Next Article in Journal
Electrochemical Synthesis of Nb-Doped BaTiO3 Nanoparticles with Titanium-Niobium Alloy as Electrode
Next Article in Special Issue
Towards a CdTe Solar Cell Efficiency Promotion: The Role of ZnO:Al and CuSCN Nanolayers
Previous Article in Journal
TiO2/SnO2 Bilayer Electron Transport Layer for High Efficiency Perovskite Solar Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Diethanolamine Modified Perovskite-Substrate Interface for Realizing Efficient ESL-Free PSCs

1
Department of Chemical & Petroleum Engineering, United Arab Emirates University, Al Ain P.O. Box 15551, United Arab Emirates
2
National Water and Energy Center, United Arab Emirates University, Al Ain P.O. Box 15551, United Arab Emirates
3
College of Physics and Energy, Fujian Normal University, Fuzhou 350007, China
4
College of Natural and Health Sciences, Zayed University, Abu Dhabi P.O. Box 144534, United Arab Emirates
5
Department of Metallurgical and Materials Engineering, Atilim University, Ankara 06836, Turkey
6
Department of Physics, United Arab Emirates University, Al Ain P.O. Box 15551, United Arab Emirates
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(2), 250; https://doi.org/10.3390/nano13020250
Submission received: 7 December 2022 / Revised: 31 December 2022 / Accepted: 4 January 2023 / Published: 6 January 2023
(This article belongs to the Special Issue State-of-the-Art Nanomaterials for Solar Cells)

Abstract

:
Simplifying device layout, particularly avoiding the complex fabrication steps and multiple high-temperature treatment requirements for electron-selective layers (ESLs) have made ESL-free perovskite solar cells (PSCs) attractive. However, the poor perovskite/substrate interface and inadequate quality of solution-processed perovskite thin films induce inefficient interfacial-charge extraction, limiting the power conversion efficiency (PCEs) of ESL-free PSCs. A highly compact and homogenous perovskite thin film with large grains was formed here by inserting an interfacial monolayer of diethanolamine (DEA) molecules between the perovskite and ITO substrate. In addition, the DEA created a favorable dipole layer at the interface of perovskite and ITO substrate by molecular adsorption, which suppressed charge recombination. Comparatively, PSCs based on DEA-treated ITO substrates delivered PCEs of up to 20.77%, one of the highest among ESL-free PSCs. Additionally, this technique successfully elongates the lifespan of ESL-free PSCs as 80% of the initial PCE was maintained after 550 h under AM 1.5 G irradiation at ambient temperature.

