Next Article in Journal
Effect of pH Hydrolysis on the Recovery of Antimony from Spent Electrolytes from Copper Production
Previous Article in Journal
Investigation of Oxidation Homogeneity in Asphalt Puck after Simulation of Long-Term Aging (Pressure Aging Vessel)
Previous Article in Special Issue
Understanding the Role of Paper-Ink Interactions on the Lightfastness of Thermochromic Prints
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Functional Materials for Fabrication of Carbon-Based Perovskite Solar Cells: Ink Formulation and Its Effect on Solar Cell Performance

by
Dena Pourjafari
1,
Nidia G. García-Peña
1,
Wendy Y. Padrón-Hernández
2,
Diecenia Peralta-Domínguez
1,
Alejandra María Castro-Chong
3,4,
Mahmoud Nabil
5,
Roberto C. Avilés-Betanzos
1 and
Gerko Oskam
1,6,*
1
Department of Applied Physics, CINVESTAV-IPN, Antigua Carretera a Progreso Km 6, Merida 97310, Yucatan, Mexico
2
Facultad de Ingeniería Química, Universidad Autónoma de Yucatán, Periférico Norte, Km 33.5, Chuburná de Hidalgo Inn, Merida 97203, Yucatan, Mexico
3
Faculty of Science, Universidad Autónoma de San Luis Potosí, Álvaro Obregón 64, Centro 78000, San Luis Potosi, Mexico
4
Engineering and Science School, Tecnológico de Monterrey, Avenida Eugenio Garza Sada 2501, Tecnológico, Monterrey 64700, Nuevo Leon, Mexico
5
Facultad de Ingeniería, Universidad Autónoma de Yucatán, Avenida Industrias No Contaminantes por Anillo Periférico Norte, Merida 97203, Yucatan, Mexico
6
Department of Physical, Chemical and Natural Systems, Universidad Pablo de Olavide, Carretera de Utrera Km 1, 41013 Seville, Spain
*
Author to whom correspondence should be addressed.
Materials 2023, 16(11), 3917; https://doi.org/10.3390/ma16113917
Submission received: 4 April 2023 / Revised: 4 May 2023 / Accepted: 16 May 2023 / Published: 23 May 2023

Abstract

:
Perovskite solar cells (PSCs) have rapidly developed into one of the most attractive photovoltaic technologies, exceeding power conversion efficiencies of 25% and as the most promising technology to complement silicon-based solar cells. Among different types of PSCs, carbon-based, hole-conductor-free PSCs (C-PSCs), in particular, are seen as a viable candidate for commercialization due to the high stability, ease of fabrication, and low cost. This review examines strategies to increase charge separation, extraction, and transport properties in C-PSCs to improve the power conversion efficiency. These strategies include the use of new or modified electron transport materials, hole transport layers, and carbon electrodes. Additionally, the working principles of various printing techniques for the fabrication of C-PSCs are presented, as well as the most remarkable results obtained from each technique for small-scale devices. Finally, the manufacture of perovskite solar modules using scalable deposition techniques is discussed.

1. Introduction

Since the first report in 2009 [1], the popularity of photovoltaic (PV) devices known as perovskite solar cells (PSCs) has skyrocketed due to their many advantages, such as the utilization of a low-cost, and earth-abundant hybrid lead halide perovskite with ambipolar transport properties [2], low-exciton binding energy, long free-charge diffusion length, bandgap tunability, and highly efficient light absorption [3,4].Remarkably, in less than 15 years, their power conversion efficiency (PCE) has increased from 3.81% [1] to 25.7% [5] in single junction cells, making them a promising complement to the currently dominating crystalline silicon-based solar cells [6]. The major drawbacks of conventional PSCs are related to low chemical stability, expensive materials, and toxicity. Degradation of the devices is caused by the chemical instability of the organic hole transport layer (HTL), metallic top electrode, and the hybrid perovskite layer upon exposure to UV or high temperature, and incorporation of oxygen and moisture during working cycles. For instance, the metal contact (gold, silver, or aluminum) may migrate into the perovskite material through the HTL after thermal stress [7]. To overcome such drawbacks and simultaneously improve the device efficiency, researchers have explored various methods, including the incorporation of electron/hole conducting nanoparticles (NPs) into the electron transport layer (ETL) and hole transport layer, modification of fabrication processes, engineering of the ETL/perovskite and perovskite/HTL interfaces, altering the perovskite precursor formulation, using hydrophobic HTL and top contact materials, and more effective sealing and encapsulation.
Among various architectures of perovskite solar cells (PSCs), the carbon-based perovskite solar cell (C-PSC) is an interesting architecture due to the low-cost, easy, scalable, and fully printable manufacturing process. The hydrophobicity of carbon materials as a top contact can significantly improve the stability of the photovoltaic device. This type of solar cell consists of a patterned transparent conductive oxide substrate (TCO), a blocking, compact layer (BL or CL), a mesoporous metal oxide as an electron transport layer (m-ETL), either a large bandgap metal oxide as an insulating separator layer or a hybrid perovskite layer, and a (mesoporous) carbon layer as a top contact; the mesoporous layers are impregnated with the hybrid perovskite absorber material. Depending on the cell architecture, one or two layers may or may not be present in the structure.
For instance, the m-ETL and/or the insulating layer are absent in the planar architecture, in which the BL acts as an ETL. Additionally, extra layers can be incorporated between the m-ETL and the perovskite layer, or between the perovskite and carbon layers to improve the device performance. C-PSCs are divided into two types: low-temperature and high-temperature. In low-temperature C-PSCs, a carbon paste with low curing temperature (<120 °C) is deposited after perovskite deposition and crystallization. In contrast, in high-temperature C-PSCs, a carbon paste is deposited and sintered at 400 °C. Then, after cooling down to room temperature, the perovskite precursor solution is deposited by drop casting and further annealed. Independent of the process fabrication temperature, the working principle of the C-PSCs is as follows: photogenerated electrons in the conduction band (CB) of the perovskite material are injected into the CB of the ETL metal oxide, while holes are extracted from the perovskite valence band (VB) by the carbon electrode. For efficient charge injection, transport, and extraction, the energy levels of the different functional layers in the cell structure must be compatible. Figure 1 shows the energy band positions of the fluorine-doped tin oxide (FTO) substrate, different electron extraction materials (ETMs), insulator layers, perovskite materials, and carbon top contact, which are mentioned in this review.
The C-PSC has achieved more than one-year stability [9,10]; however, the cell performance is not as good as other PSC architectures. Chen et al. [11], have explained several principal issues related to the lower performance of C-PSCs compared to conventional PSCs with HTL and metal contact. Briefly, there are three main reasons for lower performance of C-PCS, which are lower open circuit potential (VOC), slower hole transfer, and serious electron back transfer to the carbon electrode. The VOC is related to the difference between the electron and hole quasi-Fermi levels in the perovskite film. Various research has shown that the recombination losses occur mainly at the perovskite/ETM or perovskite/hole transport material (HTM) interfaces or close to the contacts rather than in the bulk. This can be interpreted that the electron and hole quasi-Fermi levels (Efn and Efp) in the perovskite film can be affected by band bending at the perovskite/ETM or HTM interfaces resulting in voltage losses. For the C-PSCs, the band bending can be affected by carbon electrode [12,13]. The Fermi level of carbon is usually higher than the highest occupied molecular orbital (HOMO) level of HTMs, thus the Efp in C-PSCs stands at higher energy resulting in a lower VOC. Slower hole transfer in C-PSCs is due to the lower hole selectivity at the perovskite/carbon interface. As shown in Figure 2, the difference between the Fermi-level of carbon and valence band level of the perovskite is larger than the difference between the HOMO level of the HTM and the VB level of the perovskite, which slows down the hole transfer step. In HTM-based PSCs, the lowest unoccupied molecular orbital (LUMO) level of the HTM is found at higher energy than the conduction band (CB) level of the perovskite, which could suppress the electron back transfer from the perovskite to the metal electrode. However, in C-PSCs, due to the positions of the energy levels, the charge separation rate at the perovskite/carbon interface is slower which may result in higher charge recombination and charge accumulation and, hence, larger hysteresis in the current density-voltage (J-V) curves.
Therefore, one of the main objectives of research on C-PSCs is to achieve higher VOC, and to enhance charge separation, extraction, and transport properties. These goals can be achieved by using novel ETMs, modifying the existing materials or using novel inorganic HTMs and/or cost-effective HTLs between the perovskite and carbon electrode.
One important ETM in solar cell structure is mesoporous inorganic metal oxide. The most studied mesoporous metal oxide layer is titania (TiO2) due to its extensive development in the field of dye-sensitized solar cells. TiO2 has a wide band gap, and it can be synthesized in different morphologies such as nanoparticles, nanorods, nanotubes, nanosheets, etc. Titania is commercially available in the market in the form of nanopowder, paste, and colloid and many research groups purchase it for the fabrication of electronic devices. Three big companies supply the commercial pastes typically used in third generation solar cells: Greatcell Solar Materials (formerly part of Dyesol) [14], Solaronix [15], and WonderSolar [16]. However, to improve the electronic properties of titania ETLs and enhance the TiO2/perovskite interface for a better match between the energy levels, researchers have synthesized titania nanopowders and have prepared homemade pastes. This allows to study the effect of different titania crystallographic phases, nanoparticle sizes and doping with metallic nanoparticles on the solar cell performance.
Another important layer in C-PSCs is the carbon electrode, which can be deposited using commercially available pastes. A common carbon paste consists of carbon black (CB), graphite, and organic binder material. Graphite powder provides the electrical conductivity of the carbon layer, and the morphology can be spherical or flake-shaped within the size range of 10–20 μm [17]. The carbon black nanoparticles typically have diameters around 30 nm and improve the interconnection between graphite sheets and, therefore, the hole extraction efficiency and conductivity [18]. The organic binders in the carbon paste provide adequate viscosity and adhesion of the carbon to the substrate. Commonly, ethyl cellulose, hydroxypropyl cellulose, and terpineol, are used as binder and adhesive, respectively. To enhance adhesion to the underneath separator layer (zirconia or alumina), small amounts of zirconia or alumina powder are usually considered in the paste formulation. To further improve the electrical properties at the perovskite/carbon interface, organic binders, and carbon NPs with different morphologies (such as single-walled and multi-walled carbon nanotubes [SWCNTs and MWCNTs], graphene, and carbon from biomass) have been incorporated [19,20,21,22,23,24,25]. Also, to enhance carbon conductivity, several researchers have added inorganic nanoparticles (NP) into their homemade carbon pastes [26,27,28,29,30].
To fabricate C-PSCs several printing techniques such as screen printing, inkjet printing, spin coating, slot-die coating, blade coating and spray coating have been used. Each technique has its own advantages and disadvantages. For example, the most efficient PSCs have been manufactured using spin coating; however, this technique is mainly suitable for small scale and batch fabrication. In contrast, screen printing, spray, blade coating and slot-die coating have been used in the manufacture of large photovoltaic devices, although with lower performance due to the imperfection and defects of the final deposited layers. The reports on fully printed devices by only one technique are few. Usually, for the fabrication of the entire C-PSCs more than one printing technique is used. This is due to the fabrication and operational restrictions of some printing techniques. For example, preparation of aqueous binder-free pastes suitable for screen-printing is still challenging and this technique is particularly used for deposition of non-aqueous viscous pastes. Another example is the inkjet printing technique in which the size of dispersed nanoparticles of the ink is critical to avoid nozzle blockage. Carbon pastes with micron-sized graphite sheets are not suitable to be used in this technique. In spin-coated devices, commonly all layers are deposited using this technique except the carbon layer which is deposited by screen printing, blade or brush painting and spray coating. In screen-printed devices the blocking layer is mostly deposited by spray pyrolysis and the perovskite layer by drop-casting. The choice of the printing technique highly depends on the rheological properties of the original ink, paste or colloid and the solvent in which the nanoparticles are dissolved or dispersed.
Fluid characterization may help with the right selection of printing technique. Different characterization techniques and equipment exist, such as rheology to determine fluid rheological properties (yield stress, relaxation time, thixotropic properties, etc.), viscosimetry to quantify fluid viscosity, and dynamic light scattering (DLS) to determine the size and size distribution of molecules and particles in a dispersion before layer printing. As mentioned before, these photovoltaic devices are composed of a stack of different hydrophilic (inorganic metal oxide) and hydrophobic (carbon electrode) layers. The light absorbing material, i.e., the hybrid perovskite layer must be deposited on and/or infiltrated into the layers. An appropriate interfacial contact between the perovskite and other layers is required for good performance of the device. To measure the wettability of the layers for the perovskite precursor solution, contact angle measurement is a useful technique that can be used after printing of the layers. The next step, after selection of a suitable ink, is to optimize the operational parameters of each printing technique in order to achieve uniform and pinhole-free layers.
In this review, we report on printing techniques for the deposition of functional materials for the manufacture of carbon-based perovskite solar cells. We provide important information on the principles of different printing techniques and their operational parameters. In the following sections, we focus on the extensive research literature reporting recipes and formulations for the preparation of pastes, inks, and colloids of different layers in C-PCSs. In this review, we consider neither the formulation of the perovskite precursor solution, which is covered in many reviews, nor the perovskite deposition method, which is generally drop casting or spin coating. We exclude those reports that use commercial pastes since it is beyond the scope of this review. We highlight research that achieved the fabrication of efficient C-PSCs by modification of homemade printing inks, pastes and colloids. Finally, we will mention the reports on fabrication of this type of solar modules and panels, emphasizing the scale-up opportunities and commercialization feasibility of these potentially low-cost devices.

2. Printing Techniques

2.1. Spin Coating

The spin coating technique produces thin, consistent coatings, which has enabled the quick manufacture of effective and reliable PSCs. An extensive variety of coating solutions and substrates can be used in this technique. The thickness of the coating can be controlled from a few nanometers to a few micrometers. The final coating properties such as thickness and homogeneity depend on the fluid rheological properties such as viscosity and thixotropy as well as surface tension. Other experimental parameters such as dispense volume, solution concentration, spin speed and time, and acceleration/deceleration rate impact the final coating properties [31,32].
The spin coating process consists of four main steps: deposition, spin-up, spin-off, and evaporation, which are shown in Figure 3 [33]. During the deposition step, a solution falls on a rotating substrate (dynamic spin coating) or fixed substrate (static spin coating) typically using a pipette or syringe. The solution spreads on the substrate due to the centrifugal or gravitational forces and, depending on its rheological properties, it may cover or wet the substrate completely. During the spin-up, the substrate is accelerated up to its desired rotational speed. At the very beginning, the substrate rotates faster than the fluid. While rotating, the fluid is radially driven out from the substrate due to the centrifugal forces. When the substrate reaches its desired speed, the fluid rotates at the same speed as the substrate and the film thickness decreases. In the third stage, the fluid rotates at a constant rate, flows out to the substrate perimeter, and continues thinning gradually. If the fluid contains a volatile solvent, usually it is possible to observe a color change in the deposited film. Finally, the fluid spin-off is stopped, and the thinning is dominated by solvent evaporation and film drying [32,33].
To our knowledge, the highest reported efficiency for HTM-free C-PSCs fabricated by the spin coating technique (except the carbon layer) is 17.42%, and corresponds to the following architecture: BL-TiO2/m-TiO2/MAPI/C(commercial paste) [34]. The highest efficiency for a cell with insulating layer is 9% and corresponds to BL-TiO2/m-TiO2/ZrO2 (homemade paste)/MAPI/C (homemade paste) structure [35].
As explained previously, fluid rheology has a strong influence on the quality of the deposited layer. For photovoltaic devices fabricated by spin coating, a choice of solvent for ink preparation is crucial and depends on the nature of the solute. For example, to deposit the blocking layer, a molecular ink is generally used in which the precursor is completely dissolved. For metal oxide mesoporous ETL or the insulator layer, a nano-ink is formulated in which the nanoparticles are dispersed into the appropriate solvent [36,37]. In the most common spin-coated C-PSCs, the insulating layer is not present, and the structure is FTO/BL or m-ETL/perovskite/carbon. However, as mentioned before, the carbon layer is usually not deposited using this technique. For this reason, in the following sections we present examples of reports in which the blocking and electron transport layers have been modified to achieve better performance. Finally, the work that reported strategies to improve the interface between perovskite and carbon are worth mentioning.

2.1.1. Blocking Layer

The research to improve the blocking layer goes back to the fabrication of dye-sensitized solar cells and is a well-known process. The blocking layer must be compact and pinhole-free. It is mostly TiO2, but sometimes SnO2 or ZnO are also used as a blocking layer [38,39]. A TiO2 compact layer (c-TiO2) usually is deposited from a normal TiO2 sol or a molecular solution containing a titanium precursor such as titanium isopropoxide (TTIP), titanium diisopropoxide bis acetylacetonate, titanium diisopropoxyacetate, titanium(diisopropoxide) bis(2,4-pentanedionate), and bis(2,4-pentanedionato)-bis(2-propanolato) titanium (IV) in ethanol, isopropyl alcohol (IPA), and 1-butanol [40,41,42,43,44,45,46,47,48,49,50]. To improve the electron transport properties of the blocking layer, doping with several metallic elements has been suggested [47,51,52,53]. Zhu et al. first deposited a compact titania layer of approximately 60 nm from a titania sol. After sintering, a solution of CsBr in deionized water (DI): IPA was directly spin-coated on the compact layer. The CsBr acts as a modifier resulting in an increase in the conduction band minimum (CBM) of the TiO2 ETL from −4.00 to −3.81 eV and, hence, a decrease in the work function, thus promoting favorable band alignment at the ETL/perovskite heterojunction. This doped compact layer suppresses recombination and improves charge carrier extraction and transport due to the band alignment [47]. In other work, Zhang et al. suggested Co/Eu doping of the titania compact layer. First, the TiO2 sol was obtained using a TTIP precursor solution. After the formation of titania nanoparticles, different molar ratios of Eu(NO3)3.6H2O/TiO2 and Co(NH3)6]Cl3/TiO2 were mixed, and spin coated on the FTO substrate to obtain a Co/Eu-TiO2 compact layer. They demonstrated that Co doping decreases the trap density in the TiO2 ETL and up-shifted the band edges of TiO2, while Eu doping increased the photo-response in the UV region. Combining them, the performance of the device with Co/Eu co-doped TiO2 was appreciably enhanced compared to the solar cells based on an undoped TiO2 BL, leading to a 26.4% increase in efficiency [51].

