, 310–317 (2014). [CrossRef]  

41. A. Bou, P. Torchio, D. Barakel, F. Thierry, A. Sangar, P.-Y. Thoulon, and M. Ricci, “Indium tin oxide-free transparent and conductive electrode based on SnOx | Ag | SnOx for organic solar cells,” J. Appl. Phys. 116(2), 023105 (2014). [CrossRef]  

42. A. Bou, M. Chalh, S. Vedraine, B. Lucas, D. Barakel, L. Peres, P.-Y. Thoulon, M. Ricci, and P. Torchio, “Optical role of the thin metal layer in a TiOx/Ag/TiOx transparent and conductive electrode for organic solar cells,” RSC Advances 6(109), 108034 (2016). [CrossRef]  

43. A. R. Gentle, S. D. Yambem, G. B. Smith, P. L. Burn, and P. Meredith, ““Optimized multilayer indium-free electrodes for organic photovoltaics,” Phys. Status Solidi -Appl,” Mater. Sci. 212, 348–355 (2015).

44. W. Qiu, M. Buffiere, G. Brammertz, U. W. Paetzold, L. Froyen, P. Heremans, and D. Cheyns, “High efficiency perovskite solar cells using a PCBM/ZnO double electron transport layer and a short air-aging step,” Org. Electron. 26, 30–35 (2015). [CrossRef]  

45. H. Hoppe, N. Sariciftci, and D. Meissner, “Optical constants of conjugated polymer/fullerene based bulk-heterojunction organic solar cells,” Mol. Cryst. Liq. Cryst. (Phila. Pa.) 385(1), 233–239 (2002). [CrossRef]  

46. C.-W. Chen, S.-Y. Hsiao, C.-Y. Chen, H.-W. Kang, Z.-Y. Huang, and H.-W. Lin, “Optical properties of organometal halide perovskite thin films and general device structure design rules for perovskite single and tandem solar cells,” J. Mater. Chem. A Mater. Energy Sustain. 3(17), 9152–9159 (2015). [CrossRef]  

47. Y. Yang, X. Sun, B. Chen, C. Xu, T. Chen, C. Sun, B. Tay, and Z. Sun, “Refractive indices of textured indium tin oxide and zinc oxide thin films,” Thin Solid Films 510(1-2), 95–101 (2006). [CrossRef]  

48. A. D. Rakic, A. B. Djurisic, J. M. Elazar, and M. L. Majewski, “Optical properties of metallic films for vertical-cavity optoelectronic devices,” Appl. Opt. 37(22), 5271–5283 (1998). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (10)

Fig. 1
Fig. 1 (a) Sketch of transparent electrode layer design. (b) Simulated average transmittance in the range from λ = 400 nm to 900 nm for various material combinations with optimized dielectric layer thicknesses. The materials are ordered with increasing refractive index value at λ = 550 nm.
Fig. 2
Fig. 2 Layer thickness difference Δd = dMat 1dMat 2 of various material combinations for maximum average transmittance in the spectral range from 400 nm to 900 nm.
Fig. 3
Fig. 3 (a) Sketch of transparent electrode layer design used in this study. (b) Simulated average transmittance of a Cu7.5 electrode for different thicknesses of the TiOX and AZO layers in the wavelength range of 400-900 nm.
Fig. 4
Fig. 4 Simulated (Sim.) and experimental (Exp.) transmittance T (solid) and reflectance R (dashed) spectra for two different sputter powers (P) and different Cu thicknesses for (a) TiOX;30/Cu5/AZO65, (b) TiOX;31/Cu7.5 /AZO60 and (c) TiOX;32/Cu10/AZO58.
Fig. 5
Fig. 5 (a) Average transmittance T400-900, (b) sheet resistance RS and (c) Haake’s figure of merit for T400-900 as a function of Cu thickness for 40 W (green) and 100 W (orange) Cu sputter power. The measured values of commercial ITO are also depicted.
Fig. 6
Fig. 6 SEM images of TiOx;30/Cu5 bilayer samples at different Cu sputter powers and their respective sheet resistance.
Fig. 7
Fig. 7 Average transmittance from 400 nm to 900 nm (black circles) and sheet resistance (red squares) as a function of annealing temperature for TiOx;31/Cu7.5/AZO60.
Fig. 8
Fig. 8 (a) Sketch of trilayer on different substrate configurations. (b) Direct transmittance (solid), total transmittance (circles) and specular reflectance (dashed) spectra of trilayer structure TiOx;31/Cu7.5/AZO60 on glass, PET and PET/AZO10 substrate.
Fig. 9
Fig. 9 AFM images of (a) PET, (b) PET/TiOx;30 and (c) PET/AZO10/TiOx;30 including the RMS roughness, autocorrelation length (ACL) and kurtosis.
Fig. 10
Fig. 10 (a) Sketch of the considered perovskite solar cell, (b) simulated absolute absorption in the perovskite layer of the cell in the wavelength range from 400 to 800 nm for a 7.5 nm Cu layer.

Equations (3)

Equations on this page are rendered with MathJax. Learn more.

T(α)= T TM (α) T 0 (α) 1 R TM (α) R 0 (α)
R(α)= R 0 (α)+ T 0 2 (α) R TM (α) 1 R TM (α) R 0 (α) ,
Q j (x)= 2πc ε 0 n j k j λ | E j (x) | 2 ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.