Graphical Abstract

1. Introduction

Within the last few years, PSCs have progressed significantly in their PCEs, rising from 3.8 to 25.7% [1]. PSCs are conventionally fabricated using two architectures: planar and mesoscopic [2,3,4,5,6]. Mesoscopic PSCs are generally made up of an electron-selective layer (ESL), a mesoscopic TiO2 or Al2O3 scaffold, a perovskite layer, a hole-selective layer (HSL), and back and front electrodes [7]. Further research found that PSCs can produce a PCE of over 15% utilizing a straightforward planar architecture, achieving high efficiency without the need for a mesoscopic scaffold [8,9,10,11,12,13,14,15,16,17]. Due to perovskite’s ambipolar characteristic, additional device architectural simplification is possible in order to fabricate efficient PSCs without using complicated fabrication methods [18,19]. In this scenario, HSL-free PSCs can be produced without manufacturing or utilizing costly organic HSLs [20,21].
The subsequent focus in PSCs is on the simplified fabrication technique that offers simple and inexpensive mass production in order to become economically relevant in the context of high efficiency and reasonable stabilities [22]. In this context, numerous researchers constructed ESL-free PSCs with reasonable PCEs [23]. ESL-free PSCs have a high device stability potential in addition to decent PCEs and do not need the time-consuming manufacturing of ESLs, particularly the numerous high-temperature processes required to create TiO2 ESL. Since it is known that TiO2 causes the perovskite layer to decompose by creating oxygen vacancies, the removal of the TiO2 ESL is mostly responsible for the high device stability [24]. However, because of the inefficient electron transport and the associated carrier recombination at the interface of transparent conductive oxides (TCO) and perovskite, the performance of ESL-free PSCs is still inferior to that of conventional devices [25,26,27]. An additional reason for the low device efficiency of the ESL-free PSCs is the rough and uneven perovskite film when no ESL is fabricated, which will produce many open grain boundaries and lead to severe recombination as a result of the shunt pathways between the HSL and TCO. Surface modification methods were recently suggested to reduce charge-carrier recombination at the TCO/perovskite interface and to initiate charge collection at the relevant contact [28,29].
The efficacy of the resulting ESL-free PSCs is enhanced by carefully adjusting the compositions, architecture, and morphology of the perovskite layer as well as pre-treating the TCO (UV/O3 irradiation) to prevent carrier recombination in the perovskite or at the TCO/perovskite interface [24]. These ESL-free PSCs, however, displayed relatively lower PCEs (<18%) than those of their ESL-containing counterparts. Other methods for improving device performance and surface modification of TCO substrates include bathocuproine [25], polar non-conjugated small molecule modifier [30], and hydroxyethyl-functionalized imidazolium iodide ionic liquid [31], and self-assembled fullerene monolayer [32]. Recently, PCEs of ESL-free PSCs >19% have been reported employing sodium fluoride [33] and tetramethylammonium hydroxide [34] as TCO modifiers. Nevertheless, most of the modifications in the aforementioned approaches need additional or difficult processes, which consumes time and energy. These modifiers are also regarded for their hygroscopicity, insulating properties, and high material costs.
In the present study, we used a low-cost, very simple Indium-Tin Oxide (ITO)-surface modification with DEA to enhance the ITO/perovskite interaction and perovskite crystallization. Because amine groups typically interact with the ITO surface and Pb-based substances are reported to link with –OH groups, we chose DEA, which contains both –OH and amine groups, as the contact modifier between ITO and perovskite [35]. As anticipated, the DEA molecular interlayer interacted with ITO and perovskite, which strengthened the interface contact and elevated the crystallinity of the perovskite. A dipole layer that is well-orientated for electron extraction was also produced by molecular adsorption. The electron extraction rate was significantly improved as a result, and the ESL-free PSC with the ITO/DEA/Perovskite/spiro/Au device structure demonstrated significantly better performance (PCE: 20.77%) than an ITO/Perovskite/spiro/Au device (PCE: 18.65%).

2. Experimental Procedure

2.1. Solution Preparation

Perovskite solution was obtained by mixing anhydrous dimethylformamide/dimethylsulfoxide (600 mg/78 mg) with 0.15 M of formamidinium lead iodide (FAI), 0.85 M of methylammonium lead iodide (MAI), and 1.025 M of PbI2. This solution was stirred for 4 h in a glove box. Spiro-OMeTAD was dissolved in chlorobenzene at a concentration of 80 mg/mL to make the precursor for the hole-selective layer. Tert-butylpyridine (28.5 μL in 1 mL of chlorobenzene) and lithium bis-(trifluoromethanesulfonyl) imide (8.75 mg/mL) were then added as additives, and the mixture was stirred in a glove box for 6 h.

2.2. Device Fabrication

The ITO substrates were ultrasonically cleaned with detergent solution, deionized water, acetone, ethyl alcohol, and deionized water for 20 min, respectively. The ITO substrates were exposed to UV-ozone for 20 min after drying. The cleaned ITO substrates were spin-coated with the DEA pure solvent for 30 s at 2000 rpm and then dried for 10 min at 100 °C. The bare cleaned ITO and DEA-treated ITO substrates were spin-coated with the perovskite precursor at 4000 rpm for 25 s to form the perovskite layers. Eighteen seconds before the end of the spin-coating process, 0.8 mL of diethyl ether was dripped on the perovskite surface. The resulting perovskite layers were heated for 15 min at 130 °C before naturally cooling to ambient temperature. Subsequently, both perovskite layers were spin-coated with the Spiro-OMeTAD solution for 30 s at 4000 rpm. The devices were finished by evaporating 60 nm Au electrodes onto the ITO/perovskite/spiro-OMeTAD and ITO/DEA/perovskite/spiro-OMeTAD.