2.1.2. Electron Transport Layer

A suitable TiO2 ink for spin coating is commonly obtained from a commercial paste diluted in ethanol [40,41,54,55,56,57,58]. The paste consists of the spherical nanoparticles of titania anatase phase synthesized by the sol-gel method. However, many researchers have investigated the effect of different morphologies and particle sizes on solar cell performance by preparing homemade pastes. Acchutharaman et al. synthesized different sizes and porosities of TiO2 nanospheres via a solvothermal route by varying the concentration of acetic acid. Adequate porosity of the nanoparticles forming the ETL facilitates better crystallinity of the perovskite layer. The nanoparticle size (NPs) plays an important role in the back-scattering properties of the ETL. Larger particles scatter a larger number of photons, resulting in more photons that may be absorbed by the perovskite material and, therefore, the generation of a high photocurrent in the device. Also, lower degrees of crystallinity of the NPs was shown to lead to a larger number of electron-trapped oxygen vacancies. In this study, the titania layer with largest nanoparticles with an average diameter of approximately 17 nm and better light-harvesting properties, acts as a suitable scaffold to perovskite crystal growth resulting in a decrease in recombination and an increase in short-circuit current compared to nanoparticles with an average diameter of 12.5 nm. From impedance spectroscopy (IS), the highest value for the recombination resistance (Rrec) was obtained for the solar cells with the largest particles suggesting that these cells transfer the charge carrier effectively with very low non-radiative recombination at the interfaces of the device. Therefore, cells based on the ETL with 17 nm particle size show a lower recombination rate and higher VOC with a champion efficiency of 10.6% compared to the ETL with 12.5 nm particle size with a champion efficiency of 8.2% [59].
Another modification of the titania ETL in carbon-based perovskite solar cells is changing the crystalline phase. There are four commonly known polymorphs of titania found in nature, i.e., anatase (tetragonal), rutile (tetragonal), brookite (orthorhombic), and TiO2-B (monoclinic) [60]. Rutile is the most stable polymorph of TiO2 in the bulk at all temperatures, exhibiting lower total free energy than metastable phases of anatase and brookite. Anatase is the most used polymorph for solar cell applications because of its potentially higher conduction band edge energy and lower recombination rate of electron−hole pairs [61]. All these phases have been used in the structure of the PSCs as an ETL. Figure 1 shows the energy level of different phases in alignment with MAPbI3 [8].
There are various reports on the incorporation of the other titania phases into the solar cell architecture. Xiao et al. reported on the rutile passivation of an anatase TiO2 scaffold by introducing forming an anatase/rutile mixed phase system for carbon-based perovskite solar cells. The rutile phase was formed during the hydrolysis process by immersion of the anatase films, which were spin coated on the FTO substrate, into the TiCl4 aqueous precursor solution with different concentrations. The mixed phase exhibited higher conductivity, faster charge extraction, and reduced charge recombination by passivating defects at the interface of TiO2/perovskite films, which resulted in a higher open circuit voltage up to almost 1000 mV and champion efficiency of 15.21%. In addition, benefiting from the stable and better ultraviolet light filtration properties of the rutile phase, devices exhibited excellent stability maintaining 92% of their initial efficiency after 200 days of storage in ambient air [62].
Also, a brookite phase has been used in the structure of carbon-based perovskite solar cells and was reported in the work of Bhandari et al. [63]. Brookite nanorods and cubic nanoparticles were synthesized (inset of Figure 4), and the spin coating technique was used to deposit the brookite colloidal suspensions in water onto the substrate. X-ray diffraction spectra (XRD) was used to determine the phase purity of the synthesized material (Figure 4a). The peaks at 25.34°, 31.5°, 36.2°, 40°, 47.5° and 55.5° correspond to the (120), (111), (121), (012), (022), (231) and (244) planes respectively and are the major characteristics planes for brookite TiO2 (BTO). Raman spectra was conducted to confirm XRD analysis and dismiss the possibility of overlapping of anatase at 25.28° and reflection of brookite at 25.34°. Figure 4b shows that all Raman peaks match the literature in the range of 100 cm− 1 to 700 cm− 1 which clarifies the absence of anatase, rutile and any other titanium oxide or impurities. The film thickness was adjusted through the number of coating steps. The brookite-based cells reached a champion PCE of 15% and a higher open-circuit voltage of more than 1000 mV compared to the commercial anatase titania-based cells. The solar cells composed of the nanorod shaped brookite particles show better photovoltaic performance compared to those consisting of cubic particles and commercial titania. This may be due to the large surface area of nanorods, as illustrated by nitrogen physisorption measurements. The larger surface area enhances the cohesion of the ETL/perovskite contact, lowering the number of surface defect traps and refining the morphology of perovskite. This also was confirmed by incident photon-to-current conversion efficiency (IPCE) measurements showing better coverage in the region of ~500 to 780 nm and hence a higher electron transfer rate for brookite nanorod devices due to the better quality of the perovskite/brookite interfaces (see Figure 4c).
In order to obtain more details on the charge transfer (RCt), charge recombination (Rrec), and inner resistances they carried out impedance spectroscopy. According to the IS spectra (Figure 4d), a smaller loop in the high-frequency region was obtained for the nanorod shape brookite which suggests a lower RCt value indicating a high charge transfer rate between the transport layer and perovskite. This means higher values of FF and Jsc for nanorod brookite-based devices. On the other hand, a large loop with a high value of Rrec explains a lower rate of charge recombination, increasing the VOC of the device. Hence, they concluded that the type and morphology of the active layer greatly altered the electrical transport properties of the device demonstrating the importance of grain boundary trap passivation via structural modulation of nanomaterials, which can hugely improve the performances of devices [63].
As mentioned previously, replacing the commonly used titania ETL in C-PSCs with another metal oxide is one way to improve the cell performance and overcome several disadvantages of titania such as its low bulk electron mobility (<1 cm2 V−1 s−1), structural defects acting as charge traps, the high UV-photocatalytic activity resulting in decomposition of organic compounds, and high processing temperature [60,64].
Usually, a very thin layer of SnO2 is spun onto the conductive glass substrate from a SnO2 colloidal dispersion in deionized water [65,66,67,68,69,70], followed by spin coating of the perovskite layer. In many reports, the inorganic CsPbIBr2 perovskite is used in SnO2-based devices because of the better compatibility of its conduction band edge energy with inorganic perovskite (see Figure 1) [66,71]. There are different strategies to improve the performance of SnO2-based photovoltaic devices by modifying the SnO2 layer in order to engineer the charge transport properties at the tin oxide/perovskite interface.
For example, this can be achieved by improving the SnO2 porosity or reducing the surface defect density. He et al. showed that the introduction of nanopores into the SnO2 film can improve the performance of the carbon-based PSCs. They obtained the nanoporous SnO2 film by spin coating the tin oxide colloidal solution aided by monodisperse polystyrene microspheres (PS-MS). The nanoporous structure provides more electron-transporting pathways, increases the electronic transmission efficiency, and reduces the charge transport resistance (Rct) of the PSCs. In addition, the nanoporous structure improves the crystallinity of the perovskite upper layer and reduces defects in the film. Also, the nanoporous structure can change the refraction of the incident light, and the refracted light can be introduced into the perovskite film (Figure 5e,f). Therefore, nanoporous SnO2 improves the absorption capacity of the PSCs by reducing the reflection of the light [67].
To reduce surface defects, Qiang et al. obtained a tin oxide and lithium-doped tin oxide precursor solution via a simple low-temperature sol-gel method and subsequently deposited onto the substrate using spin coating. The authors compared the SnO2 with Li-SnO2 ETL in carbon-based perovskite solar cells. They explained that Li+ doping causes SnO2 lattice defects and forms a mixed valence state structure in which Sn2+/Sn4+ coexist, which can effectively improve the conductivity of SnO2. This approach offers two advantages over using a pure SnO2 ETL: it provides additional valence electrons to the SnO2 lattice and increases the electron extraction ability of ETL resulting in a 42.3% improvement in PCE compared to the non-doped SnO2 ETL [72]. In addition to TiO2 and SnO2, other inorganic electron transport layers have been explored in PSCs structure such as zinc oxide (ZnO), niobium pentoxide (Nb2O5), Zn2SnO4 (ZSO), BaSnO3 (BSO), amorphous InGaZnO4 and etc. [60,64,73,74,75,76].
ZnO is comparable to titania, with a similar energy band alignment with perovskite, but with a much larger electron mobility. Fu et al. proposed a feasible interface engineering method using deionized (DI) water to treat the FTO substrate (Figure 6) [77]. This enhances the wettability of the FTO substrate and improves the ZnO nanoparticles spreading on FTO. ZnO nanoparticles were obtained using low-temperature synthesis and were spun onto the DI water-coated-FTO substrate. Using contact angle measurements, they illustrated that the quality of the perovskite layer directly depends on the quality of the ZnO ETL, which in turn is significantly related to the quality of the FTO substrate. Using the pre-treatment of the substrate, they obtained a smooth and uniform perovskite layer with fewer pinholes and larger grains, with a champion PCE of 12.39% compared to the non-treated FTO substrate with PCE of 10.95%.
The last but not the least ETL modification in carbon-based perovskite solar cells is the incorporation of Nb2O5 and Li-doped Nb2O5. Zhao et al. reported a high open circuit voltage of 1.505 V using Li-doped Nb2O5 as an electron transport layer in C-PSCs [74]. They used spin coating method to dope a sputtered layer of Nb2O5 on FTO with lithium using acetonitrile solutions of 0.025 M, 0.05 M, and 0.075 M lithium bis(trifluoromethanesulfonyl)imide (Li-TFSI). As in previous work, they also mentioned the importance of the nature of the substrate on the quality of the perovskite layer. The inorganic CsPbBr3 average grain size on Li: Nb2O5/FTO was larger than on Nb2O5/FTO, which was 0.98 mm and 0.78 mm respectively (Figure 7). A better and uniform distribution of perovskite precursor solution on Li: Nb2O5/FTO resulted in an improvement in perovskite crystallinity, promoting light capture and charge transport and therefore, better performance of the device. From Hall effect measurements, the electron mobility and conductivity for Nb2O5 and Li: Nb2O5-based devices were found to be 15.7 and 33.8 cm2 V−1 s −1 and 1.5 × 10−5 and 2.6 × 10−5 S cm−1 respectively. This indicates that Li-doping increases the electronic conductivity of the amorphous Nb2O5, which facilitates better charge extraction and lower recombination in the photovoltaic devices.

2.1.3. Perovskite/Carbon Interface

The present section shows the strategies that have been proposed to improve the interface between the perovskite layer and carbon electrode by incorporating carbon nanoparticles or nanotubes as well as inorganic nanoparticles during perovskite deposition. In C-PSCs, the PCE depends not only on the energy level alignment between perovskite and carbon but also on the quality of the interfacial connection between these layers to guarantee a perovskite/carbon interface without pinholes. This can be achieved by adding carbon nanomaterials into the perovskite precursor solution in order to obtain a uniform, pinhole free and less rough perovskite layer with good optical quality and high crystallinity. In the following paragraphs we will report on several proposed strategies in which a spin coating ink of perovskite precursor mixed with carbon materials was formulated to achieve (1) an improvement in perovskite crystal growth, and (2) a better connection between perovskite crystals and the top carbon electrode deposited by blade coating.
In 2017, Zheng et al. deposited boron doped multi-walled carbon nanotubes (B-MWNTs) onto the PbI2 layer followed by annealing to obtain MAPbI3 perovskite film (See Figure 8a) [78]. The incorporated boron increases the work function, carrier concentration, and conductivity of MWNTs. Therefore, C-PSCs prepared with B-MWNTs showed an enhancement in the hole extraction efficiency and transport dynamics of the counter carbon electrode resulting a PCE improvement from 10.70% (cell with the undoped MWNTs) to 14.60% due to lower number of defects and higher conductivity. In 2019, Junshuai Zhou et al. added small amounts of multi-walled carbon nanotubes (MWCNT) (0.5% and 1.0% in volume fraction) to the FAI:MABr precursor solution [21]. The incorporation of MWCNTs as templates in the perovskite precursor helps to fill the boundary between perovskite grains and reduce the transformation kinetics of the perovskite, which benefits defect-free crystal growth. The results showed an improvement of the perovskite morphology, grain size and crystallinity with a long-term stability of 93% under ambient air conditions for 22 weeks.
Gao et al. used chlorobenzene as antisolvent with four different types of carbon materials (See Figure 8b), including CNTs (particle size, 5–15 nm), acetylene black (AB, particle size, 30–40 nm), nanocarbon (NC, particle size, 40–50 nm), and graphene (particle size, 3–5 µm) [79]. They found that the order of crystallinity was graphene > CNT > AB > NC, which meant that graphene had the best graphitization and that the NC was amorphous. As a result, the conductivity of graphene was better than for the other materials. However, the results indicated that the CNTs had a better interfacial contact effect and work function, leading to a PCE of 15.09%. They argue that the morphology of the interfacial bridging of the carbon materials played a more important role than the energy band alignment and conductivity. Also, carbon quantum dots (CQDs) have been used as interconnecting agents between perovskite and carbon electrode. In 2021, Li et al. fabricated C-PSCs using ethyl acetate including CQDs as anti-solvent (See Figure 9a) [80]. The CQDs were obtained by laser irradiation and the color change of the colloidal solution from dark to light yellow indicated the formation of the product. X-ray photoelectron spectroscopy (XPS) was used to reveal the surface chemistry of EACQDs as shown in Figure 9b. C 1s and O 1s were found in the whole spectrum of EACQDs. The deconvolution of C was conducted and displayed in Figure 9c. The peaks located at about 284.8, 285.8, 286.3, and 288.8 eV can be assigned to C–C, C=C, C–O, and C=O, separately. The EACQDs inside the spin-coated perovskite film passivate the defects, while perovskite crystallization results in a slower carrier recombination rate and faster carrier transportation, XRD results of pristine and EACQDs optimized PSCs are shown in Figure 9d.
The role of CQDs also was confirmed by Xu et al., who showed that the presence of this material in deposited perovskite layers results in the formation of larger size perovskite grains and excellent crystallinity [81].

2.2. Screen Printing

Screen printing is a widely used technique for imprinting designs onto a variety of substrates, such as paper, textiles, polymers, glass, or any solid material [82]. This technique involves pushing a viscous paste (or ink) through the orifices of a stencil screen with the help of a rubber blade (known as a squeegee) [83]. The procedure is shown in Figure 10. This method has a long history, with its origins believed to be in China and Egypt [84]. Nowadays, it is a particularly useful method for manufacturing scalable, high-speed printed microelectronics, as it is cost effective, easy to operate, and can be scaled up for larger production.
The key components of the screen-printing process that affect the resulting film are (i) the stencil, (ii) the squeegee, and (iii) the inks or pastes. The stencil is usually made of cotton, natural silk, plastic (nylon or polyester), or woven metal fibers; with plastic and metallic filaments being the preferred materials due to their durability and chemical resistance. The number of openings in the screen per linear inch (mesh count) also affects the quality of the pattern and the film thickness [83]. To create the stencil, a light-sensitive emulsion is applied to the mesh, with the net covering only the non-printable areas [82]. The squeegee is a blade made from a polymer with synthetic materials such as polyvinyl or polyurethane typically used for their sharpness and durability. The squeegee is used to control the spread of paste through the mesh and to adjust the film thickness [82]. The inks or pastes need to have the correct rheology to be able to infiltrate the mesh, and are composed of a solvent or solvent mixture, material particles, and organic binders. The proportions of these components depend on the type of substrate to be imprinted. Commonly used solvents are water and organic solvents, while binders are chosen to match the desired viscosity and functionality [85]. Despite the fact that a considerable number of studies have been conducted on screen printing of electronic materials, the operator must still adjust the deposition parameters and paste formulation in order to achieve the best possible coating. Here, we will look back at several studies related to paste formulation in C-PSCs.
In 2013, Ku et al., reported the first mesoscopic C-PSC using MAPI perovskite (Figure 11) [86]. This triple-stack C-PSC was printed using the screen-printing technique and the layers were composed of TiO2, ZrO2 or alumina, and carbon. As mentioned before, researchers usually use commercially available pastes such as TiO2 paste with particle size around 30 nm [14], and it is further diluted with terpineol in 1:1 to 1:7 ratios to attain the desired thickness. Commercial ZrO2 paste is composed of particles with diameters between 20 nm to 40 nm [15] and, depending on the stencil and dilution employed, the layer thickness can be changed from 1.5 μm to 3.0 μm. Carbon paste is also available commercially [15] and, as mentioned before, it consists of graphite and carbon black in different ratios. The carbon layer thickness must be in the range of 8 μm to 20 μm.
The following section is dedicated to homemade paste formulation for the fabrication of triple-stack mesoporous TiO2/ZrO2/C perovskite solar cells. The section is divided according to the separate layers, where we will show the formulation of each homemade paste in order to enhance the properties of each coating. Most of these formulations are based on the original work from the Han group, in which a corresponding metal oxide (5% to 20%), ethyl cellulose (2.5% to 10% concentration), and terpineol (70% to 92.5%) are mixed to reach an appropriate viscosity. Also, other components such as lauric acid, turpentine alcohol, acetylacetonate, or propylene glycol monomethyl ether acetate and organic binders such as hydroxyethyl cellulose, poly(vinyl pyrrolidone), or polyacrylic resin have been used in the paste formulation.

2.2.1. Blocking Layer

Spray pyrolysis is typically the method used to deposit the bottom layer (BL) of the triple mesoscopic architecture. However, recently screen printing has also been employed to deposit this layer, demonstrating fully screen-printed C-PSCs and the scalability of this method. Hvojnik and collaborators screen-printed a compact TiO2 BL using the commercial Solaronix paste and compared the performance of the screen-printed and spin-coated BLs in C-PSCs [87]. The spin coated BL had the best result, with a PCE of 7.8%. In a recent article, Wang et al. took a different approach, screen-printing the mesoscopic triple-layer from Wonder Solar on top of an In and Al doped compact-TiO2 BL [88]. A precursor solution of AVAxMA1-xPbI3 perovskite was then drop-cast on the structure, resulting in a PCE of 15.53% with the doped BL, and 13.49% with a typical c-TiO2.

2.2.2. Electron Transport Layer

Research dedicated to this layer is mainly focused in improving the surface area and chemistry of the TiO2 film, which results in better infiltration of the perovskite precursor solution, and thus better TiO2/perovskite electrical contact. In this sense, Liu et al. developed their own TiO2 paste composed of two different NP sizes to enhance the performance of C-PSCs [89]. The TiO2 paste was fabricated from a mixture of the oxide powder, ethyl cellulose, and terpineol in a 2:1:7 ratio, respectively. The TiO2 powder portion consisted of different mass fractions of 25 nm and 100 nm particles. The perovskite precursor solution was infiltrated into the triple-stack by one-step drop casting. Solar devices with 14.4 wt.% of 100 nm TiO2 particles achieved better performance with PCE up to 10.81%; in contrast, those with 25 nm-TiO2 particles reached 9.17%. Likewise, Zhao et al. developed a nano-TiO2 slurry synthesized from the metal-organic framework (MOF) MIL-125 [90]. TiO2 in the anatase phase was obtained by sintering the MIL-125 at 450° C for 4 h. Once the particles were obtained, the slurry was prepared by mixing the powder (5 wt.%) with ethyl cellulose (2.5 wt.%) and terpineol (92.5 wt.%) as binder agent and solvent, respectively, followed by ball milling. Finally, the (MA)1-x(5-AVA)xPbI3 perovskite precursor solution was drop-cast. They obtained devices with PCE up to 13.42%. Furthermore, they synthesized another TiO2 powder from the MOF MIL-125-NH2, which was employed to obtain a paste containing 4.6 wt.% TiO2 anatase, 91.4 wt.% terpineol, 3.3 wt.% ethyl cellulose, and 7 wt.% lauric acid [91]. Fully screen-printed mesoscopic structures were further infiltrated with MAPI precursor solution in DMF. The MIL-125-NH2 -based devices achieved a champion PCE up to 12.55%. It is claimed that the cell efficiency enhancement was the result of the large specific surface area, as well as the existence of special pores that made the MOF-derived TiO2 layer more favorable to the permeation of the perovskite precursor solution.
In addition to the highly used TiO2, BaSnO3 coatings have also been considered as ETL since it has a suitable optical band gap of 3.1 eV and high electron mobility of 70 cm2/V·s [92]. According to the report, the barium stannate nanomaterial was synthesized by co-precipitation and thermal annealing under nitrogen atmosphere. The paste was fabricated from a mixture of ethanol, turpentine alcohol, acetylacetonate, and ethyl cellulose. Additionally, they obtained a NP dispersion by ultrasound, which was treated by rotary evaporation to obtain the appropriate rheology. The paste was deposited by screen printing, followed by screen printing of the ZrO2 and carbon layers. With this triple-stack architecture, the authors achieved PCEs up to 14.77%. The results were ascribed to enhanced charge extraction efficiency and suppressed charge recombination due to the presence of extensive oxygen vacancies inside the particles.

2.2.3. Insulating ZrO2 Layer

Liu and co-workers, from the Han group, constructed the triple-stack mesoscopic architecture from the Greatcell m-TiO2 paste, diluted with terpineol (3:1.2), along their own previously developed ZrO2 and carbon pastes [86]. This scaffold was employed to test the critical thickness of ZrO2 insulating layer to achieve an efficient carbon-based mesoscopic perovskite solar cell [93]. The optimal thickness for spacer layer was determined to be 1 μm, with a PCE of 10.30%. Moreover, when the MAPbI3 precursor solution was drop-cast at 70 °C (solution and substrate temperature), 12.34% PCE was achieved. The cell improvement is explained by better infiltration at 70 °C, and better quality MAPbI3 crystal formation at that temperature.
Meng and collaborators fabricated their own pastes of mesoporous TiO2, ZrO2, Al2O3, and carbon, to evaluate the substitution of ZrO2 isolating layer with Al2O3, due to its lower cost and better availability [94]. TiO2 was prepared from a mixture of 15 wt.% nanosheets of TiO2, 6 wt.% ethyl cellulose, and 79 wt.% terpineol. Al2O3 and ZrO2 nanoparticles with diameters of 20 nm and 20–40 nm, respectively, were suspended in a mixture of 9.8 wt.% of the corresponding oxide, 1.98 wt.% of ethyl cellulose, 39.22 wt.% of terpineol, and 49 wt.% of absolute ethanol; this mixture was ball milled during 5 h, followed by rotary evaporation to obtain the appropriate rheology. Carbon paste was prepared according to the procedure developed by Rong et al. [95]. Finally, the (5-AVA)x(MA)1-xPbI3 precursor solution was drop-cast on top of the carbon-based architecture. The zirconia-based spacer solar cells achieved PCE of 8.72%; whereas the alumina-based devices showed an efficiency of 5.48%. These results were explained by the morphology of the films: the larger nanoparticle diameter of the ZrO2 material generates bigger pores that allowed better perovskite infiltration. The substitution of the ZrO2 scaffold layer with Al2O3 was further investigated by Xiong and co-workers [96]. The authors prepared their own pastes from recipes previously developed by Han’s group [86]. The mesoscopic TiO2 layer was screen-printed using a paste prepared with P25 Degussa powder. To obtain the Al2O3 layer, they deposited a thin coating of aluminum by vacuum techniques on top of TiO2 and sintered it at 500 °C for 40 min (scheme shown in Figure 12). Following, ZrO2 and C pastes were screen-printed according to the typical procedure. Infiltration was carried out from a (5-AVA)x(MA)1-xPbI3 precursor solution in γ-butyrolactone (GBL) by drop casting. Solar cells with 10 nm Al thickness and a 1 μm ZrO2 layer achieved the best results (PCE of 14.26%) along with a negligible hysteresis effect.