2.3. Characterizations

A Thermo Fisher Scientific (Shanghai, China) ESCALAB 250Xi under 10−9 Torr vacuum with a monochromatic Al-Kα X-ray source was used to collect the XPS spectra. A scanning electron microscope (Hitachi, Chiyoda, Japan, S-4800) and an atomic force microscope (Agilent Keysight (Beijing, China) AFM-5500) were utilized to study the morphologies. The perovskite’s crystallinity was examined using an X-ray diffractometer (Bruker D8 Advance (Bruker AXS Inc.), Cu-Kα radiation of 0.15406 nm). The UPS spectra were obtained using the ESCALAB 250Xi Contact angle images were obtained with an Ossila Contact Angle Goniometer. The FT-IR spectra were collected by using a Jasco (Tokyo, Japan) FT-IR spectrometer. Using a UV-Vis spectrophotometer (UV-2600), the absorption spectrum was measured. The steady-state PL spectra of the produced samples were looked at utilizing an Edinburgh PLS 980. Utilizing a transient state spectrophotometer (Edinburgh Institute F900) and a 485 nm laser, the TRPL decay of the perovskite layers was monitored. Under AM 1.5 G illumination with a power intensity of 100 mW cm−2, the J-V characteristic curves were measured with a source meter (Keithley (Cleveland, OH, USA) 2400) utilizing forward (−0.1 to 1.2 V) or reverse (1.2 to −0.1 V) scans from a solar simulator (XES-301S+EL-100). The delay time was set to 10 ms, and the step voltage was set to 12 mV. The EQE was calculated using QE-R systems (Enli Tech., Shanghai, China). The EIS measurement was carried out using an electrochemical workstation (Zahner Zennium). The EIS data were fitted with help of ZSim-software version 3.20 with equivalent circuit parameters of R1 = 0.116, C = 0.00001 and R2 = 54.1 (for ITO-based device), and R1 = 0.116, C = 0.000001 and R2 = 29.6 (for DEA-treated ITO-based device).

3. Results and Discussion

The cleaned ITO substrates were first spin-coated with DEA at 2000 rpm for 30 s, then dried at 100 °C for 10 min. X-ray photoelectron spectroscopy (XPS) was carried out to investigate the effect of DEA treatment on the surface chemistry of ITO. The XPS survey spectra of bare ITO and DEA-treated ITO are presented in Figure 1. All spectra were calibrated using the C 1s peak of 284.9 eV binding energy as a reference point. The bare ITO and DEA-treated ITO share the same peaks except at 400.4 eV, which is assigned to N 1s, showing that the DEA successfully adsorbed on the ITO substrate, as can be seen in Figure 1a,b. Figure 1c displays the high-resolution N 1s spectra of ITO/DEA, which can be divided into three peaks at 401.4, 400.4, and 399.7 eV. Sigma-state, including species containing N, can be attributed to the peaks at 401.4 eV and 400.4 eV. Additionally, it has been stated that an amine group correlates to the peak at 399.7 eV [36,37]. The shift in the binding energy, as observed in the Sn 3d (Figure 1d) signals, is indicative of a chemical interaction between DEA and the ITO substrate. This suggests a dipole interaction between ITO and DEA that is mediated by an electric field, which lowers the attractive force of the Sn nuclei and increases the Sn atom’s outer-shell electron density [38].
To investigate the impact of DEA coating on the surface morphology of ITO, scanning electron microscopy (SEM) was conducted on bare ITO and DEA-treated ITO. Figure 2a,b displays the surface SEM images of ITO and ITO/DEA, respectively. As can be seen, both surfaces are compact without pinholes. However, because the DEA layer is so thin, it is hard to see the DEA-coated ITO surface clearly. Figure S1 shows that this thin layer of DEA had little influence on ITO transparency across a wide spectrum of light, indicating a great potential for high photocurrent output. The morphologies of the bare ITO and DEA-treated ITO were investigated with the help of atomic force microscopy (AFM), as shown in Figure 2c,d. The alteration in the morphology following DEA coating is not very noticeable, similar to the SEM images. However, in the DEA-treated ITO substrate, a reasonable decrease in the root-mean-square (RMS) of surface roughness from 2.8 to 0.9 nm is observed. The DEA coating may considerably minimize the surface defect states and promote the deposition of high-quality perovskite thin film because of the reduced roughness. We also used Kelvin probe force microscopy (KPFM) to examine bare and DEA-treated ITO, as displayed in Figure S2. This revealed that the surface potential distribution is consistent with the aforementioned surface morphologies, and the average potential of the ITO/DEA substrate is −110 mV compared to the 0.18 mV of bare ITO, which indicates desirable Fermi-level of the ITO/DEA and better band alignment with perovskite [24,39].
In order to get highly efficient PSCs in ESL-free architectures, one of the main obstacles is the energy-level mismatch between ITO and the perovskite, so it is essential to investigate how DEA affects the energy levels of ITO. Figure 3a,b illustrates the outcomes of an ultraviolet photoelectron spectroscopy (UPS) experiment. For the purpose of calculating Ecutoff (secondary electron edges) and Eonset (Fermi edges), the intercepts of the tangents of the peaks with the extrapolated baselines were used. By estimating the difference between the incident photon energy (light source He I, 21.22 eV) and Ecutoff, the Fermi-level (EF) values of the bare and DEA-treated ITO were quantitatively determined. The UPS show an Ecutoff of 16.64 eV and 17.21 eV for bare ITO and ITO/DEA, respectively. The EF values of bare ITO and ITO/DEA were estimated to be −4.58 and −4.01 eV, respectively, using the following equation.
E F = E c u t o f f 21.22   eV  
The EF of the ITO was significantly shifted by −0.57 eV after DEA treatment, which would be advantageous for the cathode’s ability to collect electrons. This is because, following DEA treatment, a favorable dipole moment formed on the surface of the ITO. As stated in previous studies, Pb-based compounds interact with –OH groups, while the amine groups in DEA tend to link with the ITO surface [34,35]. As can be seen, the DEA modification resulted in a −0.57 eV shift in the Fermi level of ITO, which should be attributed to a chemical reaction between DEA and ITO, as supported by the FT-IR data (Figure S4). The appearance of the N-atom signal in Figure 1 demonstrates the interaction between the N-atom of DEA and ITO. The -NH- (from the amine group) should function as an electron donor to engage in chemisorption and modify the energy levels of ITO. To examine the surface wettability of ITO treated with DEA, the water contact angle was measured (Figure 3c). Since DEA possesses hydrophilic –OH and amine groups, the water contact angles of the ITO surface decrease from 70.1° to 59.6°, showing greater wetting capabilities of DEA-treated ITO. The superior wettability can decrease the nucleation barrier and aid in the formation of a high-quality perovskite layer.