2.2.4. p-Type Mesoporous Metal Oxide: Hole Transporting/Extracting Materials

In triple-stack carbon-based solar cells, a low-cost, mesoporous and printable hole transport layer such as nickel oxide can be used between the spacer and carbon layers in order to improve the performance. All studies developed their own NiO paste using approximately the same recipe i.e., 5.0% to 15% of NiO, 3.0% to 7.0% of ethyl cellulose, and 76.0% to 90.0% of terpineol. Another p-type metal oxide that can work as hole extracting layer is Co2O3 [97]. Behrouznejad et al. developed their own pastes from commercial powders of TiO2 (30 nm) and ZrO2 (45 nm) from Sigma-Aldrich [98]. The carbon paste is composed of 5 wt.% carbon black, 5 wt.% Al2O3, 10 wt.% graphite, 20 wt.% ethyl cellulose, and 60 wt.% alpha-terpineol. They used NiOx (30 nm in diameter, from Sigma-Aldrich) paste that worked as a hole extraction layer. When the Al2O3 layer presented a thickness of 450 nm, which is the critical parameter in this architecture, a PCE up to 12.12% was obtained. The authors conclude that the NiOx layer, with both Ni3+ and Ni2+ species detected by XRD analyses of the initial nickel oxide (Figure 13), helps to increase the VOC up to 945 mV due to hole hopping from Ni3+ sites to Ni2+ sites. This layer shifts the hole Fermi-level downward, resulting in an increment in VOC. Following this strategy, Tao et al. developed a triple-stack architecture supplemented with a NiO/graphene coating as hole-transporting layer [99]. The authors employed a commercial TiO2 paste, and prepared their own Al2O3, NiO/graphene and C pastes. The Al2O3 and NiO/graphene slurries were obtained from a mixture of the corresponding oxide powder, ethyl cellulose and terpineol in a 2:1:27 ratio, diluted with ethanol, and ball milled for 10 h. In order to optimize the rheological properties, the ethanol was rotary evaporated. For NiO paste, different amounts of graphene were added and extra terpineol was added to fix the solid contents at 15%. Carbon paste was prepared from a blend of graphite, carbon black, ethyl cellulose, and ZrO2 with a 7:2:1:1 ratio, which was ball milled for 10 h. TiO2 paste was spin-coated at a speed of 2500 rpm for 30 s, whereas Al2O3, NiO and carbon layers were screen-printed to reach the stack architecture. The device was finally infiltrated by a two-step process with PbI2 in DMF, followed by immersion of the cell into a MAI in IPA solution. The best results were obtained for a NiO/graphene ratio of 100:1 that achieved a device efficiency of 11.72%. Graphene was added in small quantities since its sole purpose was to improve the conductivity and hole extraction ability of nickel oxide layer. Noteworthy, the authors determined that the addition of graphene into the NiO coating made the cell more resistant to humidity, hence it stabilized the cell.
Icli et al. developed a carbon-based mesoscopic substrate using a m-TiO2/MgO/NiO/carbon array [100]. They synthesized MgO, TiO2 doped with yttrium (Y:TiO2) and NiO doped with lithium (Li:NiO) by a methanol combustion method in a flame spray apparatus that guaranteed the incorporation of the Li+ and Y3+ cations into the corresponding oxide matrix. Then, all oxide pastes were prepared from a mixture of the corresponding oxide powders, ethyl cellulose, and terpineol in a 2:1:10 ratio. The carbon layer was deposited from commercial paste. The architecture was screen printed and infiltrated by a one step drop casting process. The best device performance was obtained for the stacked architecture with 500 nm, 500 nm, and 1 μm thickness for the TiO2/MgO/Li:NiO layers, respectively achieving a PCE of 9.63%. In contrast, when the Y:TiO2 was used in the stack, the efficiency lowered to just 2.78% due to formation of rutile phase that has worse device performance. Therefore, they concluded that lithium incorporation into the NiO matrix increased the charge collection and transport properties and reduced the overall resistance of the cell.
Bashir et al. employed this methodology to build the triple-stack mesoporous carbon-based scaffold [101]. Between the ZrO2 and carbon layers, an 80 nm thin sheet of Cu:NiOx/NiOx was inserted as hole conducting layer. The copper oxide paste was prepared by the mixture of Cu:NiOx or NiOx nanoparticles, with ethyl cellulose and terpineol in a 1:10:8 ratio, which was ball-milled, and rotary evaporated to obtain the desired viscosity. The MAPbI3 perovskite (5% AVAI) in GBL was infiltrated in one step. Devices of an active area of 0.09 cm2 and 0.8 cm2 were fabricated, and for a 0.8 cm2 cell a record efficiency of 12.79% was obtained; in addition, the cell was stable for up to 6 months without any encapsulation. Using the same approach, the authors also tested a thin layer of Co2O3 spinel to enhance the hole extraction efficiency in the solar cell [97]. The Co2O3 paste was the result of a mixture of 10 wt.% oxide, 10 wt.% ethyl cellulose, and 80% terpineol, which was diluted in ethanol, ground by ball milling, and finally rotary evaporated to obtain the appropriate rheological properties to be screen-printed. The initial paste generated layers with 340 nm in thickness, thus different dilutions with terpineol were performed to achieve different layer thickness. The layer resulting from dilution with 1:5 ratio achieved the best results (shown in Figure 14); the same architecture without the spinel layer, achieved PCE of 11.25%.

2.2.5. Carbon Layer

As mentioned before, due to inefficient hole extraction and the relatively low open-circuit voltage (VOC) of the carbon-based perovskite solar cells, significant efforts are focused on the engineering of the carbon layer. Duan and co-workers modified the carbon layer by the addition of boron-doped graphite, which was synthesized using 5 wt.% boron carbide (B4C) with the aim to improve the work function of carbon electrode from 4.81 to 5.10 eV [102]. The paste was prepared from a mixture of boron-doped graphite, carbon black, hydroxypropyl cellulose, ZrO2, and terpineol in a 13:4:2:2:60 ratio that was ball milled for 12 h. Then, the typical triple-stack architecture was screen printed and infiltrated with the (5-AVA)x(MA)1-xPbI3 precursor solution by drop casting. With this modification, the device presented an efficiency of up to 13.6%. In contrast, when the carbon layer was fabricated without boron, the resulting cell presented a PCE of up to 12.4%. The improvement in the work function facilitated hole extraction from the perovskite and shortened the PL lifetime, which may be translated into a longer carrier recombination lifetime. Chen et al. further modified the work function of the carbon counter electrode with chloride (C-Clx) to enhance the efficiency of the resulting carbon-based perovskite solar cell [103]. The graphite was modified via reaction of the pristine graphite (XFNano) with hydrochloric acid, washed, and later sodium hypochlorite (NaClO) was added to obtain the C-Clx. The resulting product was used to fabricate the carbon paste from a blend of 1 g carbon black, 3 g of the attained C-Clx graphite, 0.4 g ZrO2 powder, 0.5 g ethyl cellulose and 10 g terpineol. The mixture was ball milled for 12 h. The triple-stack architecture was screen printed and infiltrated with the typical drop casting method. The chlorinated carbon with the C-Cl0.4 formula achieved a PCE of 11.75%, which was 1.46 times higher than the pristine device. Chu et al. developed a double carbon layer to improve the conductivity of the electrode [104]. A bottom coating composed of carbon black was first deposited, which was designed to guarantee an intimate contact between the perovskite layer and the electrode. The upper layer, which was a 3:1 mixture of graphite: carbon black, ensured the proper conductivity of the coating. Polyacrylic resin and propylene glycol monomethyl ether acetate were used as binder and solvent, respectively, and ball milling was employed to properly homogenize the paste. With this design, the authors were able to ensure good conductivity with a device efficiency of 14.1%, a VOC of 1.01 V, and excellent stability. Likewise, Jiang a co-workers constructed a double-layer carbon electrode to take advantage of the work function of both layers to enhance the open-circuit voltage and to reduce the series resistance (Rs) of the carbon-based perovskite solar cell [105]. They prepared a counter electrode with two layers: a high-temperature mesoporous carbon layer (m-C) with high work function and large surface area as a first layer, which was responsible for charge extraction. The second layer was a low-temperature highly conductive carbon layer (c-C), which was responsible for charge collection (see Figure 15). Moreover, they modified another m-C paste by the addition of Co2O3, CuO, MoO3, and NiO p-type oxide powders. The m-C paste was prepared from a mixture of 2 g carbon black (30 nm particle size), 6.5 g graphite, 30 mL terpineol, 1 g ZrO2 (20 nm particle size) and 1 g hydroxypropyl cellulose, which was ball-milled for 10 h. Then the p-type m-C paste was obtained by adding different oxide ratios to the former paste. The highly conductive c-C paste was obtained from a blend of 1.5 g carbon black (30 nm particle size), 4.5 g graphite, 10 g terpineol, 1.83 mL titanium (IV) isopropoxide, and 200 μL acetic acid that was ball-milled for 10 h. The champion solar cell achieved 14.15% efficiency using modified m-C paste with 50% NiO as the lower carbon layer. In this structure the ZrO2 layer thickness was 3 μm.
Another work based on the screen-printed double-layer of the carbon electrode is reported by Gong and collaborators. They implemented a carbon black interlayer between the CsPbI2Br perovskite layer and the carbon electrode in a planar architecture [106]. The interlayer was fabricated with the mixture of carbon black, ethyl cellulose and terpineol, which was screen-printed with a 325 mesh screen to ensure a proper thickness. With this thin carbon black coating, the authors facilitated the hole extraction due to larger contact area and suitable energy band alignment in the perovskite—carbon black interface. In addition, the research revealed diffusion of carbon black, which resulted in the enhancement of the cell efficiency. Therefore, a PCE of 13.13% was achieved for the inorganic-based PSC with a thin interlayer of 4 μm. In later work, the authors implemented an extra interlayer composed of CuSCN, which afforded a record device efficiency of 15.70% [107].
Hatala et al. tested different compositions of carbon pastes to study its effect on the performance of the carbon-based perovskite solar cell [108]. They changed three parameters: the carbon black concentration, the polymer binder, and the solvents. Graphite powder with particle size < 20 μm, and conductive carbon black with average particle size of 45 nm were used. Ethyl cellulose or poly(vinyl pyrrolidone) as binder agents and diethylene glycol butyl ether, 4-hydroxy-4-methyl-2-pentanone (HMP) and ethyl alcohol as solvents were employed. The pastes were ball milled to achieve a homogeneous mixture. A higher amount of carbon black enhanced viscosity, whereas the lowest sheet resistance was observed with a 3:1 graphite—carbon black ratio. The total carbon amount increased to 35 wt.% resulting in an improvement in conductivity. In addition, a higher concentration of ethyl cellulose increased the viscosity of the paste and enhanced its conductivity. In contrast, a proper ethyl cellulose: poly(vinyl pyrrolidone) mixture resulted in less surface roughness of the layer.
Raptis and collaborators employed aluminium and copper grids integrated into the same double-carbon layer to enhance the conductivity of this coating [109]. They prepared their own low-temperature carbon paste from a mixture of ethyl cellulose binder (12.5%), a mixture of carbon allotropes (29.4%), and terpineol. The triple-stack mesoporous architecture was screen printed, annealed, and finally infiltrated with a (5-AVA)x(MA)1-xPbI3 precursor solution by drop casting. Then, aluminium or copper grids (25 μm thick) were firmly placed on top of the mesoporous carbon layer and the low-temperature carbon layer was deposited to allow the incorporation of the grid into the carbon layer underneath. When a copper grid was used in the carbon-based solar cell, up to 13.15% efficiency was obtained. In contrast, the Al grid resulted in worse performance. In a similar approach, Zhang et al. enhanced the hole extraction efficiency of the carbon electrode via addition of a liquid metal (GaInSn) dopant in the carbon layer, which may also enhance conductivity [110]. The LM-modified carbon paste was fabricated from the mixture of graphite, carbon black, ZrO2, and terpineol in a 15:5:4:75 ratio that was ball milled for 24 h. The liquid metal was added in different concentrations of 0.6, 1.2, 2.0, and 3.5% before the milling process. Then the triple-stack mesoscopic architecture was screen printed using the typical method, followed by drop casting of the (5-AVA)x(MA)1-xPbI3 perovskite precursor solution. The modified PSC with 1.2% liquid metal achieved an efficiency up to 13.51% and showed a negligible hysteresis in the J-V curves.
In another study, the effect of different types of carbon black on device performance was investigated [111]. The carbon paste was formulated from a mixture of graphite powder (10 μm flake size), carbon black, ZrO2, ethyl cellulose, and terpineol with a 7:1:1:3:23.27 ratio, which was ball-milled for 10 h at 300 rpm. Three types of carbon black, Vulcan carbon (75.1 nm, CABOT), mesoporous carbon (58.0 nm, Sigma-Aldrich) and Super P carbon (47.8 nm, TimCal) were tested, which have different surface area, pore volume, micropore area, and adsorption-desorption pore radius. The best results were achieved for Vulcan carbon, with a device efficiency of 12.55%. The authors conclude that the higher porosity and pore volume in Vulcan black carbon allowed decent wettability of the perovskite precursor solution, excellent infiltration, and strong adhesion within the device stacks. Further, Bogachuk et al. employed different graphite sources in the carbon paste to evaluate the importance of graphite in the sheet resistance of the carbon-based cathode [112]. The authors based their carbon paste formulation on Elcocarb B/SP from Solaronix. They used six different types of natural and synthesized graphite: Flaky (V14, Ito graphite Co., Tado-cho Kuwana-city, Japan), scaly (SPR150, Ito graphite Co., Tado-cho Kuwana-city, Japan), pyrolytic (PC30, Ito graphite Co., Ltd., Tado-cho Kuwana-city, Japan), amorphous (Oriental Industry Co., Ltd., Shenzhen, China), needle graphite (Oriental Industry Co., Ltd., Shenzhen, China), and artificial fine powders (UFG30, Showa Denko K.K., Tokyo, Japan). Among these graphite powders, scaly graphite showed better performance, as it achieved up to 14.28% due to better transport properties and lower series resistance losses. The same carbon paste recipe was used by Tsuji et al. to expand on the investigation with the implementation of extra amorphous graphite [113], which was compared with pyrolytic graphite. Once more, the latter graphite demonstrated better performance with an efficiency of 13.45%.
Phillips and co-workers developed their own carbon inks, with optimum graphite to carbon black ratio to enhance conductivity and hence performance [114]. They found that the best conductivity, with a resistivity of just 0.029 Ω cm, was obtained with a graphite to carbon black (20–50 nm) ratio of 2.6:1. This mixture was prepared in a 29.4% total carbon concentration and 70.6 wt.% for resin VINNOL 15 wt.% dispersed in 4-hydroxy-4-methylpenan-2-one. Later, the same group expanded their investigation to the effect of photonic flash annealing with subsequent compression rolling on carbon-based inks [115]. The authors prepared three different inks with three different compositions: ink with graphite only, graphite nanoplatelets and the previously developed combination of graphite with carbon black. The ink was formulated as vinyl resin VINNOL (Wacker Chemie AG) 15% in 4-hydroxy-4-methylpentan-2-one. The graphite only inks were formulated in a 22.5 wt.% carbon concentration and 77.5 wt.% dispersed polymer, whereas mixture of graphite and carbon black was prepared with a total carbon concentration of 29.4% (2.6 parts graphite to 1 part carbon black) and 70.6 wt.% resin in solvent. Initial mixing of powders in VINNOL was performed manually, followed by homogenization in a three-roll milling. Photonic annealing used for degradation of the binder and subsequent compression rolling resulted in a reduction in printed film thickness, roughness, and an improvement in particle orientation. These results provided better electrical performance for all printed inks.
Also, several research groups have obtained carbon materials from biomass and tested them in C-PSCs to analyze the device performance. Mali and collaborators fabricated a carbon hole transporting layer from aloe vera bio-carbon [116]. The biochar was obtained from the aloe vera gel that was carbonized under an argon stream at 1000 °C to ensure proper graphitization. Then, the carbon paste was manufactured simply by ball-milling, mixing the bio-carbon with chlorobenzene in a 1:2 wt.%/vol % ratio. With this simple manufacture, the resulting paste was appropriate for screen-printing on top of a typical m-TiO2/m-ZrO2 stack. The substrate was infiltrated by drop casting with a solution of MAI and PbI2 in γ-butyrolactone. This device presented an efficiency of up to 12.58%. For comparison, they also fabricated spiro-MeOTAD/Au-based spin-coated PSCs, which achieved only 3.2% higher PCE than carbon biomass-based C-PSC.

2.3. Inkjet Printing

Inkjet printing technology has been widely used in photovoltaic device fabrication since it allows to print the active materials on the solar cell stack [117]. Inkjet printing provides flexibility in thickness and shape, and it is considered a material-efficient technology. The key requirements for ink materials are low viscosity, fine particle size if the ink is a colloid or a suspension, and low volatility to avoid nozzle clogging [118,119]. Inkjet printing can be divided into two modes of operation, continuous (CIJ) and Drop-on-Demand (DOD). In a CIJ printer, charged droplets are controlled by an electrostatic field during the printing process. The imaging droplets pass through to the substrate while other droplets are deflected to an ink catcher for re-use (Figure 14a). A DOD printer ejects ink droplets when a pulse of voltage is applied, without any drop deflecting and catching. Generally speaking, the CIJ printer has a higher jetting frequency, and the DOD printer has a much simpler inkjet head structure [118,120]. Their working principles are schematically shown in Figure 16. There are two principal mechanisms of propelling ink drops in an inkjet printer head, piezoelectric propelling, and thermal bubble propelling. In the piezoelectric inkjet head, a voltage is applied to the piezoelectric pressure transducer to make it bend or change shape, which repels the ink out of the nozzle, as schematically shown in Figure 14b. Drop on demand inkjet printing can be used for the fabrication of C-PSCs by printing all the oxide layers in the stack as well as the organo-metal halide absorber [121,122,123].
In 2007, the first organic solar cell fully fabricated by inkjet printing technique was reported with low efficiencies from 3 to 4% [125,126,127]. In 2014, researchers started using inkjet printing technology to print the functional materials to fabricate PSCs [128,129,130]. Even though studies on inkjet-printed PSCs are few, substantial achievements have been reported [131,132]. Efficiencies higher than 21% have been recently reported for conventional metal electrode devices [133]. Here, we provide several examples of inkjet-printed C-PSCs and the effect of the ink or printing parameters on the device performance. In 2016, Hashmi et al. reported infiltration of the perovskite precursor solution using inkjet printing in the HTM-free carbon counter electrode-based PSCs [134]. The inkjet infiltration of the perovskite precursor solution was performed for a triple-stack (TiO2/ZrO2/Carbon) mesoporous substrate using the perovskite precursor ink containing 5-AVAI. They used 5-AVAI because it significantly slows down perovskite crystal growth before and after the deposition of the precursor ink, thus preventing the inkjet printer cartridge from clogging, and also provides an opportunity for precise patterning and controlled volume dispensing of precursor ink. The fabricated devices had an efficiency of 8.1% when infiltrated with perovskite precursor in an area of 0.16 cm2 using an inkjet printer, and after three weeks of storage in the dark under vacuum showed a high stability and a performance increase to 9.53%.
In 2020, Verma et al. achieved inkjet-printed carbon-based solar cells with four layers out of five, i.e., c-TiO2, m-TiO2, and m-ZrO2, as well as the perovskite precursor ink, using environmentally friendly non-halogenated solvents [135]. For the c-TiO2, they formulated an ink from titanium diisopropoxide bis(acetylacetonate) (TAA) using an appropriate dilution with binary mixtures such as ethylene glycol: iso-propanol (IPA), tetralin: lPA, terpineol: IPA, ethyleneglycol: ethanol, tetralin: ethanol, and terpineol: ethanol. The TAA/terpineol: xylene mixture 1:16 vol/vol yielded stable jetting as well as very homogeneous wet and dry film formation. The final ink composition had a viscosity of 2.1 mPas and a surface tension of 23.1 mN/m. The m-TiO2 ink was developed by diluting the commercial screen printing paste (Solaronix) into a terpineol: IPA mixture. For the m-ZrO2, the Solaronix paste was diluted with a binary solvent. The ink with 6 vol of screen-printing paste in 5:5 (vol:vol) of binary solvents gave stable jetting with no satellite formation as well as homogenous layer formation. The oxide layers and the carbon electrode were first sintered at elevated temperature and then infiltrated with the standard MAPI by inkjet printing at room temperature. After infiltration and drying, the devices were post-treated in damp heat for 100 h. As shown in Figure 17, this treatment increases the crystallinity of the perovskite layer. They obtained solar cell devices of an active area of 1.5 cm2 with an efficiency of 9.1%.
In 2021, Karavioti et al. provided a direct comparison between spin-coating and inkjet-printing techniques for the fabrication of fully ambient air-processed perovskite absorbent layers for C-based HTL-free PSCs [136]. The results showed that the inkjet-printed perovskite layer presented a discontinuous morphology due to a “coffee-ring” effect unlike in the case of spin-coating, where a uniform morphology was obtained. This is mainly because in inkjet-printing, the applied perovskite precursor solutions should be of a much higher concentration compared to the corresponding solutions used for spin coating. In inkjet printing, the solvent evaporation rate is lower than in the case of spin coating, leading to a poor crystal structure, whereas in the case of spin-coating most of the solvent is quickly removed through centrifugal forces. Another reason is related to the wettability of the substrate by the ink, in particular, over-wetting. Moreover, the FTO/c-TiO2/m-TiO2/perovskite system fabricated by inkjet printing presented a lower light-harvesting efficiency (LHE) compared to cells fabricated by spin coating. These differences had an impact on the electrical characteristics of the solar cells. The devices fabricated by the inkjet printing technique achieved an efficiency of 8.40% compared to 10% for the solar cells fabricated by spin coating.
Chalkias et al. devised a strategy based on the jet-ability of the ink and on the wettability of the substrate with the ink to reduce “coffee ring defects” that appeared during ambient air inkjet printing processing of the PSCs [137]. With different concentrations of the perovskite precursor ink (up to the limits determined by the solubility of the solutes in the solvent used), better charge transportation and recombination kinetics and an improvement of the external and internal quantum efficiency of the developed solar cells were achieved. C-based HTL-free PSCs using the optimized ink (1.8 M of perovskite precursor) reached an efficiency >12% for a cell with an active area of 0.34 cm2, which is among the highest reported values for the architecture c-TiO2, m-TiO2 and perovskite layers deposited by inkjet printing. Finally, by the combination of inkjet printing and screen printing perovskite sub-modules of 34.2 cm2 with an average efficiency value of 9.09% were developed to demonstrate the scalability of the technology. While the ink properties play an essential role in determining the solar cell performance, the fluid dynamics calculation of ink is crucial. However, inkjet printing of carbon and metal oxides for electrodes remains a challenge and requires further ink and process development.

2.4. Slot-Die Coating

Slot-die coating is an attractive, low-cost printing technique that is used for the deposition of the perovskite and other layers because of its high uniformity in large-scale production and thickness control over a broad range. Nevertheless, slot die coating has not been considered in the literature as a promising coating technique for C-PSCs [138]. As illustrated in Figure 18, a standard slot-die coating process includes an ink reservoir and slot-die head positioned towards the substrate, then the ink is pumped into the head using a syringe pump, with the ink forced out of a narrow slit along the length of the coating head. The ink then creates a liquid bridge between the substrate and slot-die head, and deposition happens upon moving the substrate across the coating head [138]. Slot-die coating covers inks within a wide viscosity window ranging from 1 to 10,000 mPa s, and final dry thickness from a few nanometers to tens of microns [139]. In this technique, the determination of the operational limits to set parameters such as coating speed, flow rate and coating gap must be selected carefully to avoid coating defects [140].
Khambunkoed et al. fabricated carbon-based methylammonium-free PSCs in 2021, utilizing amorphous zinc tin oxide (ZTO) as the ETL [141]. The device was produced with a homemade slot-die system, and the inks used were prepared from commercial materials. Amorphous ZTO was chosen due to its beneficial optical and electronic properties, such as high electrical conductivity, high electron mobility, and high transparency. The device had an optimum PCE of 9.92% when the ZTO thickness was around 48 nm. This PCE value was comparable to that of a PSC with spin-coated ZTO. The short circuit current density, open circuit voltage, and fill factor of the device were 18.86 mA/cm2, 0.94 V, and 56%, respectively.

2.5. Blade Coating

Doctor blade coating is one of the widely used techniques to produce thin films on large area surfaces. The operation principle ensembles the screen-printing approach. According to Figure 19, in the doctor blade process, the precursor solution is applied between the coating head and the substrate; then, the substrate is coated by moving the blade or knife across the substrate and the solvent is evaporated through heat treatment to obtain the film [142]. Furthermore, the controllable substrate temperature can impact the quality of the film. The inks/pastes used in these processes usually require large amounts of binders and thickeners to produce the high viscosities (1000–10,000 mPa s) required for reproducible and reliable production of films. Viscosities can be increased with the addition of polymeric additives such as glycerol or ethylene glycol or ethyl cellulose [143]. Coating thickness and speed play a vital role in determining interfaces quality. This technique is a relatively simple and scalable fabrication method for C-PSCs [144].
Syrrokostas et al. reported on the fabrication of C-PSCs with double-layered ZrO2 films using the doctor blade method [145]. This included the use of commercial zirconia nanoparticles of both nano and micro sizes, which acted as spacer and light scattering material to increase the light harvesting efficiency and yield an enhanced photocurrent density. The UV-Vis absorption spectra of the N_ZrO2 and DL_ZrO2 films (see Figure 20) reveal an increased absorbance, covering all the visible and the near-infrared region, from 400 nm to 800 nm, in the second case. The increased absorbance for the DL_ZrO2 film is ascribed to the higher scattering ability owing to the size of the particles. The presence of the large ZrO2 particles in the double layered spacer film results in an increase of the optical path length, due to the improved scattering efficiency; this resulted in a 30% improvement in efficiency.
Tian et al. reported further advances by producing C-PSCs using commercial materials and incorporating fullerene C60 as the ETL, in addition to the modification of the ITO substrate with a self-assembled monolayer of 3-aminopropyl triethoxysilane (APTES) [146]. This combination enabled interfacial charge extraction, ultimately resulting in reduced hysteresis in the J-V curve. They report a remarkable PCE of 18.64%, one of the highest reported for C-PSCs, which was found to be stable for over 3000 h. Bidikoudi et al. explored the use of different ammonium iodide (AI) precursors as additive and post-treatment agent in the multiple-cation mixed halide perovskite precursor solution, which enabled the deposition of homogeneous films of ZrO2 and carbon with commercial materials [147]. This led to an increase in power conversion efficiency by 33%, with further post-treatment with AI and ethyl-AI raising the efficiencies to 9.77% and 9.15%, respectively. The authors highlighted the importance of selecting the right materials for each part of the solar cell and the substantial influence of the precursor and post-treatment solutions in the C-PSCs fabrication process.

2.6. Spray Coating

Spray coating is an effective technique to produce ETL, HTL and other films, as well as the perovskite layer, on various kinds of substrates [130,148]. It is recognized as an attractive method for fabricating large-area, high-throughput, and low-cost PSCs and modules. The spray coating technique can be classified into different methods, including ultrasonic spraying, pneumatic spraying and electrospraying. The classification is based on the droplet dispersion over the substrate [149]. Figure 21 represents the spray coating method. During the spray deposition process, the hot substrate comes in contact with the precursor solution droplets; this leads to heating up and, consequently, drying, yielding faster and uniform crystallization and in turn, better surface morphology [150].
In 2017, Yang et al. reported a simple ultrasound spray deposition method to prepare pure MWCNT films to improve the contact at the perovskite/carbon interface [151]; the c-TiO, m-TiO2 and PbI2 layers were deposited on FTO/glass by spin coating. Commercial MWCNTs were dispersed in chlorobenzene to form a homogeneous ink, and the MWCNT layer was deposited onto the perovskite layer using a home-made compressed air gun-based spray system. The PCE of the cell with an active area of 0.08 cm2 was 14.07%. Moreover, they deposited a sub-monolayer of NiO nanoparticles prior to the MWCNT deposition to enhance the hole extraction efficiency and reduce the recombination rate obtaining up to 15.80% of efficiency, which is among the highest efficiency reported for C-PSCs.
Another interesting report was published by Wu et al. on the use of electrospray (ES) coating to fabricate C-PSCSs [152]. This technique utilizes a high voltage to atomize a flowing solution into charged micro-droplets and can be used to deposit any layer in the perovskite solar cells. ES is distinguished from other techniques due to the advantages of versatility, scalability, designability, and low materials waste [153]. Wu et al. developed an electrospray deposition technique for continuously printing C-based HTL-free PSCs. All the layers, i.e., TiO2, perovskite and carbon were continuously printed through ES in the ambient atmosphere at low temperature. Commercial precursor solution, paste and solvents were used to prepare the dispersions. The fully ES printed carbon-based PSCs showed VOC of 1.03 V, JSC of 24.35 mA cm−2, FF of 57.7% and PCE of 14.41%. Even though this technology is not widely used for perovskite fabrication, the electrospray-assisted technique is promising for C-PSCs commercial manufacture.

3. Scale-Up

Although the PCE of PSCs is high enough in small areas, it is necessary to fabricate large-scale perovskite solar modules for future commercialization of this PV technology. However, scientific challenges, technical and techno-economical issues are magnified while increasing the substrate size for the fabrication of larger devices. For a successful scale-up, novel abundant and non-toxic materials, new device architectures, easy and low-cost fabrication processes, and adequate choice of printing technique compatible with the ink formulations and substrate size are required. In addition, in a production line of large-scale devices, all manufacturing steps should be compatible and match with a continuous flow process to avoid a slow production rate and, hence, higher manufacturing cost [154,155]
One example of a conflicting fabrication technique with large-scale substrates is spin coating. Unfortunately, spin coating technique works well for small devices with less than 1 cm2. Another problem that is magnified when up scaling the perovskite PV technology is lead toxicity. Still, lead-based perovskite solar cells possess by far the highest efficiencies. To overcome this issue, investigation on lead-free perovskite PVs and effective encapsulation of the entire device should be the important subjects of future research. As mentioned before, the unstable metal top electrode and degradation of perovskite material, which are the main reasons of device instability, have motivated researchers to find stable top electrode alternatives, such as carbon, and new perovskite formulations. Han et al. demonstrated that the carbon-based perovskite solar cell as a fully printable architecture is a promising candidate for large-scale applications related to the easy and cost-effective manufacturing process using stable and abundant materials [86]. Although the carbon-based perovskite devices have shown potential for scale-up, one barrier for large-scale application is the multiple sintering processes required in their manufacture. The sintering process helps to eliminate all organic binders from the printing pastes and results in a better interconnection between nanoparticles. However, it is costly and time-consuming, which reduces the production throughput. One solution is developing low-temperature processing materials such as aqueous-based inks and low-temperature carbon pastes. Unfortunately, the efficiency of low-temperature-based PSCs is still low. This could be due to the adhesion problems of low-temperature and aqueous-based inks to the substrate and poor particle interconnections. There are different deposition methods, including slot-die coating, screen printing, spray coating, roll-to-roll printing that can be used for the fabrication of large area C-PSCs. Rong et al. compared different fabrication methods for large-scale fabrication of PSCs and suggested that among all these methods, screen printing and slot-die coating would be the most promising methods [156]. In the following section we present several examples of large-area C-PSCs fabricated by different printing techniques.

3.1. Screen-Printed Modules

Screen printing is the most employed method to build modules and mini-modules because of the easy scalability of this technique. Priyadarshi et al. first fabricated mini-modules of 70 cm2 of active area using the screen printing technique [157]. The device was fabricated employing the commercial pastes for compact TiO2 (Greatcell), mesoporous TiO2 (Greatcell), mesoporous ZrO2 (Solaronix) and two homemade carbon pastes consisting of graphite and carbon particles of less than 500 nm in size. The two carbon pastes had different amounts of ethylcellulose as binder agent. They found that lower concentration of binder enhanced the conductivity of the layer. A third commercial carbon paste (Greatcell) was used to compare with the homemade pastes. The best result was a PCE of 10.03% corresponding to the commercial carbon. Later, the same group improved the previously fabricated mini-modules by inserting an 80 nm thin sheet of Cu:NiOx/NiOx as hole conducting layer between ZrO2 and carbon layers [101]. The MAPbI3 perovskite precursor with 5% 5-AVAI in GBL solvent was infiltrated in one step. Large-area devices with 70 cm2 of active area were fabricated reaching stable performance for up to 6 months without any encapsulation. For these large-area devices the precursor solution of MAPbI3 perovskite (3% 5-AVAI) was deposited in one step by slot-die coating with a width of 200 mm for the slot die head. In contrast, when a thin layer of spinel Co2O3 was used in these mini-modules, a device efficiency of 11.39% was obtained [97].
In a recent paper, the group further improved the performance of the 70 cm2 mini modules [4]. They modified the formulation of the TiO2 commercial paste HPX-100 (Cristal) with variable amounts (from 4.0% to 5.0%) of CsI, CsBr, and CsCl. To compare the efficiency of the homemade pastes, the authors tested the commercially available HPX100CsI and HPX100CsBr, which are modified with CsI and CsBr, respectively, and the typical commercial paste from Greatcell. A PCE of up to 11.55% was obtained for the TiO2 paste modified with 4% CsBr. The resulting mini-modules had negligible J-V curve hysteresis, good reproducibility (Figure 22), and excellent environmental stability.
Hu and co-workers [8] constructed modules with an area of 10 × 10 cm2 from the pastes initially developed by the Han group [10,86]. These modules possess an active area of 49 cm2 and were infiltrated with a (5-AVA)x(MA)1-xPbI3 perovskite precursor solution in γ-butyrolactone. When the modules were composed of 10 sub-cells, the device reached a PCE up of 10.4%, VOC of 9.3 V, JSC of 2.0 mA/cm2, and FF of 0.56, with outdoor stability of 1 month and shelf-life stability of >1 year. Grancini et al. further expanded the stability of the 10 × 10 cm2 modules to >10,000 h under continuous irradiation with zero loss in performance measured under controlled standard conditions 9. The fabricated modules had an active area of 46.7 cm2 and were infiltrated with the same 3D/2D perovskite precursor solution in GVL. These modules reached an efficiency of 11.164%, VOC of 7.05 V, JSC of 2.247 mA/cm2, and FF of 0.704.
Raptis and co-workers employed copper and aluminium grids to enhance the conductivity of the carbon electrode in mini-modules of 11.7 cm2 [109]. The authors employed commercial TiO2 paste from Greatcell, ZrO2 paste from Solaronix, and carbon paste from Gwent electronic materials, to fabricate the typical triple-stack mesoporous architecture. The substrates were infiltrated with (5-AVA)x(MA)1-xPbI3 precursor solution by drop casting. Then, they prepared a low-temperature carbon ink that was used to secure the copper or aluminium grid (25 μm thickness). This carbon ink was a mixture of ethyl cellulose binder (12.5%), carbon allotropes (29.4%), 1-butanol, and terpineol. With this architecture, they obtained a PCE of 11.05%, VOC of 5.11 V, ISC of 42.73 mA, and FF of 59.18%, for copper grid reinforced mini modules. In comparison, when a device was evaluated before implementation of copper grid, a PCE of 7.70% (VOC of 5.02 V, ISC of 41.35 mA, and FF of 43.44%) was obtained.
Lee and collaborators further developed mini-modules with an active area of 9.75 cm2 under indoor conditions [158]. When an LED light of 1000 lux was used to test the device, the device achieved a Pmax of 683 μW, VOC of 3.60 V, ISC of 269 μA, and FF of 70.5%. In a recent article, the group up-scaled the mini-module area to 220 cm2 [159]. The modules were infiltrated with a green precursor solution of (5-AVA)x(MA)1-xPbI3 perovskite in GVL:methanol in a 9:1 volume ratio. Outstandingly, the module achieved a PCE up to 9.05%, VOC of 18.4 V, JSC of 21.34 mA/cm2, and FF of 51.0% in forward scan.
Bogachuk and co-workers employed commercial pastes to manufacture mini-modules of 10 × 10 cm2 geometrical area, and 56.8 cm2 active area [160]. These devices, with initial conversion efficiency of 11.1%, were tested using the “hotspot test” implemented in the IEC 61215–2:2016 norm to evaluate the stability of solar cells. These modules were tested for 1 h at 50 °C, and no visible signs of degradation were observed. Noteworthy, the performance of the devices slightly improved (OCE: 12.1%) after 1 h of hotspot test.

3.2. Other Printing Techniques

In regard to other printing techniques, fewer articles have been devoted to scaling up C-PSCs. Most of the larger area devices were created using a combination of screen printing and other methods to print or infiltrate one part of the PV architecture. Duong et al. revealed a blade coating procedure to examine the effect on the morphology and photoluminescence properties of perovskite films via a spray-assisted deposition technique using a multistep spin–spray [161]. Initially, c-TiO2 was spun coated onto the FTO substrates from titanium diisopropoxide bis(acetylacetonate) precursor solution. Subsequently, m-TiO2 layer was applied on the TiO2 BL/FTO substrate by spin coating utilizing TiO2 commercial paste diluted in methanol. Then, solutions of PbI2 in N,N-dimethylformamide (DMF) were spin-coated on the top of the m-TiO2, followed by the incorporation of MAI/IPA into the PbI2 layer using an air gun at various substrate temperatures and spray speeds. Lastly, a conductive carbon-coated copper foil was attached directly to the surface of the perovskite layer. This led to a void-free MAPI bilayer with a large grain size, which reduces moisture absorption at the grain boundaries, delays degradation of the organic/inorganic components, and boosts carrier extraction in the fabrication of PSCs. Solar devices with an area of 2.25 cm2 exhibited a maximum PCE of 10.58%.
Haruta et al. proposed a one-step blade coating method for the deposition of CsPbBr3 thin films using the mist deposition method to ensure efficient carrier transport [162]. The precursor solution was created by mixing CsPbBr3 powder with DMSO and DMF. DMF was added to reduce the solvent viscosity and thus enable mist generation. This CsPbBr3-based device exhibited remarkable PV performance (PCE of 8.3%), which is the highest for a CsPbBr3-based PSC fabricated through a one-step solution process and is comparable to those devices made by two-step solution processes.
Inkjet-printed PSCs are susceptible to “coffee ring” defects, however Chalkias et al. proposed a strategy to reduce their appearance during ambient air inkjet printing processing [137]. By changing the concentration of the perovskite precursor ink, better charge transportation and recombination kinetics were achieved, leading to improved external and internal quantum efficiency. Combining inkjet printing and screen printing, the authors developed HTL-free C-PSCs sub-modules with an average efficiency of 9.09%, covering an area of 34.2 cm2, which demonstrates the scalability of this technology. Small-area devices, using the optimized perovskite precursor ink, exhibited an efficiency of more than 12% with 0.34 cm2 of active area; this is one of the highest reported values for the inkjet-printed c-TiO2/m-TiO2/perovskite/carbon architecture.
Verma et al. fabricated C-PSCs using the slot-die technique, depositing a perovskite precursor solution on the stripes of c-TiO2, m-TiO2, m-ZrO2, and carbon in a substrate area of 100 cm2 and active area of 1.5 cm2 for each stripe [163]. To develop the ink rheology for each layer, they initially utilized commercial materials, diluting the inks for slot-die coating with a lower volume fraction of binders. For the c-TiO2 ink, they combined a low boiling point and high boiling point solvent to produce a homogenous film, whereas for the m-TiO2, m-ZrO2, and carbon inks, alcohol solvents such as i-PrOH and EtOH were used. The efficiency of the best performing cells was over 12%, comparable to the values reported for cells manufactured by screen printing. Additionally, they demonstrated the shelf-life stability of fully slot-die coated cells exceeding 1 year under ambient conditions.