The surface morphologies and cross-sectional images of the perovskite layers prepared on bare ITO and ITO/DEA substrates were taken with help of SEM, as depicted in Figure 4. It is evident that the perovskite layers that were formed on ITO/DEA have densely packed, considerably larger grains. In order to facilitate charge transfer, the perovskite layer formed on ITO/DEA substrate displays entire coverage without any grain boundary gaps. The larger perovskite grains and improved morphology on the ITO/DEA substrate can be attributed to the favorable interaction of the –OH group with Pb2+ [35], which promotes the formation of larger grains and passivation of the perovskite’s grain boundaries. A schematic representation of the DEA interaction is shown in Figure S3a, where the –OH groups of the DEA would interact with the Pb-atoms in the perovskite thin film while the N-atom of DEA tends to create a chemical connection with the Sn atom on the ITO surface. According to other researchers, the interaction between the –OH group and Pb2+ can improve the penetration of PbI2 into a TiO2 mesoporous layer [40]. Through the direct mixing of DEA solution and PbI2 powder, we were able to further test the possible interaction between DEA and PbI2. Intriguingly, adding PbI2 to DEA has significantly altered the appearance, likely creating new products. As seen in Figure S3b, the colorless DEA solution and the yellow PbI2 powder were combined to create an entirely distinct product that was white in color. The white mixture turns into a colorless liquid after heating (Figure S3c). To further confirm these interactions, Fourier-transform infrared (FT-IR) spectra in the wavenumber range of 600 to 4000 cm−1 were carried out, as shown in Figure S4. The spectrum of the ITO substrate shows typical broad bands below 800 cm−1, which is assigned to the stretching vibration of In-O or Sn-O bonds [41]. The DEA-modified ITO shows broad and strong bands below 800 cm−1. In addition to the broad and strong bands, the DEA-modified ITO displays an additional sharp and strong peak at 630 cm−1, which is ascribable to the stretching vibration of the Sn-N bond. Therefore, it is reasonable to assume that the DEA and ITO substrate were linked by a Sn-N bond, exposing the –OH groups for interaction with the perovskite layer that was later added. The DEA and PbI2/DEA were characterized using FT-IR spectroscopy, and the variation in –OH absorption peak was monitored. Broad absorption peaks can be seen in the DEA and DEA/PbI2 spectra between 1052 and 1458 cm−1, which corresponds to the scissoring vibration of the –OH group, as seen in Figure S4b. The peak locations altered in the PbI2/DEA sample from 1052–1458 cm−1 to 1044–1448 cm−1, indicating a connection between PbI2 and the –OH groups of DEA. As seen in Figure S3b, the addition of PbI2 to DEA has significantly altered the appearance, perhaps creating new compounds. These observations point to a significant interaction between DEA and PbI2. In perovskite deposition, the exposed –OH groups on the ITO surface would interact favorably with the PbI2, which is present in the perovskite precursor solution. An even and complete coverage of the perovskite layer would benefit from this advantageous interaction, ensuring good interface contact with the ITO/DEA substrate.
The formation of high-quality perovskite layers may improve the performance of corresponding PSCs by facilitating charge transport and reducing recombination losses [42]. In addition to morphological characteristics, X-ray diffraction (XRD) and UV-vis absorption spectra were used to study how DEA deposition affected the structure and optical performance of the perovskite layers. The high-intensity peaks from the XRD pattern of the perovskite layer formed on ITO/DEA displayed slightly greater crystallinity compared to bare ITO, as shown in Figure 5a. Additionally, the perovskite layer formed on the ITO/DEA substrate exhibits slightly greater absorption capabilities throughout the entire spectral region as evidenced by UV-vis absorbance spectra in Figure 5b. Steady-state photoluminescence (PL) and time-resolved PL (TRPL) measurements were used to analyze the charge-carrier dynamics of perovskite layers grown on glass, bare ITO, and ITO/DEA substrates, as shown in Figure 5c,d. When compared to perovskite layers formed on glass and glass/ITO, the PL spectra in Figure 5c show a considerable drop in intensity for the ITO/DEA substrate, suggesting higher electron transfer to the corresponding layer [43] and lower recombination of charges [44]. A significant quenching effect for the perovskite layer deposited on the ITO/DEA substrate can be seen in the TRPL spectra given in Figure 5d, demonstrating the improved electron extraction capability from the perovskite layer to the DEA-treated substrate. The results were fitted by a biexponential decay function with the following equation.
Y = k 1 e x τ 1 + k 2 e x τ 2
where τ1 and τ2 represent decay factors that are closely related to non-radiative recombination by defects and a component of bulk recombination, respectively. Table S1 shows decreased values for τ1 and τ2 for the perovskite layer deposited on the ITO/DEA substrate, indicating minimized recombination losses.
We fabricated PSCs with and without DEA for the evaluation of photovoltaic parameters after carefully examining the morphology and substantial role of the DEA at the ITO/perovskite interface. The schematic diagram of the as-prepared device and energy levels of each layer are depicted in Figure S5. The cross-sectional SEM images, current-voltage (J-V) characteristic curves, and photovoltaic parameters of the as-prepared PSCs are depicted in Figure 6 and Table 1, respectively. The PSC employing bare ITO displays an open-circuit voltage (Voc) of 1.096 V, a short-circuit current density (Jsc) of 22.94 mA cm−2, a fill factor (FF) of 74.19%, and a power conversion efficiency (PCE) of 18.65% under reverse scan condition. In comparison, the DEA-modified PSC shows improved performance of up to PCE of 20.77%, a Voc of 1.109 V, a Jsc of 24.45 mA cm−2, and an FF of 76.60% under the same scan condition. It is noteworthy that the DEA-modified PSC exhibits significantly better performance and less hysteresis in the J-V curves. The better layer quality and improved charge-carrier transport at the ITO/DEA/perovskite interface might be primarily responsible for the improved performance and decreased hysteresis, which is consistent with the dark J-V curves, as discussed below, where we observed a decrease in the density of trap states in the DEA-modified PSCs.