4. Summary and Future Scope

Carbon-based perovskite solar cells can be fabricated from low-cost and easy printing techniques. For each technique, printing parameters must be optimized to obtain highly efficient devices. In addition to the printing parameters, ink formulation and properties play a crucial role in the quality of deposited layers. Table 1 compares the printing parameters, ink properties, and the highest reported efficiency of the C-PSC device.
Although carbon materials have been identified as an alternative to metal electrodes in order to address some of the drawbacks of metal top contacts in perovskite solar cells (PSCs), such as instability and high cost, a significant obstacle to their commercialization is the low device performance due to the poor adhesion of the carbon with the lower layers and inadequate contact with the perovskite phase, leading to significant charge recombination. This is attributed to the morphology of the perovskite layer and the inherent physical properties of carbon materials, such as crystal structure, size, and electrical conductivity. Therefore, for C-PSCs to become a mature technology for commercialization, device stability (lifetime), processing cost, and safety are essential criteria that must be taken into account. Various strategies have been proposed to improve the performance and stability of C-PSCs through interface engineering between the perovskite and carbon layers, such as by: (i) inserting an interlayer to improve the carbon work function; (ii) reducing the energy level mismatch and increasing the contact area; and (iii) modifying the perovskite precursor composition using antisolvent including carbon nanoparticles during perovskite deposition to eliminate the pores between the perovskite and carbon top electrode. Using these strategies, the fabrication of C-PSCs with excellent power conversion efficiency has been achieved. Low-temperature C-PSCs are also being investigated, as they offer advantages in terms of process cost since curing temperatures lower than the perovskite decomposition temperature are needed, thus eliminating the need for infiltrating the precursor solution through a thick stack of carbon and spacer layers such as ZrO2 or Al2O3.
Concerning process safety and non-toxicity, development of lead-free C-PSCs as well as incorporation of green solvents, such as GVL-based precursor solutions, are attractive strategies to be further investigated. Also, recently the utilization of biomass such as aloe vera plant, bamboo chopsticks, peanut shell, phragmites australis and corn stalk bio-carbons has grabbed the attention because of the simple, inexpensive synthesis routes of these materials.
Although device stability is one of the important advantages generally reported for the C-PSC technology, up to now, the research community has been focused mostly on efficiency improvement. This is necessary since this configuration is quite recent, and more time is needed to fully examine the device lifetime at longer periods of time (several years). In addition to the device performance, C-PSC manufacturing cost and toxicity must be the subject of investigations in order to increase the technological readiness level and the potential of this technology for commercialization.

Author Contributions

Funding acquisition and resources: G.O. Scientific guidance: D.P. and G.O. Writing the manuscript, graphs and schematic designs: All authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by CONACYT under the FORDECYT-PRONACES projects 318703 and 848260, and the ROYAL SOCIETY international collaboration award, ICA\R1\191321. G.O. gratefully acknowledges support from the Ministerio de Universidades and Universidad Pablo de Olavide (Spain) through the Beatriz Galindo program under project BEAGAL 18/00077 and grant BGP 18/00060.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data is available within the manuscript.