To further confirm the charge-carrier dynamics and charge-carrier recombination in the devices with bare ITO and ITO/DEA, the dark J-V measurement was carried out. As seen in Figure 7a, the device with DEA has a lower current density than the device with bare ITO. This shows that the DEA/perovskite interface has efficient charge transport. In order to quantitatively assess the trap density in the as-prepared devices, we also fabricated electron-only devices (ITO/Perovskite/PCBM/Au and ITO/DEA/perovskite/PCBM/Au), as shown in Figure 7b. The trap density (Ntrap) can be determined with the equation
N t r a p = 2 ε 0 ε V T F L e L 2  
where ε 0 , ε , VTFL, e, and L are the vacuum permittivity, relative dielectric constant, trap-field limit voltage, elementary charge, and perovskite layer thickness, respectively. Since Ntrap and VTFL are linearly related, a device with a smaller VTFL will exhibit fewer defects overall. As seen in Figure 7b, the device with DEA exhibits a significantly lower VTFL (0.25 V) than the device with bare ITO (0.33 V), indicating DEA successfully reduces the defect density. The smooth morphology with fewer grain boundaries and the high-quality crystallinity of the perovskite layer deposited on the ITO/DEA substrate are responsible for the reduced trap density. Electrochemical impedance spectroscopy (EIS) measurements were conducted in order to learn more about how DEA affects the interfacial charge transport and recombination processes in the PSCs. Figure 7c displays the Nyquist plots of the corresponding devices with a bias voltage of 0.9 V under dark conditions and a frequency range of 1 Hz to 1 MHz. At higher frequencies, the semicircle stands in for the charge transport resistance (Rct), which originates from the substrate and perovskite contact. Because ITO/DEA made better contact with the perovskite surface, the DEA-treated PSC showed a lower Rct value (29.76 kΩ) than the ITO-based device (Rct of 54.06 kΩ), indicating more efficient electron transfer and electron extraction at the corresponding interface. Additionally, measurements of external quantum efficiency (EQE) were made in order to assess how well the J-V and EQE curves agreed. The matching Jsc values integrated from the EQE curves (Figure 7d) for both devices are consistent with J-V characteristic curves. Here, we noticed a minor discrepancy between the EQEs of the PSCs and the absorption spectra of the perovskite layers prepared on ITO and DEA-treated ITO substrates in the range of 400–500 nm. This can be attributed to the fact that the absorption spectra do not represent an external property of the PSCs but rather an internal property of the perovskites (originating from the distinctive energy-band structure) [45,46,47]. Moreover, the EQE spectrum encompasses all excitonic and/or thermal broadening effects, along with the architectural features of the PSC, and this reflects the occupied density of states as specified in the entire PSC rather than in individual perovskites.
Furthermore, we acquired performance data from 9 PSCs with bare ITO and ITO/DEA, respectively, to show the reproducibility of the devices. The ITO/DEA-based PSCs exhibit improved performance and reproducibility when compared to PSCs with bare ITO, as shown in Figure S6. In addition to the high PCEs, the PSCs should have long-term stability. Figure 8a illustrates the tracking of the steady-state PCEs of the PSCs with ITO/DEA and bare ITO under AM 1.5 G at maximum power at 0.9 V bias voltage and ambient temperature. After 120 s of continuous illumination, the ITO/DEA-based device showed a stabilized PCE of 19.71%, which is considerably near the highest PCE of 20.77% for the same PSC. Additionally, for 550 h of AM 1.5 G irradiation, the stabilities of the unsealed PSCs made with bare ITO and ITO/DEA substrates were evaluated every 34 h. After each stability test, the tested PSCs were kept in a glass oven at 60% humidity and 25 °C. Figure 8b demonstrates how the stability of the ITO/DEA-based PSCs exceeded that of the ITO-based devices. After 550 h, the ITO/DEA-based PSC still had 80% of the original PCE. The bare ITO-based PSCs under the same testing conditions only retained 37% of the initial PCE, demonstrating poor stability and quick degradation. The degradation of ITO-based PSCs is associated with rapid erosion because the perovskite layer formed on the bare ITO substrate has a defective surface morphology. As mentioned above, high-quality perovskite layers were created on the DEA-treated ITO substrates, with compact, large grains, and reduced grain boundaries. Starting from the grain boundaries makes it considerably easier to decompose perovskite films. Less oxygen, water, etc. will intermix within the grain boundaries of the perovskite films because there are fewer grain boundaries. The stability of PSCs made using DEA-treated ITO substrates is ultimately improved by the perovskite film with fewer grain boundaries, which effectively shields perovskite films from water penetration. Conversely, erosion occurs more rapidly in the perovskite thin films generated using the ITO substrates because these have a considerably rougher surface with smaller grain sizes and numerous grain boundaries.