Acknowledgments

We would like to acknowledge Beatriz Heredia Cervera and Daniel Aguilar for technical assistance, and support from the “Laboratorio Nacional de Nano y Biomateriales” (LANNBIO)—CINVESTAV-Mérida.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal Halide Perovskites as Visible-Light Sensitizers for Photovoltaic Cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. [Google Scholar] [CrossRef]
  2. Lin, Q.; Armin, A.; Burn, P.L.; Meredith, P. Organohalide Perovskites for Solar Energy Conversion. Acc. Chem. Res. 2016, 49, 545–553. [Google Scholar]
  3. Stefanelli, M.; Vesce, L.; Di Carlo, A. Upscaling of Carbon-Based Perovskite Solar Module. Nanomaterials 2023, 13, 313. [Google Scholar]
  4. Keremane, K.S.; Prathapani, S.; Haur, L.J.; Bruno, A.; Priyadarshi, A.; Adhikari, A.V.; Mhaisalkar, S.G. Improving the Performance of Carbon-Based Perovskite Solar Modules (70 cm2) by Incorporating Cesium Halide in Mesoporous TiO2. ACS Appl. Energy Mater. 2021, 4, 249–258. [Google Scholar] [CrossRef]
  5. Best Research-Cell Efficiency Chart|Photovoltaic Research|NREL. Available online: https://www.nrel.gov/pv/cell-efficiency.html (accessed on 26 March 2023).
  6. R&D for Energy Transition-Fraunhofer Institute for Solar Energy Systems ISE-Fraunhofer ISE. Available online: https://www.ise.fraunhofer.de/en.html (accessed on 21 February 2023).
  7. Wang, R.; Mujahid, M.; Duan, Y.; Wang, Z.; Xue, J.; Yang, Y. A Review of Perovskites Solar Cell Stability. Adv. Funct. Mater. 2019, 29, 1808843. [Google Scholar] [CrossRef]
  8. Lee, J.-W.; Meng, L.; Yang, Y. Semiconducting Metal Oxides for High Performance Perovskite Solar Cells. In The Future of Semiconductor Oxides in Next-Generation Solar Cells; Elsevier: Amsterdam, The Netherlands, 2018; pp. 241–265. [Google Scholar]
  9. Grancini, G.; Roldán-Carmona, C.; Zimmermann, I.; Mosconi, E.; Lee, X.; Martineau, D.; Narbey, S.; Oswald, F.; De Angelis, F.; Graetzel, M.; et al. One-Year Stable Perovskite Solar Cells by 2D/3D Interface Engineering. Nat. Commun. 2017, 8, 15684. [Google Scholar] [CrossRef]
  10. Hu, Y.; Si, S.; Mei, A.; Rong, Y.; Liu, H.; Li, X.; Han, H. Stable Large-Area (10  ×  10 cm2) Printable Mesoscopic Perovskite Module Exceeding 10% Efficiency. Solar RRL 2017, 1, 1600019. [Google Scholar] [CrossRef]
  11. Chen, H.; Yang, S. Methods and Strategies for Achieving High-Performance Carbon-Based Perovskite Solar Cells without Hole Transport Materials. J. Mater. Chem. A Mater. 2019, 7, 15476–15490. [Google Scholar] [CrossRef]
  12. Hu, H.; Birkhold, S.; Sultan, M.; Fakharuddin, A.; Koch, S.; Schmidt-Mende, L. Surface Band Bending Influences the Open-Circuit Voltage of Perovskite Solar Cells. ACS Appl. Energy Mater. 2019, 2, 4045–4052. [Google Scholar] [CrossRef]
  13. Caprioglio, P.; Stolterfoht, M.; Wolff, C.M.; Unold, T.; Rech, B.; Albrecht, S.; Neher, D. On the Relation between the Open-Circuit Voltage and Quasi-Fermi Level Splitting in Efficient Perovskite Solar Cells. Adv. Energy Mater. 2019, 9, 1901631. [Google Scholar] [CrossRef]
  14. Greatcell Solar Materials. Available online: https://www.greatcellsolarmaterials.com/ (accessed on 13 February 2023).
  15. Solaronix—Innovative Solutions for Solar Professionals. Available online: https://www.solaronix.com/ (accessed on 13 February 2023).
  16. Wonder Solar Powering the World. Available online: http://www.wondersolar.cn/ (accessed on 13 February 2023).
  17. Zhang, L.; Liu, T.; Liu, L.; Hu, M.; Yang, Y.; Mei, A.; Han, H. The Effect of Carbon Counter Electrodes on Fully Printable Mesoscopic Perovskite Solar Cells. J. Mater. Chem. A Mater. 2015, 3, 9165–9170. [Google Scholar] [CrossRef]
  18. Que, M.; Zhang, B.; Chen, J.; Yin, X.; Yun, S. Carbon-Based Electrodes for Perovskite Solar Cells. Mater. Adv. 2021, 2, 5560–5579. [Google Scholar] [CrossRef]
  19. Wei, Z.; Chen, H.; Yan, K.; Zheng, X.; Yang, S. Hysteresis-Free Multi-Walled Carbon Nanotube-Based Perovskite Solar Cells with a High Fill Factor. J. Mater. Chem. A Mater. 2015, 3, 24226–24231. [Google Scholar] [CrossRef]
  20. Wang, Y.; Zhao, H.; Mei, Y.; Liu, H.; Wang, S.; Li, X. Carbon Nanotube Bridging Method for Hole Transport Layer-Free Paintable Carbon-Based Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2019, 11, 916–923. [Google Scholar] [CrossRef]
  21. Zhou, J.; Wu, J.; Li, N.; Li, X.; Zheng, Y.-Z.; Tao, X. Efficient All-Air Processed Mixed Cation Carbon-Based Perovskite Solar Cells with Ultra-High Stability. J. Mater. Chem. A Mater. 2019, 7, 17594–17603. [Google Scholar] [CrossRef]
  22. Zhang, C.; Wang, S.; Zhang, H.; Feng, Y.; Tian, W.; Yan, Y.; Bian, J.; Wang, Y.; Jin, S.; Zakeeruddin, S.M.; et al. Efficient Stable Graphene-Based Perovskite Solar Cells with High Flexibility in Device Assembling via Modular Architecture Design. Energy Environ. Sci. 2019, 12, 3585–3594. [Google Scholar] [CrossRef]
  23. Liu, J.; Lei, M.; Wang, G. Communication—Enhancing Hole Extraction in Carbon-Based CsPbBr3 Inorganic Perovskite Solar Cells without Hole-Transport Layer. ECS J. Solid State Sci. Technol. 2020, 9, 041004. [Google Scholar] [CrossRef]
  24. Bhandari, S.; Roy, A.; Ali, M.S.; Mallick, T.K.; Sundaram, S. Cotton Soot Derived Carbon Nanoparticles for NiO Supported Processing Temperature Tuned Ambient Perovskite Solar Cells. Sci. Rep. 2021, 11, 23388. [Google Scholar] [CrossRef]
  25. Pitchaiya, S.; Eswaramoorthy, N.; Natarajan, M.; Santhanam, A.; Asokan, V.; Madurai Ramakrishnan, V.; Rangasamy, B.; Sundaram, S.; Ravirajan, P.; Velauthapillai, D. Perovskite Solar Cells: A Porous Graphitic Carbon Based Hole Transporter/Counter Electrode Material Extracted from an Invasive Plant Species Eichhornia Crassipes. Sci. Rep. 2020, 10, 6835. [Google Scholar] [CrossRef]
  26. Geng, C.; Xie, Y.; Wei, P.; Liu, H.; Qiang, Y.; Zhang, Y. An Efficient Co-NC Composite Additive for Enhancing Interface Performance of Carbon-Based Perovskite Solar Cells. Electrochim. Acta 2020, 358, 136883. [Google Scholar] [CrossRef]
  27. Guo, M.; Wei, C.; Liu, C.; Zhang, K.; Su, H.; Xie, K.; Zhai, P.; Zhang, J.; Liu, L. Composite Electrode Based on Single-Atom Ni Doped Graphene for Planar Carbon-Based Perovskite Solar Cells. Mater. Des. 2021, 209, 109972. [Google Scholar] [CrossRef]
  28. Xie, Y.; Cheng, J.; Liu, H.; Liu, J.; Maitituersun, B.; Ma, J.; Qiang, Y.; Shi, H.; Geng, C.; Li, Y.; et al. Co-Ni Alloy@carbon Aerogels for Improving the Efficiency and Air Stability of Perovskite Solar Cells and Its Hysteresis Mechanism. Carbon 2019, 154, 322–329. [Google Scholar] [CrossRef]
  29. Chen, L.; Duan, Q.; Dong, W.; Zhu, A.; Zhang, A.; Zhang, X.; Zhong, J.; Huang, F.; Cheng, Y.; Xiao, J. Sprayed and Mechanical-Modified Graphite Layer as Transferred Electrode for High-Efficiency Perovskite Solar Cells. Carbon 2023, 202, 161–166. [Google Scholar] [CrossRef]
  30. Chu, L.; Liu, W.; Qin, Z.; Zhang, R.; Hu, R.; Yang, J.; Yang, J.; Li, X. Boosting Efficiency of Hole Conductor-Free Perovskite Solar Cells by Incorporating p-Type NiO Nanoparticles into Carbon Electrodes. Sol. Energy Mater. Sol. Cells 2018, 178, 164–169. [Google Scholar] [CrossRef]
  31. Taylor, J.F. Spin Coating: An Overview. Met. Finish. 2001, 99, 16–21. [Google Scholar] [CrossRef]
  32. Sahu, N.; Parija, B.; Panigrahi, S. Fundamental Understanding and Modeling of Spin Coating Process: A Review. Indian J. Phys. 2009, 83, 493–502. [Google Scholar] [CrossRef]
  33. Scriven, L.E. Physics and Applications of DIP Coating and Spin Coating. MRS Proc. 1988, 121, 717. [Google Scholar] [CrossRef]
  34. Yang, Y.-B.; Chen, P.; Li, H.-S.; Zhao, Q.; Li, T.-T.; Wu, Y.; Zhang, Y.; Gao, X.-P.; Li, G.-R. Reversible Degradation in Hole Transport Layer-Free Carbon-Based Perovskite Solar Cells. Solar RRL 2022, 6, 2200281. [Google Scholar] [CrossRef]
  35. Dileep, R.; Kesavan, G.; Reddy, V.; Rajbhar, M.K.; Shanmugasundaram, S.; Ramasamy, E.; Veerappan, G. Room-Temperature Curable Carbon Cathode for Hole-Conductor Free Perovskite Solar Cells. Solar Energy 2019, 187, 261–268. [Google Scholar] [CrossRef]
  36. Li, J.; Rossignol, F.; Macdonald, J. Inkjet Printing for Biosensor Fabrication: Combining Chemistry and Technology for Advanced Manufacturing. Lab Chip 2015, 15, 2538–2558. [Google Scholar] [CrossRef]
  37. Singh, M.; Haverinen, H.M.; Dhagat, P.; Jabbour, G.E. Inkjet Printing-Process and Its Applications. Adv. Mater. 2010, 22, 673–685. [Google Scholar] [CrossRef]
  38. Yang, Y.; Hoang, M.T.; Yao, D.; Pham, N.D.; Tiong, V.T.; Wang, X.; Sun, W.; Wang, H. High Performance Carbon-Based Planar Perovskite Solar Cells by Hot-Pressing Approach. Sol. Energy Mater. Sol. Cells 2020, 210, 110517. [Google Scholar] [CrossRef]
  39. Jin, J.; Yang, M.; Deng, W.; Xin, J.; Tai, Q.; Qian, J.; Dong, B.; Li, W.; Wang, J.; Li, J. Highly Efficient and Stable Carbon-Based Perovskite Solar Cells with the Polymer Hole Transport Layer. Solar Energy 2021, 220, 491–497. [Google Scholar] [CrossRef]
  40. Yin, R.; Wang, K.; Huo, X.; Sun, Y.; Sun, W.; Gao, Y.; You, T.; Yin, P. A One-Step Ionic Liquid Interface-to-Bulk Modification for Stable Carbon-Based CsPbI3 Perovskite Solar Cells with Efficiency Over 15%. Adv. Mater. Interfaces 2022, 9, 2201488. [Google Scholar] [CrossRef]
  41. Zamanpour, F.; Behrouznejad, F.; Ghavaminia, E.; Khosroshahi, R.; Zhan, Y.; Taghavinia, N. Fast Light-Cured TiO2 Layers for Low-Cost Carbon-Based Perovskite Solar Cells. ACS Appl. Energy Mater. 2021, 4, 7800–7810. [Google Scholar] [CrossRef]
  42. Yin, R.; Wang, K.-X.; Cui, S.; Fan, B.-B.; Liu, J.-W.; Gao, Y.-K.; You, T.-T.; Yin, P.-G. Dual-Interface Modification with BMIMPF6 for High-Efficiency and Stable Carbon-Based CsPbI2Br Perovskite Solar Cells. ACS Appl. Energy Mater. 2021, 4, 9294–9303. [Google Scholar] [CrossRef]
  43. Zhang, H.; Xiao, Y.; Qi, F.; Liu, P.; Wang, Y.; Li, F.; Wang, C.; Fang, G.; Zhao, X. Near-Infrared Light-Sensitive Hole-Transport-Layer Free Perovskite Solar Cells and Photodetectors with Hexagonal NaYF4:Yb3+, Tm3+ @SiO2 Upconversion Nanoprism-Modified TiO2 Scaffold. ACS Sustain. Chem. Eng. 2019, 7, 8236–8244. [Google Scholar] [CrossRef]
  44. Fan, W.; Wei, Z.; Zhang, Z.; Qiu, F.; Hu, C.; Li, Z.; Xu, M.; Qi, J. High-Performance Carbon-Based Perovskite Solar Cells through the Dual Role of PC61BM. Inorg. Chem. Front. 2019, 6, 2767–2775. [Google Scholar] [CrossRef]
  45. Zhu, J.; Tang, M.; He, B.; Shen, K.; Zhang, W.; Sun, X.; Sun, M.; Chen, H.; Duan, Y.; Tang, Q. Ultraviolet Filtration and Defect Passivation for Efficient and Photostable CsPbBr3 Perovskite Solar Cells by Interface Engineering with Ultraviolet Absorber. Chem. Eng. J. 2021, 404, 126548. [Google Scholar] [CrossRef]
  46. Wang, C.; Long, Y.; Liu, X.; Fu, S.; Wang, J.; Zhang, J.; Hu, Z.; Zhu, Y. A Dual Promotion Strategy of Interface Modification and Ion Doping for Efficient and Stable Carbon-Based Planar CsPbBr3 Perovskite Solar Cells. J. Mater. Chem. C Mater. 2020, 8, 17211–17221. [Google Scholar] [CrossRef]
  47. Zhu, W.; Zhang, Z.; Chai, W.; Zhang, Q.; Chen, D.; Lin, Z.; Chang, J.; Zhang, J.; Zhang, C.; Hao, Y. Band Alignment Engineering Towards High Efficiency Carbon-Based Inorganic Planar CsPbIBr2 Perovskite Solar Cells. ChemSusChem 2019, 12, 2318–2325. [Google Scholar] [CrossRef]
  48. Zhu, W.; Zhang, Z.; Chen, D.; Chai, W.; Chen, D.; Zhang, J.; Zhang, C.; Hao, Y. Interfacial Voids Trigger Carbon-Based, All-Inorganic CsPbIBr2 Perovskite Solar Cells with Photovoltage Exceeding 1.33 V. Nanomicro Lett. 2020, 12, 87. [Google Scholar] [CrossRef]
  49. Zhu, W.; Chai, W.; Deng, M.; Chen, D.; Chen, D.; Zhang, J.; Zhang, C.; Hao, Y. Flux-Mediated Growth Strategy Enables Low-Temperature Fabrication of High-Efficiency All-Inorganic CsPbIBr2 Perovskite Solar Cells. Electrochim. Acta 2020, 330, 135325. [Google Scholar] [CrossRef]
  50. Chalkias, D.A.; Karavioti, A.; Papanicolaou, G.C.; Stathatos, E. Stability Assessment of Carbon-Based Hole-Transport-Layer-Free Perovskite Solar Cells under Accelerated Ageing: A Combined Experimental and Predictive Modelling Analysis. Electrochim. Acta 2022, 427, 140905. [Google Scholar] [CrossRef]
  51. Zhang, Z.; Qiu, F.; Shen, T.; Xu, L.; Ma, J.; Qi, J. Co/Eu Co-doped Electron Transport Layer Enhances Charge Extraction and Light Absorption for Efficient Carbon-based HTM-free Perovskite Solar Cells. Int. J. Energy Res. 2021, 45, 5224–5234. [Google Scholar] [CrossRef]
  52. Liu, D.; Li, S.; Zhang, P.; Wang, Y.; Zhang, R.; Sarvari, H.; Wang, F.; Wu, J.; Wang, Z.; Chen, Z.D. Efficient Planar Heterojunction Perovskite Solar Cells with Li-Doped Compact TiO2 Layer. Nano Energy 2017, 31, 462–468. [Google Scholar] [CrossRef]
  53. Liu, X.; Li, J.; Liu, Z.; Tan, X.; Sun, B.; Xi, S.; Shi, T.; Tang, Z.; Liao, G. Vapor-Assisted Deposition of CsPbIBr2 Films for Highly Efficient and Stable Carbon-Based Planar Perovskite Solar Cells with Superior Voc. Electrochim. Acta 2020, 330, 135266. [Google Scholar] [CrossRef]
  54. Yang, Y.; Liu, Z.; Ng, W.K.; Zhang, L.; Zhang, H.; Meng, X.; Bai, Y.; Xiao, S.; Zhang, T.; Hu, C.; et al. An Ultrathin Ferroelectric Perovskite Oxide Layer for High-Performance Hole Transport Material Free Carbon Based Halide Perovskite Solar Cells. Adv. Funct. Mater. 2019, 29, 1806506. [Google Scholar] [CrossRef]
  55. Mashreghi, A.; Maleki, K.; Moradzadeh, M. Two-Phase Synthesized Cu2ZnSnS4 Nanoparticles as Inorganic Hole-Transporting Material of Paintable Carbon-Based Perovskite Solar Cells. Solar Energy 2020, 201, 547–554. [Google Scholar] [CrossRef]
  56. Fu, X.; Zhou, K.; Zhou, X.; Ji, H.; Min, Y.; Qian, Y. Surface Passivation for Enhancing Photovoltaic Performance of Carbon-Based CsPbI3 Perovskite Solar Cells. J. Solid State Chem. 2022, 308, 122891. [Google Scholar] [CrossRef]
  57. Wang, G.; Liu, J.; Lei, M.; Zhang, W.; Zhu, G. Optimizing the Substrate Pre-Heating and Post-Annealing Temperatures for Fabricating High-Performance Carbon-Based CsPbIBr2 Inorganic Perovskite Solar Cells. Electrochim. Acta 2020, 349, 136354. [Google Scholar] [CrossRef]
  58. Arora, N.; Dar, M.I.; Akin, S.; Uchida, R.; Baumeler, T.; Liu, Y.; Zakeeruddin, S.M.; Grätzel, M. Low-Cost and Highly Efficient Carbon-Based Perovskite Solar Cells Exhibiting Excellent Long-Term Operational and UV Stability. Small 2019, 15, 1904746. [Google Scholar] [CrossRef]
  59. Acchutharaman, K.R.; Santhosh, N.; Isaac Daniel, R.; Senthil Pandian, M.; Ramasamy, P. Enhanced Electron Harvesting in next Generation Solar Cells by Employing TiO2 Nanoparticles Prepared through Hydrolysis Catalytic Process. Ceram. Int. 2021, 47, 21263–21270. [Google Scholar] [CrossRef]
  60. Lin, L.; Jones, T.W.; Yang, T.C.; Duffy, N.W.; Li, J.; Zhao, L.; Chi, B.; Wang, X.; Wilson, G.J. Inorganic Electron Transport Materials in Perovskite Solar Cells. Adv. Funct. Mater. 2021, 31, 2008300. [Google Scholar] [CrossRef]
  61. Bai, Y.; Mora-Seró, I.; De Angelis, F.; Bisquert, J.; Wang, P. Titanium Dioxide Nanomaterials for Photovoltaic Applications. Chem. Rev. 2014, 114, 10095–10130. [Google Scholar] [CrossRef]
  62. Xiao, Y.; Wang, C.; Kondamareddy, K.K.; Liu, P.; Qi, F.; Zhang, H.; Guo, S.; Zhao, X.-Z. Enhancing the Performance of Hole-Conductor Free Carbon-Based Perovskite Solar Cells through Rutile-Phase Passivation of Anatase TiO2 Scaffold. J. Power Sources 2019, 422, 138–144. [Google Scholar] [CrossRef]
  63. Bhandari, S.; Roy, A.; Mallick, T.K.; Sundaram, S. Morphology Modulated Brookite TiO2 and BaSnO3 as Alternative Electron Transport Materials for Enhanced Performance of Carbon Perovskite Solar Cells. Chem. Eng. J. 2022, 446, 137378. [Google Scholar] [CrossRef]
  64. Balis, N.; Zaky, A.A.; Perganti, D.; Kaltzoglou, A.; Sygellou, L.; Katsaros, F.; Stergiopoulos, T.; Kontos, A.G.; Falaras, P. Dye Sensitization of Titania Compact Layer for Efficient and Stable Perovskite Solar Cells. ACS Appl. Energy Mater. 2018, 1, 6161–6171. [Google Scholar] [CrossRef]
  65. Deng, F.; Li, X.; Lv, X.; Zhou, J.; Chen, Y.; Sun, X.; Zheng, Y.-Z.; Tao, X.; Chen, J.-F. Low-Temperature Processing All-Inorganic Carbon-Based Perovskite Solar Cells up to 11.78% Efficiency via Alkali Hydroxides Interfacial Engineering. ACS Appl. Energy Mater. 2020, 3, 401–410. [Google Scholar] [CrossRef]
  66. Han, S.; Zhang, H.; Li, Y.; Wang, R.; He, Q. Solution-Processed Amino Acid Modified SnO2 Electron Transport Layer for Carbon-Based CsPbIBr2 Perovskite Solar Cells. Mater. Sci. Semicond. Process. 2021, 133, 105964. [Google Scholar] [CrossRef]
  67. He, Q.; Zhang, H.; Han, S.; Xing, Y.; Li, Y.; Zhang, X.; Wang, R. Improvement of Nanopore Structure SnO2 Electron-Transport Layer for Carbon-Based CsPbIBr2 Perovskite Solar Cells. Mater. Sci. Semicond. Process. 2022, 148, 106787. [Google Scholar] [CrossRef]
  68. Qi, X.; Wang, J.; Tan, F.; Dong, C.; Liu, K.; Li, X.; Zhang, L.; Wu, H.; Wang, H.-L.; Qu, S.; et al. Quantum Dot Interface-Mediated CsPbIBr2 Film Growth and Passivation for Efficient Carbon-Based Solar Cells. ACS Appl. Mater. Interfaces 2021, 13, 55349–55357. [Google Scholar] [CrossRef]
  69. He, Q.; Zhang, H.; Han, S.; Wang, R.; Li, Y.; Zhang, X.; Xing, Y. Improvement of Thiourea (Lewis Base)-Modified SnO2 Electron-Transport Layer for Carbon-Based CsPbIBr2 Perovskite Solar Cells. ACS Appl. Energy Mater. 2021, 4, 10958–10967. [Google Scholar] [CrossRef]
  70. Xing, Y.; Zhang, H.; Yan, Z.; Guo, Y.; Qin, W.; Feng, Y.; Liu, S.; He, Q.; Han, S.; Zhu, H. LiF-Modified SnO2 Electron Transport Layer Improves the Performance of Carbon-Based All-Inorganic CsPbIBr2 Perovskite Solar Cells. Energy Fuels 2022, 36, 13179–13186. [Google Scholar] [CrossRef]
  71. Guo, Z.; Teo, S.; Xu, Z.; Zhang, C.; Kamata, Y.; Hayase, S.; Ma, T. Achievable High Voc of Carbon Based All-Inorganic CsPbIBr2 Perovskite Solar Cells through Interface Engineering. J. Mater. Chem. A Mater. 2019, 7, 1227–1232. [Google Scholar] [CrossRef]
  72. Qiang, Y.; Xie, Y.; Qi, Y.; Wei, P.; Shi, H.; Geng, C.; Liu, H. Enhanced Performance of Carbon-Based Perovskite Solar Cells with a Li+-Doped SnO2 Electron Transport Layer and Al2O3 Scaffold Layer. Solar Energy 2020, 201, 523–529. [Google Scholar] [CrossRef]
  73. Xie, L.; Zhang, L.; Hua, Y.; Ding, L. Inorganic Electron-Transport Materials in Perovskite Solar Cells. J. Semicond. 2022, 43, 040201. [Google Scholar] [CrossRef]
  74. Zhao, F.; Guo, Y.; Wang, X.; Tao, J.; Li, Z.; Zheng, D.; Jiang, J.; Hu, Z.; Chu, J. Efficient Carbon-Based Planar CsPbBr3 Perovskite Solar Cells with Li-Doped Amorphous Nb2O5 Layer. J. Alloys Compd. 2020, 842, 155984. [Google Scholar] [CrossRef]
  75. Wang, C.; Zhang, J.; Jiang, L.; Gong, L.; Xie, H.; Gao, Y.; He, H.; Fang, Z.; Fan, J.; Chao, Z. All-Inorganic, Hole-Transporting-Layer-Free, Carbon-Based CsPbIBr2 Planar Solar Cells with ZnO as Electron-Transporting Materials. J. Alloys Compd. 2020, 817, 152768. [Google Scholar] [CrossRef]
  76. Zhang, J.; Wang, C.; Fu, H.; Gong, L.; He, H.; Fang, Z.; Zhou, C.; Chen, J.; Chao, Z.; Fan, J. Low-Temperature Preparation Achieving 10.95%-Efficiency of Hole-Free and Carbon-Based All-Inorganic CsPbI3 Perovskite Solar Cells. J. Alloys Compd. 2021, 862, 158454. [Google Scholar] [CrossRef]
  77. Fu, H.; Zhang, J.; Li, Y.; Gong, L.; He, H.; Fang, Z.; Zhou, C.; Chen, J.; Fan, J. A Facile Interface Engineering Method to Improve the Performance of FTO/ZnO/CsPbI3−xBrx (x<1)/C Solar Cells. J. Mater. Sci. Mater. Electron. 2022, 33, 3711–3725. [Google Scholar] [CrossRef]
  78. Zheng, X.; Chen, H.; Li, Q.; Yang, Y.; Wei, Z.; Bai, Y.; Qiu, Y.; Zhou, D.; Wong, K.S.; Yang, S. Boron Doping of Multiwalled Carbon Nanotubes Significantly Enhances Hole Extraction in Carbon-Based Perovskite Solar Cells. Nano Lett. 2017, 17, 2496–2505. [Google Scholar] [CrossRef]
  79. Gao, L.; Hu, J.; Meng, F.; Zhou, Y.; Li, Y.; Wei, G.; Ma, T. Comparison of Interfacial Bridging Carbon Materials for Effective Carbon-Based Perovskite Solar Cells. J. Colloid Interface Sci. 2020, 579, 425–430. [Google Scholar] [CrossRef]
  80. Li, S.; Li, Y.; Liu, K.; Chen, M.; Peng, W.; Yang, Y.; Li, X. Laser Fabricated Carbon Quantum Dots in Anti-Solvent for Highly Efficient Carbon-Based Perovskite Solar Cells. J. Colloid Interface Sci. 2021, 600, 691–700. [Google Scholar] [CrossRef]
  81. Xu, T.; Wan, Z.; Tang, H.; Zhao, C.; Lv, S.; Chen, Y.; Chen, L.; Qiao, Q.; Huang, W. Carbon Quantum Dot Additive Engineering for Efficient and Stable Carbon-Based Perovskite Solar Cells. J. Alloys Compd. 2021, 859, 157784. [Google Scholar] [CrossRef]