4. Conclusions

We used DEA to generate Sn-N and –OH–Pb chemical linkages between the perovskite layer and the ITO substrate. The interface contact, perovskite crystallinity, and film quality have all increased due to these chemical interactions, and the perovskite layer’s grain boundaries and trap states have decreased. Based on the creation of interfacial dipole layers, our results imply that the addition of DEA can modify the band misfit at the ITO/perovskite contact, hence reducing the voltage deficits. Transient PL, TRPL, and IS measurements revealed that the rates of electron extraction and charge transfer have dramatically improved. The PCE of the PSCs has significantly increased as a result, rising from about 18.65% to 20.77%. Additionally, DEA modification has increased the stability of the PSCs, with 80% PCE retention even after 550 h without encapsulation. This is likely because the DEA modification improved the interfacial contact and thin-film quality of the perovskite. Our research opens up scientific pathways for the facile fabrication of ESL-free PSCs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13020250/s1, Figure S1: Transmittance spectra of bare ITO and DEA-treated ITO substrates; Figure S2: KPFM images (a,b) and corresponding potential difference curves (c,d) of the bare ITO and DEA-modified ITO substrates; Figure S3: An strengthened link at the ITO/perovskite interface is highlighted by a coupling between the –OH group and Pb in the perovskite in this schematic example of surface modification of ITO-substrate with DEA (a), Photos of DEA liquid, PbI2 powder and PbI2/DEA mixture (b), Photos of DEA liquid, PbI2 powder and PbI2/DEA mixture under 135 °C (c); Figure S4: FT-IR spectroscopy of the ITO and ITO/DEA substrates (a), and DEA liquid and PbI2/DEA mixture (b); Figure S5: Schematic illustration of as-prepared PSC (a) and energy band diagrams of the corresponding layers (b); Figure S6: Statistical analyses of the ESLs-free PSCs prepared with bare ITO and DEA-modified ITO substrates, Voc (a), Jsc (b), FF (c), and PCE (d); Table S1: TRPL data for the perovskite layers prepared on bare ITO and DEA-modified substrates.