  82. Novaković, D.; Kašiković, N.; Vladić, G.; Pál, M. Screen Printing, Printing on Polymers: Fundamentals and Applications. In Printing on Polymers; Elsevier: Amsterdam, The Netherlands, 2016; pp. 247–261. [Google Scholar]
  83. Licari, J.J.; Enlow, L.R. Thick Film Processes. In Hybrid Microcircuit Technology Handbook; Elsevier: Amsterdam, The Netherlands, 1998; pp. 104–171. [Google Scholar]
  84. Singh, S.; Wang, J.; Cinti, S. Review—An Overview on Recent Progress in Screen-Printed Electroanalytical (Bio)Sensors. ECS Sens. Plus 2022, 1, 023401. [Google Scholar] [CrossRef]
  85. Kapur, N.; Abbott, S.J.; Dolden, E.D.; Gaskell, P.H. Predicting the Behavior of Screen Printing. IEEE Trans. Compon. Packag. Manuf. Technol. 2013, 3, 508–515. [Google Scholar] [CrossRef]
  86. Ku, Z.; Rong, Y.; Xu, M.; Liu, T.; Han, H. Full Printable Processed Mesoscopic CH3NH3PbI3/TiO2 Heterojunction Solar Cells with Carbon Counter Electrode. Sci. Rep. 2013, 3, 3132. [Google Scholar] [CrossRef]
  87. Hvojnik, M.; Popovičová, A.M.; Pavličková, M.; Hatala, M.; Gemeiner, P.; Müllerová, J.; Tomanová, K.; Mikula, M. Solution-Processed TiO2 Blocking Layers in Printed Carbon-Based Perovskite Solar Cells. Appl. Surf. Sci. 2021, 536, 147888. [Google Scholar] [CrossRef]
  88. Wang, D.; Kang, J.; Zhang, Z.; Zhu, G.; Zhang, X.; Xiong, J.; Zhang, J. Efficient Printable Carbon-Based Mesoscopic Perovskite Solar Cells Based on Aluminum and Indium Doped TiO2 Compact Layer. Mater. Lett. 2022, 322, 132427. [Google Scholar] [CrossRef]
  89. Liu, Z.; Bi, S.; Hou, G.; Ying, C.; Su, X. Dual-Sized TiO2 Nanoparticles as Scaffold Layers in Carbon-Based Mesoscopic Perovskite Solar Cells with Enhanced Performance. J. Power Sources 2019, 430, 12–19. [Google Scholar] [CrossRef]
  90. Zhao, X.; Zhao, J.; He, J.; Li, B.; Zhang, Y.; Hu, J.; Wang, H.; Zhang, D.; Liu, Q. Porous Anatase TiO2 Nanocrystal Derived from the Metal–Organic Framework as Electron Transport Material for Carbon-Based Perovskite Solar Cells. ACS Appl. Energy Mater. 2020, 3, 6180–6187. [Google Scholar] [CrossRef]
  91. Li, B.; Zhao, J.; Lu, Q.; Zhou, S.; Wei, H.; Lv, T.; Zhang, Y.; Zhang, J.; Liu, Q. Carbon-Based Printable Perovskite Solar Cells with a Mesoporous TiO2 Electron Transporting Layer Derived from Metal–Organic Framework NH2-MIL-125. Energy Technol. 2021, 9, 2000957. [Google Scholar] [CrossRef]
  92. Liu, J.; Guan, Y.; Liu, S.; Li, S.; Gao, C.; Du, J.; Qiu, C.; Li, D.; Zhang, D.; Wang, X.; et al. Modulating Oxygen Vacancies in BaSnO3 for Printable Carbon-Based Mesoscopic Perovskite Solar Cells. ACS Appl. Energy Mater. 2021, 4, 11032–11040. [Google Scholar] [CrossRef]
  93. Liu, T.; Liu, L.; Hu, M.; Yang, Y.; Zhang, L.; Mei, A.; Han, H. Critical Parameters in TiO2/ZrO2/Carbon-Based Mesoscopic Perovskite Solar Cell. J. Power Sources 2015, 293, 533–538. [Google Scholar] [CrossRef]
  94. Meng, Z.; Guo, D.; Yu, J.; Fan, K. Investigation of Al2O3 and ZrO2 Spacer Layers for Fully Printable and Hole-Conductor-Free Mesoscopic Perovskite Solar Cells. Appl. Surf. Sci. 2018, 430, 632–638. [Google Scholar] [CrossRef]
  95. Rong, Y.; Han, H. Monolithic Quasi-Solid-State Dye-Sensitized Solar Cells Based on Graphene-Modified Mesoscopic Carbon-Counter Electrodes. J. Nanophotonics 2013, 7, 073090. [Google Scholar] [CrossRef]
  96. Xiong, Y.; Zhu, X.; Mei, A.; Qin, F.; Liu, S.; Zhang, S.; Jiang, Y.; Zhou, Y.; Han, H. Bifunctional Al2O3 Interlayer Leads to Enhanced Open-Circuit Voltage for Hole-Conductor-Free Carbon-Based Perovskite Solar Cells. Solar RRL 2018, 2, 1800002. [Google Scholar] [CrossRef]
  97. Bashir, A.; Shukla, S.; Lew, J.H.; Shukla, S.; Bruno, A.; Gupta, D.; Baikie, T.; Patidar, R.; Akhter, Z.; Priyadarshi, A.; et al. Spinel Co3O4 Nanomaterials for Efficient and Stable Large Area Carbon-Based Printed Perovskite Solar Cells. Nanoscale 2018, 10, 2341–2350. [Google Scholar] [CrossRef]
  98. Behrouznejad, F.; Tsai, C.-M.; Narra, S.; Diau, E.W.-G.; Taghavinia, N. Interfacial Investigation on Printable Carbon-Based Mesoscopic Perovskite Solar Cells with NiOx/C Back Electrode. ACS Appl. Mater. Interfaces 2017, 9, 25204–25215. [Google Scholar] [CrossRef]
  99. Tao, L.; Zhang, Y.; Chen, H.; Wang, K.; Zhou, X. Printable Commercial Carbon Based Mesoscopic Perovskite Solar Cell Using NiO/Graphene as Hole-Transport Materials. ECS J. Solid State Sci. Technol. 2021, 10, 105003. [Google Scholar] [CrossRef]
  100. Icli, K.C.; Ozenbas, M. Modification of TiO2 and NiO Charge Selective Mesoporous Layers Using Excessive Y and Li Additions for Carbon Based Perovskite Solar Cells. J. Power Sources 2021, 506, 230229. [Google Scholar] [CrossRef]
  101. Bashir, A.; Lew, J.H.; Shukla, S.; Gupta, D.; Baikie, T.; Chakraborty, S.; Patidar, R.; Bruno, A.; Mhaisalkar, S.; Akhter, Z. Cu-Doped Nickel Oxide Interface Layer with Nanoscale Thickness for Efficient and Highly Stable Printable Carbon-Based Perovskite Solar Cell. Solar Energy 2019, 182, 225–236. [Google Scholar] [CrossRef]
  102. Duan, M.; Tian, C.; Hu, Y.; Mei, A.; Rong, Y.; Xiong, Y.; Xu, M.; Sheng, Y.; Jiang, P.; Hou, X.; et al. Boron-Doped Graphite for High Work Function Carbon Electrode in Printable Hole-Conductor-Free Mesoscopic Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2017, 9, 31721–31727. [Google Scholar] [CrossRef]
  103. Chen, X.; Lu, L.; Gu, D.; Zhang, X.; Yu, H.; Chen, F.; Rui, Y.; Hou, J.; Yang, Y. Chlorine Management of a Carbon Counter Electrode for High Performance Printable Perovskite Solar Cells. J. Mater. Chem. C Mater. 2021, 9, 8615–8622. [Google Scholar] [CrossRef]
  104. Chu, Q.-Q.; Sun, Z.; Ding, B.; Moon, K.; Yang, G.-J.; Wong, C.-P. Greatly Enhanced Power Conversion Efficiency of Hole-Transport-Layer-Free Perovskite Solar Cell via Coherent Interfaces of Perovskite and Carbon Layers. Nano Energy 2020, 77, 105110. [Google Scholar] [CrossRef]
  105. Jiang, P.; Xiong, Y.; Xu, M.; Mei, A.; Sheng, Y.; Hong, L.; Jones, T.W.; Wilson, G.J.; Xiong, S.; Li, D.; et al. The Influence of the Work Function of Hybrid Carbon Electrodes on Printable Mesoscopic Perovskite Solar Cells. J. Phys. Chem. C 2018, 122, 16481–16487. [Google Scholar] [CrossRef]
  106. Gong, S.; Li, H.; Chen, Z.; Shou, C.; Huang, M.; Yang, S. CsPbI2Br Perovskite Solar Cells Based on Carbon Black-Containing Counter Electrodes. ACS Appl. Mater. Interfaces 2020, 12, 34882–34889. [Google Scholar] [CrossRef]
  107. Wang, J.; Gong, S.; Chen, Z.; Yang, S. Vacuum-Assisted Drying Process for Screen-Printable Carbon Electrodes of Perovskite Solar Cells with Enhanced Performance Based on Cuprous Thiocyanate as a Hole Transporting Layer. ACS Appl. Mater. Interfaces 2021, 13, 22684–22693. [Google Scholar] [CrossRef]
  108. Hatala, M.; Gemeiner, P.; Hvojnik, M.; Mikula, M. The Effect of the Ink Composition on the Performance of Carbon-Based Conductive Screen Printing Inks. J. Mater. Sci. Mater. Electron. 2019, 30, 1034–1044. [Google Scholar] [CrossRef]
  109. Raptis, D.; Stoichkov, V.; Meroni, S.M.P.; Pockett, A.; Worsley, C.A.; Carnie, M.; Worsley, D.A.; Watson, T. Enhancing Fully Printable Mesoscopic Perovskite Solar Cell Performance Using Integrated Metallic Grids to Improve Carbon Electrode Conductivity. Curr. Appl. Phys. 2020, 20, 619–627. [Google Scholar] [CrossRef]
  110. Zhang, Y.; Zhao, J.; Zhang, J.; Jiang, X.; Zhu, Z.; Liu, Q. Interface Engineering Based on Liquid Metal for Compact-Layer-Free, Fully Printable Mesoscopic Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2018, 10, 15616–15623. [Google Scholar] [CrossRef]
  111. Kajal, P.; Lew, J.H.; Kanwat, A.; Rana, P.J.S.; Nutan, G.V.; Koh, T.M.; Mhaisalkar, S.G.; Powar, S.; Mathews, N. Unveiling the Role of Carbon Black in Printable Mesoscopic Perovskite Solar Cells. J. Power Sources 2021, 501, 230019. [Google Scholar] [CrossRef]
  112. Bogachuk, D.; Tsuji, R.; Martineau, D.; Narbey, S.; Herterich, J.P.; Wagner, L.; Suginuma, K.; Ito, S.; Hinsch, A. Comparison of Highly Conductive Natural and Synthetic Graphites for Electrodes in Perovskite Solar Cells. Carbon 2021, 178, 10–18. [Google Scholar] [CrossRef]
  113. Tsuji, R.; Bogachuk, D.; Martineau, D.; Wagner, L.; Kobayashi, E.; Funayama, R.; Matsuo, Y.; Mastroianni, S.; Hinsch, A.; Ito, S. Function of Porous Carbon Electrode during the Fabrication of Multiporous-Layered-Electrode Perovskite Solar Cells. Photonics 2020, 7, 133. [Google Scholar] [CrossRef]
  114. Phillips, C.; Al-Ahmadi, A.; Potts, S.-J.; Claypole, T.; Deganello, D. The Effect of Graphite and Carbon Black Ratios on Conductive Ink Performance. J. Mater. Sci. 2017, 52, 9520–9530. [Google Scholar] [CrossRef]
  115. Potts, S.-J.; Lau, Y.C.; Dunlop, T.; Claypole, T.; Phillips, C. Effect of Photonic Flash Annealing with Subsequent Compression Rolling on the Topography, Microstructure and Electrical Performance of Carbon-Based Inks. J. Mater. Sci. 2019, 54, 8163–8176. [Google Scholar] [CrossRef]
  116. Mali, S.S.; Kim, H.; Patil, J.V.; Hong, C.K. Bio-Inspired Carbon Hole Transporting Layer Derived from Aloe Vera Plant for Cost-Effective Fully Printable Mesoscopic Carbon Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2018, 10, 31280–31290. [Google Scholar] [CrossRef]
  117. Gizachew, Y.T.; Escoubas, L.; Simon, J.J.; Pasquinelli, M.; Loiret, J.; Leguen, P.Y.; Jimeno, J.C.; Martin, J.; Apraiz, A.; Aguerre, J.P. Towards Ink-Jet Printed Fine Line Front Side Metallization of Crystalline Silicon Solar Cells. Sol. Energy Mater. Sol. Cells 2011, 95, S70–S82. [Google Scholar] [CrossRef]
  118. Tekin, E.; Smith, P.J.; Schubert, U.S. Inkjet Printing as a Deposition and Patterning Tool for Polymers and Inorganic Particles. Soft Matter. 2008, 4, 703. [Google Scholar] [CrossRef]
  119. Medvedeva, J.E.; Buchholz, D.B.; Chang, R.P.H. Recent Advances in Understanding the Structure and Properties of Amorphous Oxide Semiconductors. Adv. Electron. Mater. 2017, 3, 1700082. [Google Scholar] [CrossRef]
  120. Calvert, P. Inkjet Printing for Materials and Devices. Chem. Mater. 2001, 13, 3299–3305. [Google Scholar] [CrossRef]
  121. Mathies, F.; Eggers, H.; Richards, B.S.; Hernandez-Sosa, G.; Lemmer, U.; Paetzold, U.W. Inkjet-Printed Triple Cation Perovskite Solar Cells. ACS Appl. Energy Mater. 2018, 1, 1834–1839. [Google Scholar] [CrossRef]
  122. Liang, C.; Li, P.; Gu, H.; Zhang, Y.; Li, F.; Song, Y.; Shao, G.; Mathews, N.; Xing, G. One-Step Inkjet Printed Perovskite in Air for Efficient Light Harvesting. Solar RRL 2018, 2, 1700217. [Google Scholar] [CrossRef]
  123. Kipphan, H. Print Media and Electronic Media. In Handbook of Print Media; Springer: Berlin/Heidelberg, Germany, 2001; pp. 1005–1026. [Google Scholar]
  124. Hoath, S.D. Fundamentals of Inkjet Printing; Hoath, S.D., Ed.; Wiley: Hoboken, NJ, USA, 2016; ISBN 9783527337859. [Google Scholar]
  125. Eggenhuisen, T.M.; Galagan, Y.; Biezemans, A.F.K.V.; Slaats, T.M.W.L.; Voorthuijzen, W.P.; Kommeren, S.; Shanmugam, S.; Teunissen, J.P.; Hadipour, A.; Verhees, W.J.H.; et al. High Efficiency, Fully Inkjet Printed Organic Solar Cells with Freedom of Design. J. Mater. Chem. A Mater. 2015, 3, 7255–7262. [Google Scholar] [CrossRef]
  126. 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]
  127. Jung, S.; Sou, A.; Banger, K.; Ko, D.-H.; Chow, P.C.Y.; McNeill, C.R.; Sirringhaus, H. All-Inkjet-Printed, All-Air-Processed Solar Cells. Adv. Energy Mater. 2014, 4, 1400432. [Google Scholar] [CrossRef]
  128. Li, Z.; Li, P.; Chen, G.; Cheng, Y.; Pi, X.; Yu, X.; Yang, D.; Han, L.; Zhang, Y.; Song, Y. Ink Engineering of Inkjet Printing Perovskite. ACS Appl. Mater. Interfaces 2020, 12, 39082–39091. [Google Scholar] [CrossRef]
  129. Rong, Y.; Hu, Y.; Mei, A.; Tan, H.; Saidaminov, M.I.; Seok, S.I.; McGehee, M.D.; Sargent, E.H.; Han, H. Challenges for Commercializing Perovskite Solar Cells. Science 2018, 361, eaat8235. [Google Scholar] [CrossRef]
  130. Min, H.; Lee, D.Y.; Kim, J.; Kim, G.; Lee, K.S.; Kim, J.; Paik, M.J.; Kim, Y.K.; Kim, K.S.; Kim, M.G.; et al. Perovskite Solar Cells with Atomically Coherent Interlayers on SnO2 Electrodes. Nature 2021, 598, 444–450. [Google Scholar] [CrossRef]
  131. Li, P.; Liang, C.; Bao, B.; Li, Y.; Hu, X.; Wang, Y.; Zhang, Y.; Li, F.; Shao, G.; Song, Y. Inkjet Manipulated Homogeneous Large Size Perovskite Grains for Efficient and Large-Area Perovskite Solar Cells. Nano Energy 2018, 46, 203–211. [Google Scholar] [CrossRef]
  132. Li, S.-G.; Jiang, K.-J.; Su, M.-J.; Cui, X.-P.; Huang, J.-H.; Zhang, Q.-Q.; Zhou, X.-Q.; Yang, L.-M.; Song, Y.-L. Inkjet Printing of CH3NH3PbI3 on a Mesoscopic TiO2 Film for Highly Efficient Perovskite Solar Cells. J. Mater. Chem. A Mater. 2015, 3, 9092–9097. [Google Scholar] [CrossRef]
  133. Eggers, H.; Schackmar, F.; Abzieher, T.; Sun, Q.; Lemmer, U.; Vaynzof, Y.; Richards, B.S.; Hernandez-Sosa, G.; Paetzold, U.W. Inkjet-Printed Micrometer-Thick Perovskite Solar Cells with Large Columnar Grains. Adv. Energy Mater. 2020, 10, 1903184. [Google Scholar] [CrossRef]
  134. Hashmi, S.G.; Martineau, D.; Li, X.; Ozkan, M.; Tiihonen, A.; Dar, M.I.; Sarikka, T.; Zakeeruddin, S.M.; Paltakari, J.; Lund, P.D.; et al. Air Processed Inkjet Infiltrated Carbon Based Printed Perovskite Solar Cells with High Stability and Reproducibility. Adv. Mater. Technol. 2017, 2, 1600183. [Google Scholar] [CrossRef]
  135. Verma, A.; Martineau, D.; Abdolhosseinzadeh, S.; Heier, J.; Nüesch, F. Inkjet Printed Mesoscopic Perovskite Solar Cells with Custom Design Capability. Mater. Adv. 2020, 1, 153–160. [Google Scholar] [CrossRef]
  136. Karavioti, A.; Chalkias, D.A.; Katsagounos, G.; Mourtzikou, A.; Kalarakis, A.N.; Stathatos, E. Toward a Scalable Fabrication of Perovskite Solar Cells under Fully Ambient Air Atmosphere: From Spin-Coating to Inkjet-Printing of Perovskite Absorbent Layer. Electronics 2021, 10, 1904. [Google Scholar] [CrossRef]
  137. Chalkias, D.A.; Mourtzikou, A.; Katsagounos, G.; Karavioti, A.; Kalarakis, A.N.; Stathatos, E. Suppression of Coffee-Ring Effect in Air-Processed Inkjet-Printed Perovskite Layer toward the Fabrication of Efficient Large-Sized All-Printed Photovoltaics: A Perovskite Precursor Ink Concentration Regulation Strategy. Solar RRL 2022, 6, 2200196. [Google Scholar] [CrossRef]
  138. Patidar, R.; Burkitt, D.; Hooper, K.; Richards, D.; Watson, T. Slot-Die Coating of Perovskite Solar Cells: An Overview. Mater Today Commun. 2020, 22, 100808. [Google Scholar] [CrossRef]
  139. Chang, H.-M.; Chang, Y.-R.; Lin, C.-F.; Liu, T.-J. Comparison of Vertical and Horizontal Slot Die Coatings. Polym. Eng. Sci. 2007, 47, 1927–1936. [Google Scholar] [CrossRef]
  140. Ding, X.; Liu, J.; Harris, T.A.L. A Review of the Operating Limits in Slot Die Coating Processes. AIChE J. 2016, 62, 2508–2524. [Google Scholar] [CrossRef]
  141. Khambunkoed, N.; Wongratanaphisan, D.; Gardchareon, A.; Chattrapiban, N.; Homnan, S.; Songsiritthigul, P.; Ruankham, P. Slot-Die-Coated Zinc Tin Oxide Film for Carbon-Based Methylammonium-Free Perovskite Solar Cells. Surf. Rev. Lett. 2021, 28, 2150109. [Google Scholar] [CrossRef]
  142. Francis, L.F.; Roberts, C.C. Dispersion and Solution Processes. In Materials Processing; Elsevier: Amsterdam, The Netherlands, 2016; pp. 415–512. [Google Scholar]
  143. Di Risio, S.; Yan, N. Piezoelectric Ink-Jet Printing of Horseradish Peroxidase: Effect of Ink Viscosity Modifiers on Activity. Macromol. Rapid Commun. 2007, 28, 1934–1940. [Google Scholar] [CrossRef]
  144. Dai, X.; Deng, Y.; Van Brackle, C.H.; Huang, J. Meniscus Fabrication of Halide Perovskite Thin Films at High Throughput for Large Area and Low-Cost Solar Panels. Int. J. Extrem. Manuf. 2019, 1, 022004. [Google Scholar] [CrossRef]
  145. Syrrokostas, G.; Leftheriotis, G.; Yannopoulos, S.N. Double-Layered Zirconia Films for Carbon-Based Mesoscopic Perovskite Solar Cells and Photodetectors. J. Nanomater. 2019, 2019, 8348237. [Google Scholar] [CrossRef]
  146. Tian, T.; Zhong, J.; Yang, M.; Feng, W.; Zhang, C.; Zhang, W.; Abdi, Y.; Wang, L.; Lei, B.; Wu, W. Interfacial Linkage and Carbon Encapsulation Enable Full Solution-Printed Perovskite Photovoltaics with Prolonged Lifespan. Angew. Chem. Int. Ed. 2021, 60, 23735–23742. [Google Scholar] [CrossRef]
  147. Bidikoudi, M.; Simal, C.; Dracopoulos, V.; Stathatos, E. Exploring the Effect of Ammonium Iodide Salts Employed in Multication Perovskite Solar Cells with a Carbon Electrode. Molecules 2021, 26, 5737. [Google Scholar] [CrossRef]
  148. Chou, L.-H.; Yu, Y.-T.; Osaka, I.; Wang, X.-F.; Liu, C.-L. Spray Deposition of NiOx Hole Transport Layer and Perovskite Photoabsorber in Fabrication of Photovoltaic Mini-Module. J. Power Sources 2021, 491, 229586. [Google Scholar] [CrossRef]
  149. Roy, P.; Kumar Sinha, N.; Tiwari, S.; Khare, A. A Review on Perovskite Solar Cells: Evolution of Architecture, Fabrication Techniques, Commercialization Issues and Status. Solar Energy 2020, 198, 665–688. [Google Scholar] [CrossRef]
  150. Amratisha, K.; Ponchai, J.; Kaewurai, P.; Pansa-ngat, P.; Pinsuwan, K.; Kumnorkaew, P.; Ruankham, P.; Kanjanaboos, P. Layer-by-Layer Spray Coating of a Stacked Perovskite Absorber for Perovskite Solar Cells with Better Performance and Stability under a Humid Environment. Opt. Mater. Express 2020, 10, 1497. [Google Scholar] [CrossRef]
  151. Yang, Y.; Chen, H.; Zheng, X.; Meng, X.; Zhang, T.; Hu, C.; Bai, Y.; Xiao, S.; Yang, S. Ultrasound-Spray Deposition of Multi-Walled Carbon Nanotubes on NiO Nanoparticles-Embedded Perovskite Layers for High-Performance Carbon-Based Perovskite Solar Cells. Nano Energy 2017, 42, 322–333. [Google Scholar] [CrossRef]
  152. Wu, C.; Wang, K.; Jiang, Y.; Yang, D.; Hou, Y.; Ye, T.; Han, C.S.; Chi, B.; Zhao, L.; Wang, S.; et al. All Electrospray Printing of Carbon-Based Cost-Effective Perovskite Solar Cells. Adv. Funct. Mater. 2021, 31, 2006803. [Google Scholar] [CrossRef]
  153. Kavadiya, S.; Niedzwiedzki, D.M.; Huang, S.; Biswas, P. Electrospray-Assisted Fabrication of Moisture-Resistant and Highly Stable Perovskite Solar Cells at Ambient Conditions. Adv. Energy Mater. 2017, 7, 1700210. [Google Scholar] [CrossRef]
  154. Pourjafari, D.; Meroni, S.M.P.; Peralta Domínguez, D.; Escalante, R.; Baker, J.; Saadi Monroy, A.; Walters, A.; Watson, T.; Oskam, G. Strategies towards Cost Reduction in the Manufacture of Printable Perovskite Solar Modules. Energies 2022, 15, 641. [Google Scholar] [CrossRef]
  155. Meroni, S.M.P.; Worsley, C.; Raptis, D.; Watson, T.M. Triple-Mesoscopic Carbon Perovskite Solar Cells: Materials, Processing and Applications. Energies 2021, 14, 386. [Google Scholar] [CrossRef]
  156. Rong, Y.; Ming, Y.; Ji, W.; Li, D.; Mei, A.; Hu, Y.; Han, H. Toward Industrial-Scale Production of Perovskite Solar Cells: Screen Printing, Slot-Die Coating, and Emerging Techniques. J. Phys. Chem. Lett. 2018, 9, 2707–2713. [Google Scholar] [CrossRef]
  157. Priyadarshi, A.; Haur, L.J.; Murray, P.; Fu, D.; Kulkarni, S.; Xing, G.; Sum, T.C.; Mathews, N.; Mhaisalkar, S.G. A Large Area (70 cm2) Monolithic Perovskite Solar Module with a High Efficiency and Stability. Energy Environ. Sci. 2016, 9, 3687–3692. [Google Scholar] [CrossRef]
  158. Lee, H.K.H.; Barbé, J.; Meroni, S.M.P.; Du, T.; Lin, C.-T.; Pockett, A.; Troughton, J.; Jain, S.M.; De Rossi, F.; Baker, J.; et al. Outstanding Indoor Performance of Perovskite Photovoltaic Cells—Effect of Device Architectures and Interlayers. Solar RRL 2019, 3, 1800207. [Google Scholar] [CrossRef]
  159. Worsley, C.; Raptis, D.; Meroni, S.M.P.; Patidar, R.; Pockett, A.; Dunlop, T.; Potts, S.J.; Bolton, R.; Charbonneau, C.M.E.; Carnie, M.; et al. Green Solvent Engineering for Enhanced Performance and Reproducibility in Printed Carbon-Based Mesoscopic Perovskite Solar Cells and Modules. Mater. Adv. 2022, 3, 1125–1138. [Google Scholar] [CrossRef]
  160. Bogachuk, D.; Saddedine, K.; Martineau, D.; Narbey, S.; Verma, A.; Gebhardt, P.; Herterich, J.P.; Glissmann, N.; Zouhair, S.; Markert, J.; et al. Perovskite Photovoltaic Devices with Carbon-Based Electrodes Withstanding Reverse-Bias Voltages up to–9 V and Surpassing IEC 61215:2016 International Standard. Solar RRL 2022, 6, 2100527. [Google Scholar] [CrossRef]
  161. Duong, T.-T.; Hoang, P.H.; Nhan, L.T.; Van Duong, L.; Nam, M.H.; Tuan, L.Q. Multistep Spin–Spray Deposition of Large-Grain-Size CH3NH3PbI3 with Bilayer Structure for Conductive-Carbon-Based Perovskite Solar Cells. Curr. Appl. Phys. 2019, 19, 1266–1270. [Google Scholar] [CrossRef]
  162. Haruta, Y.; Ikenoue, T.; Miyake, M.; Hirato, T. One-Step Coating of Full-Coverage CsPbBr3 Thin Films via Mist Deposition for All-Inorganic Perovskite Solar Cells. ACS Appl. Energy Mater. 2020, 3, 11523–11528. [Google Scholar] [CrossRef]
  163. Verma, A.; Martineau, D.; Hack, E.; Makha, M.; Turner, E.; Nüesch, F.; Heier, J. Towards Industrialization of Perovskite Solar Cells Using Slot Die Coating. J. Mater. Chem. C Mater. 2020, 8, 6124–6135. [Google Scholar] [CrossRef]
  164. Xia, Z.; Cai, T.; Li, X.; Zhang, Q.; Shuai, J.; Liu, S. Recent Progress of Printing Technologies for High-Efficient Organic Solar Cells. Catalysts 2023, 13, 156. [Google Scholar] [CrossRef]
  165. Wiklund, J.; Karakoç, A.; Palko, T.; Yiğitler, H.; Ruttik, K.; Jäntti, R.; Paltakari, J. A Review on Printed Electronics: Fabrication Methods, Inks, Substrates, Applications and Environmental Impacts. J. Manuf. Mater. Process. 2021, 5, 89. [Google Scholar] [CrossRef]
  166. Khan, Y.; Thielens, A.; Muin, S.; Ting, J.; Baumbauer, C.; Arias, A.C. A New Frontier of Printed Electronics: Flexible Hybrid Electronics. Adv. Mater. 2020, 32, 1905279. [Google Scholar] [CrossRef]
  167. Rolston, N.; Scheideler, W.J.; Flick, A.C.; Chen, J.P.; Elmaraghi, H.; Sleugh, A.; Zhao, O.; Woodhouse, M.; Dauskardt, R.H. Rapid Open-Air Fabrication of Perovskite Solar Modules. Joule 2020, 4, 2675–2692. [Google Scholar] [CrossRef]
  168. Bishop, J.E.; Smith, J.A.; Lidzey, D.G. Development of Spray-Coated Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2020, 12, 48237–48245. [Google Scholar] [CrossRef]
  169. Dou, B.; Moore, D.T.; Whitaker, J.B.; Shaheen, S.E.; Barnes, F.S.; Zhu, K.; van Hest, M.F.A.M. One-Step High-Throughput Blade Coating of Perovskite Solar Cells. In Proceedings of the 2018 IEEE 7th World Conference on Photovoltaic Energy Conversion (WCPEC) (A Joint Conference of 45th IEEE PVSC, 28th PVSEC & 34th EU PVSEC), Waikoloa Village, HI, USA, 10–15 June 2018; pp. 2799–2802. [Google Scholar]
  170. Marques, A.S.; Faria, R.M.; Freitas, J.N.; Nogueira, A.F. Low-Temperature Blade-Coated Perovskite Solar Cells. Ind. Eng. Chem. Res. 2021, 60, 7145–7154. [Google Scholar] [CrossRef]
  171. Chang, X.; Fan, Y.; Zhao, K.; Fang, J.; Liu, D.; Tang, M.-C.; Barrit, D.; Smilgies, D.-M.; Li, R.; Lu, J.; et al. Perovskite Solar Cells toward Eco-Friendly Printing. Research 2021, 2021, 9671892. [Google Scholar] [CrossRef]
  172. Galagan, Y.; Di Giacomo, F.; Gorter, H.; Kirchner, G.; de Vries, I.; Andriessen, R.; Groen, P. Roll-to-Roll Slot Die Coated Perovskite for Efficient Flexible Solar Cells. Adv. Energy Mater. 2018, 8, 1801935. [Google Scholar] [CrossRef]
  173. Parida, B.; Singh, A.; Soopy, A.K.K.; Sangaraju, S.; Sundaray, M.; Mishra, S.; Liu, S.; Najar, A. Recent Developments in Upscalable Printing Techniques for Perovskite Solar Cells. Adv. Sci. 2022, 9, 2200308. [Google Scholar] [CrossRef]
  174. Hu, Y.; Chu, Y.; Wang, Q.; Zhang, Z.; Ming, Y.; Mei, A.; Rong, Y.; Han, H. Standardizing Perovskite Solar Modules beyond Cells. Joule 2019, 3, 2076–2085. [Google Scholar] [CrossRef]
  175. Mathies, F.; List-Kratochvil, E.J.W.; Unger, E.L. Advances in Inkjet-Printed Metal Halide Perovskite Photovoltaic and Optoelectronic Devices. Energy Technol. 2020, 8, 1900991. [Google Scholar] [CrossRef]
  176. Bishop, J.E.; Mohamad, D.K.; Wong-Stringer, M.; Smith, A.; Lidzey, D.G. Spray-Cast Multilayer Perovskite Solar Cells with an Active-Area of 1.5 cm2. Sci. Rep. 2017, 7, 7962. [Google Scholar] [CrossRef]
  177. Giuri, A.; Saleh, E.; Listorti, A.; Colella, S.; Rizzo, A.; Tuck, C.; Esposito Corcione, C. Rheological Tunability of Perovskite Precursor Solutions: From Spin Coating to Inkjet Printing Process. Nanomaterials 2019, 9, 582. [Google Scholar] [CrossRef]
  178. Cao, K.; Zuo, Z.; Cui, J.; Shen, Y.; Moehl, T.; Zakeeruddin, S.M.; Grätzel, M.; Wang, M. Efficient Screen Printed Perovskite Solar Cells Based on Mesoscopic TiO2/Al2O3/NiO/Carbon Architecture. Nano Energy 2015, 17, 171–179. [Google Scholar] [CrossRef]
Figure 1. Energy band diagram of various metal oxides used in perovskite solar cells with carbon electrode. Reproduced with permission from Ref. [8].
Figure 1. Energy band diagram of various metal oxides used in perovskite solar cells with carbon electrode. Reproduced with permission from Ref. [8].
Materials 16 03917 g001
Figure 2. Conventional PSC (left) and C-PSC (right) working principles upon light illumination. Efn and Efp represent the electron and hole quasi-Fermi levels, respectively. Reproduced with permission from Ref. [11].
Figure 2. Conventional PSC (left) and C-PSC (right) working principles upon light illumination. Efn and Efp represent the electron and hole quasi-Fermi levels, respectively. Reproduced with permission from Ref. [11].
Materials 16 03917 g002
Figure 3. Schematic view of the spin coating stages (ω is the angular velocity).
Figure 3. Schematic view of the spin coating stages (ω is the angular velocity).
Materials 16 03917 g003
Figure 4. (a) X-ray diffraction pattern; (b) Raman spectrum with inset of SEM image of brookite nanorods and TEM image of brookite nanocubes; (c) EQE of the champion devices comparing brookite nanorods and nanocubes with commercial anatase; and (d) IS representation along with the fitted electrical circuit diagrams. Reproduced with permission from Ref. [63].
Figure 4. (a) X-ray diffraction pattern; (b) Raman spectrum with inset of SEM image of brookite nanorods and TEM image of brookite nanocubes; (c) EQE of the champion devices comparing brookite nanorods and nanocubes with commercial anatase; and (d) IS representation along with the fitted electrical circuit diagrams. Reproduced with permission from Ref. [63].
Materials 16 03917 g004
Figure 5. SEM images of (a) CsPbIBr2 cell based on F–SnO2 film; (b) CsPbIBr2 cell based on NP-SnO2 film; (c) cross section of CsPbIBr2 film based on F–SnO2 film; (d) cross section of CsPbIBr2 cell based on NP-SnO2 film; (e) CsPbIBr2 based on F–SnO2 film; (f) the light path in the CsPbIBr2 cell based on NP-SnO2 film. Reproduced with permission from Ref. [67].
Figure 5. SEM images of (a) CsPbIBr2 cell based on F–SnO2 film; (b) CsPbIBr2 cell based on NP-SnO2 film; (c) cross section of CsPbIBr2 film based on F–SnO2 film; (d) cross section of CsPbIBr2 cell based on NP-SnO2 film; (e) CsPbIBr2 based on F–SnO2 film; (f) the light path in the CsPbIBr2 cell based on NP-SnO2 film. Reproduced with permission from Ref. [67].
Materials 16 03917 g005
Figure 6. Schematic view of the contact angles of ZnO NPs solution dropped on FTO substrates with DI treatment and without DI treatment and the contact angles of the perovskite precursor solution dropped on ZnO/FTO films with and without DI treatment.
Figure 6. Schematic view of the contact angles of ZnO NPs solution dropped on FTO substrates with DI treatment and without DI treatment and the contact angles of the perovskite precursor solution dropped on ZnO/FTO films with and without DI treatment.
Materials 16 03917 g006
Figure 7. XRD spectra of the CsPbBr3 films prepared on different ETLs with corresponding SEM image of CsPbBr3 film grown on Nb2O5/FTO and Li: Nb2O5/FTO. Reproduced with permission from Ref. [74].
Figure 7. XRD spectra of the CsPbBr3 films prepared on different ETLs with corresponding SEM image of CsPbBr3 film grown on Nb2O5/FTO and Li: Nb2O5/FTO. Reproduced with permission from Ref. [74].
Materials 16 03917 g007
Figure 8. Schematic of perovskite cell fabrication processes using different nanomaterials dispersed in the perovskite layer. (a) B-MWNTs on PbI2 layer followed by MAPbI3 perovskite film, reprinted with permission from [78]; Copyright {2017} American Chemical Society. (b) CNPs dispersed in the antisolvent as chlorobenzene; reproduced with permission from Ref. [79].
Figure 8. Schematic of perovskite cell fabrication processes using different nanomaterials dispersed in the perovskite layer. (a) B-MWNTs on PbI2 layer followed by MAPbI3 perovskite film, reprinted with permission from [78]; Copyright {2017} American Chemical Society. (b) CNPs dispersed in the antisolvent as chlorobenzene; reproduced with permission from Ref. [79].
Materials 16 03917 g008
Figure 9. (a) CQDs dispersed in the antisolvent as ethyl acetate, (b) XPS spectra of EACQDs, (c) deconvolution of C, and (d) XRD image of pristine and 0.01 EACQDs optimized PSCs, reproduced with permission from Ref. [80].
Figure 9. (a) CQDs dispersed in the antisolvent as ethyl acetate, (b) XPS spectra of EACQDs, (c) deconvolution of C, and (d) XRD image of pristine and 0.01 EACQDs optimized PSCs, reproduced with permission from Ref. [80].
Materials 16 03917 g009
Figure 10. Schematic illustration of the basic principles of screen printing. Reproduced with permission from Ref. [82].
Figure 10. Schematic illustration of the basic principles of screen printing. Reproduced with permission from Ref. [82].
Materials 16 03917 g010
Figure 11. (A) Scheme of fully printable triple-stack mesoporous carbon-based perovskite solar cells (PSC), (B) Band gap diagram of each layer present in the carbon-based PSC, and (C) crystalline structure of MAPbI3 perovskite. Reproduced with permission from Ref. [86].
Figure 11. (A) Scheme of fully printable triple-stack mesoporous carbon-based perovskite solar cells (PSC), (B) Band gap diagram of each layer present in the carbon-based PSC, and (C) crystalline structure of MAPbI3 perovskite. Reproduced with permission from Ref. [86].
Materials 16 03917 g011
Figure 12. (A) Implementation of an extra Al2O3 layer with m-TiO2 and m-ZrO2 via a vacuum technique, (B) infiltration of perovskite precursor solution through the layers, and (C) energy band diagram of hole-conductor-free C-PSC with Al2O3 interlayer. Reproduced with permission from Ref. [96].
Figure 12. (A) Implementation of an extra Al2O3 layer with m-TiO2 and m-ZrO2 via a vacuum technique, (B) infiltration of perovskite precursor solution through the layers, and (C) energy band diagram of hole-conductor-free C-PSC with Al2O3 interlayer. Reproduced with permission from Ref. [96].
Materials 16 03917 g012
Figure 13. XRD analyses of black NiOx acquired from Sigma-Aldrich: (a) powder of NiO, Ni(OH)2, NiOOH, and Ni2O3, and (b) mixture of Ni2+ and Ni3+.Reproduced with permission from [98].
Figure 13. XRD analyses of black NiOx acquired from Sigma-Aldrich: (a) powder of NiO, Ni(OH)2, NiOOH, and Ni2O3, and (b) mixture of Ni2+ and Ni3+.Reproduced with permission from [98].
Materials 16 03917 g013
Figure 14. Cross-section of the perovskite/triple-stack mesoporous carbon-based solar cell with Co2O3 hole-transporting layer: (a) low magnification, and (b) high magnification, (c) J-V characteristics of standard carbon cell with an aperture area of 0.09 cm2 (under 1 Sun) without and with Co3O4 layer, and (d) J-V characteristics of module with an active area of 70 cm2 and with Co3O4. Reproduced with permission from Ref. [97].
Figure 14. Cross-section of the perovskite/triple-stack mesoporous carbon-based solar cell with Co2O3 hole-transporting layer: (a) low magnification, and (b) high magnification, (c) J-V characteristics of standard carbon cell with an aperture area of 0.09 cm2 (under 1 Sun) without and with Co3O4 layer, and (d) J-V characteristics of module with an active area of 70 cm2 and with Co3O4. Reproduced with permission from Ref. [97].
Materials 16 03917 g014
Figure 15. Construction of the: (a) single carbon electrode composed of a mesoporous carbon layer, and (b) double-layer carbon electrode composed of a mesoporous carbon below a conductor carbon layer. Reproduced with permission from Ref. [105].
Figure 15. Construction of the: (a) single carbon electrode composed of a mesoporous carbon layer, and (b) double-layer carbon electrode composed of a mesoporous carbon below a conductor carbon layer. Reproduced with permission from Ref. [105].
Materials 16 03917 g015
Figure 16. Working principles of inkjet printers: (a) CIJ; and (b) DOD. Reproduced with permission from Ref [124].
Figure 16. Working principles of inkjet printers: (a) CIJ; and (b) DOD. Reproduced with permission from Ref [124].
Materials 16 03917 g016
Figure 17. X-ray diffraction (XRD) patterns after infiltrating of CH3NH3PbI3 (MAPI) and annealing. The dominating (101) reflection at 2θ = 25.3° is attributed to anatase TiO2 (inset). At reduced intensity scale, the characteristic reflections of the cubic phase of the MAPI perovskite crystal are clearly visible (blue squares). The absence of a reflection peak at 2θ = 12.5° reveals that all the precursor ink has been transformed to the hybrid metal halide perovskite and that no un-desirable PbI2 is present. Reproduced with permission from Ref. [135].
Figure 17. X-ray diffraction (XRD) patterns after infiltrating of CH3NH3PbI3 (MAPI) and annealing. The dominating (101) reflection at 2θ = 25.3° is attributed to anatase TiO2 (inset). At reduced intensity scale, the characteristic reflections of the cubic phase of the MAPI perovskite crystal are clearly visible (blue squares). The absence of a reflection peak at 2θ = 12.5° reveals that all the precursor ink has been transformed to the hybrid metal halide perovskite and that no un-desirable PbI2 is present. Reproduced with permission from Ref. [135].
Materials 16 03917 g017
Figure 18. Schematic illustration of the main components in a slot-die processing set-up.
Figure 18. Schematic illustration of the main components in a slot-die processing set-up.
Materials 16 03917 g018
Figure 19. Schematic illustration of the “doctor blade” film deposition process.
Figure 19. Schematic illustration of the “doctor blade” film deposition process.
Materials 16 03917 g019
Figure 20. UV-Vis absorbance spectra of ZrO2 film (N_ZrO2) and double-layered film (DL_ZrO2) deposited on glass. Reproduced with permission from Ref. [145].
Figure 20. UV-Vis absorbance spectra of ZrO2 film (N_ZrO2) and double-layered film (DL_ZrO2) deposited on glass. Reproduced with permission from Ref. [145].
Materials 16 03917 g020
Figure 21. Schematic illustration of the spray coating deposition process.
Figure 21. Schematic illustration of the spray coating deposition process.
Materials 16 03917 g021
Figure 22. (a) Performance of 70 cm2 mini-module with CsBr-modified TiO2 layer in comparison with a standard carbon-based perovskite solar cell; (b) Image of a fully printable mini-module. Reproduced with permission from Ref. [4].
Figure 22. (a) Performance of 70 cm2 mini-module with CsBr-modified TiO2 layer in comparison with a standard carbon-based perovskite solar cell; (b) Image of a fully printable mini-module. Reproduced with permission from Ref. [4].
Materials 16 03917 g022
Table 1. Printing parameters, ink properties and the highest reported efficiency of the C-PSCs [125,140,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178].
Table 1. Printing parameters, ink properties and the highest reported efficiency of the C-PSCs [125,140,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178].
(1) Method(2) Printing Parameters(3) Layers Ink PropertiesRate of Ink Usage (U) & Waste (W) (4) Film Thickness Range (µm) Printing Speed (m/min)Throughput (m2/min)ScalabilityStructureThe Highest Reported PCE (%)/Active Area (cm2)Ref.
Viscosity (Pa.s)Surface
Tension (mN/m)
Max Particle
Size (nm)
Spin coatingspinning mode (dynamic, static), spin speed and time, and acceleration/deceleration rate, drop volumeETL, HTL, PerovskiteLowUp to 25Up to 50U: low, W: very high0.1–0.5Very low-batch sizeVery low -batch sizeNo BL-TiO2 (spin coating)/m-TiO2 (spin coating)/MAPI (spin coating)/C (blade coating)17.42/0.1[34]
BL-TiO2 (spin-coating)/m-TiO2 (spin-coating)/ZrO2 (spin-coating)/MAPbI3: EACQDs as antisolvent (spin-coating)/B-MWNTs (drop casting)/carbon (screen printed)15.14/0.06[80]
BL-TiO2 (spray pyrolysis)/m-TiO2 (spin coating)/Al2O3 (spin-coating)/MAPbI3 (PbI2 embedded with B-MWNTs, spin-coating)/boron (drop casting)15.23/No data[78]
Screen printingmesh and squeegee type, mesh tension, distance between the screen and substrate, angle between the squeegee and screen, pressure and speed of the squeegee, amount of loaded pasteETL, HTL, Perovskite, CarbonModerate-High
2.5- 5
Up to 5020–20000 (1/10th of mesh opening)U: high, W: low0.5–20Flatbed: up to 35 Rotary: up to 100Up to 70YesBL-TiO2 (spray pyrolysis)/BaSnO3 (homemade)/m-ZrO2/C (all screen- printed)/PbI2.MAI, MACl (drop casting)14.77/0.1[92]
BL-TiO2 (spray pyrolysis)/m-TiO2/Al2O3 (homemade)/NiO (homemade)/C (all screen-printed)/MAPbI3 (drop casting)15.03/0.7[178]
BL-TiO2/m-TiO2/ZrO2/Cu:NiOx (homemade)/C (all screen-printed)/(5-AVA)x(MA)1−xPbI3 (drop casting)12.79/0.8[101]
BL-TiO2 (spray pyrolysis)/m-TiO2/ZrO2/C (homemade boron-doped graphite) (all screen-printed)/(5-AVA)x(MA)1−xPbI3 (drop casting)13.6/No data[102]
Inkjet printingjetting voltage, jetting frequency, drop spacing, amount of loaded ink, substrate temperatureETL, HTL, PerovskiteLow (0.001–0.05)Up to 40Up to 50 (1/10th of nozzle diameter)U: high, W: low0.1- 1Up to 10Up to 6 Limited BL-TiO2 (spray pyrolysis)/m-TiO2 (screen printing)/m-ZrO2 (screen printing)/MAPI (inkjet printing)/C (screen printing)9.53/0.16[134]
BL-TiO2/m-TiO2/m-ZrO2/MAPI/C
All layers were deposited by inkjet printing except carbon layer by screen printing)
9.10/1.5[135]
Slot-die coatingcoating speed, flow rate, coating gap, substrate temperature, ETL, HTL, PerovskiteLow-High (0.001–5) Up to 25Up to 200U: high, W: low0.1- 2Up to 5No dataYesBL-TiO2/ZTO/CsFA perovskite/spiro- OMeTAD/C
(All layers were deposited by slot die)
9.92/No data[141]
Blade coatingcoating speed, coating gap (distance between blade and substrate), amount of loaded ink, substrate temperature,ETL, HTL, Perovskite, CarbonLow (0.001–0.05)Up to 2520-few micrometresU: high, W: moderate0.1- 10Up to 2Up to 1.5 (theoretical data)YesITO/APTES-linked C60 (ETL)/MAPI/C
(All layers were deposited by blade coating)
18.64/0.08[146]
BL-TiO2 (spin coating)/m-TiO2 (spin coating)/m-ZrO2 (blade coating)/MAPI (drop casting)/NiO NPs (spin coating)/MWCNT as C (electrospray)15.80/0.08[151]
Spray coatingnozzle diameter and type, spray speed, distance and angle between nozzle and substrate, air or gas pressure, flow rate, substrate temperatureBL, ETL, HTL, Perovskite, Carbon (rarely)Low-Moderate (0.001–2.5)Up to 2020-few µm *(usually 1/10th of nozzle diameter)U: low to high, W: low to high **0.05- 1Up to 12No data YesBL-TiO2/FA0.85MA0.15PbI2.85Br0.15/C
(All layers were deposited by electrospray)
14.41/0.096[152]
1-All methods are compatible with commonly used substrates for C-PSCs, i.e., glass and plastic (flexible) substrates. 2-Parameters which should be controlled during printing to obtain desirable thickness and quality of the deposited layer. 3-The most common layers in C-PSCs that can be deposited by the corresponding printing technique. 4-Depending on the layer. * Depending on the spray type. ** Depending on the substrate area and spray system.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pourjafari, D.; García-Peña, N.G.; Padrón-Hernández, W.Y.; Peralta-Domínguez, D.; Castro-Chong, A.M.; Nabil, M.; Avilés-Betanzos, R.C.; Oskam, G. Functional Materials for Fabrication of Carbon-Based Perovskite Solar Cells: Ink Formulation and Its Effect on Solar Cell Performance. Materials 2023, 16, 3917. https://doi.org/10.3390/ma16113917

AMA Style

Pourjafari D, García-Peña NG, Padrón-Hernández WY, Peralta-Domínguez D, Castro-Chong AM, Nabil M, Avilés-Betanzos RC, Oskam G. Functional Materials for Fabrication of Carbon-Based Perovskite Solar Cells: Ink Formulation and Its Effect on Solar Cell Performance. Materials. 2023; 16(11):3917. https://doi.org/10.3390/ma16113917

Chicago/Turabian Style

Pourjafari, Dena, Nidia G. García-Peña, Wendy Y. Padrón-Hernández, Diecenia Peralta-Domínguez, Alejandra María Castro-Chong, Mahmoud Nabil, Roberto C. Avilés-Betanzos, and Gerko Oskam. 2023. "Functional Materials for Fabrication of Carbon-Based Perovskite Solar Cells: Ink Formulation and Its Effect on Solar Cell Performance" Materials 16, no. 11: 3917. https://doi.org/10.3390/ma16113917

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