Author Contributions

Conceptualization, S.S., I.M.O. and S.A.; Supervision and validation, I.M.O., S.A. and I.B.S.; Formal analysis, D.W.; J.P.; I.B.S. and I.M.O.; Investigation, S.S.; Methodology, S.S.; Writing, S.S.; Review & editing, I.M.O. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the UAEU-Strategic research program under Grant no. 12R128.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data presented in the manuscript can be obtained from the corresponding authors by reasonable request.

Acknowledgments

The authors acknowledge the UAEU-Strategic research program for providing financial support.

Conflicts of Interest

The authors do not have any conflicts of interest to disclose.

References

  1. Green, M.; Dunlop, E.; Hohl-Ebinger, J.; Yoshita, M.; Kopidakis, N.; Bothe, K.; Hinken, D.; Rauer, M.; Hao, X. Solar cell efficiency tables (Version 60). Prog. Photovolt. Res. Appl. 2022, 30, 687–701. [Google Scholar] [CrossRef]
  2. Sajid, S.; Elseman, A.; Huang, H.; Ji, J.; Dou, S.; Jiang, H.; Liu, X.; Wei, D.; Cui, P.; Li, M. Breakthroughs in NiOx-HTMs towards stable, low-cost and efficient perovskite solar cells. Nano Energy 2018, 51, 408–424. [Google Scholar] [CrossRef]
  3. Sajid, S.; Elseman, A.; Ji, J.; Dou, S.; Wei, D.; Huang, H.; Cui, P.; Xi, W.; Chu, L.; Li, Y.; et al. Computational Study of Ternary Devices: Stable, Low-Cost, and Efficient Planar Perovskite Solar Cells. Nano-Micro Lett. 2018, 10, 51. [Google Scholar] [CrossRef] [Green Version]
  4. Sajid, S.; Khan, S.; Khan, A.; Khan, D.; Issakhov, A.; Park, J. Antisolvent-fumigated grain growth of active layer for efficient perovskite solar cells. Sol. Energy 2021, 225, 1001–1008. [Google Scholar] [CrossRef]
  5. Sajid, S.; Elseman, A.; Wei, D.; Ji, J.; Dou, S.; Huang, H.; Cui, P.; Li, M. NiO@carbon spheres: A promising composite electrode for scalable fabrication of planar perovskite solar cells at low cost. Nano Energy 2019, 55, 470–476. [Google Scholar] [CrossRef]
  6. Sajid, S.; Alzahmi, S.; Salem, I.; Obaidat, I. Guidelines for Fabricating Highly Efficient Perovskite Solar Cells with Cu2O as the Hole Transport Material. Nanomaterials 2022, 12, 3315. [Google Scholar] [CrossRef]
  7. Im, J.-H.; Lee, C.-R.; Lee, J.-W.; Park, S.-W.; Park, N.-G. 6.5% efficient perovskite quantum-dot-sensitized solar cell. Nanoscale 2011, 3, 4088–4093. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Elseman, A.; Shalan, A.; Sajid, S.; Rashad, M.; Hassan, A.; Li, M. Copper-Substituted Lead Perovskite Materials Constructed with Different Halides for Working (CH3NH3)2CuX4-Based Perovskite Solar Cells from Experimental and Theoretical View. ACS Appl. Mater. Interfaces 2018, 10, 14. [Google Scholar] [CrossRef] [PubMed]
  9. Sajid, S.; Elseman, A.; Ji, J.; Dou, S.; Huang, H.; Cui, P.; Wei, D.; Li, M. Novel hole transport layer of nickel oxide composite with carbon for high-performance perovskite solar cells. Chin. Phys. B 2018, 27, 17305. [Google Scholar] [CrossRef]
  10. Wei, D.; Ma, F.; Wang, R.; Dou, S.; Cui, P.; Huang, H.; Ji, J.; Jia, E.; Jia, X.; Sajid, S.; et al. Ion-Migration Inhibition by the Cation–π Interaction in Perovskite Materials for Efficient and Stable Perovskite Solar Cells. Adv. Mater. 2018, 30, 1707583. [Google Scholar] [CrossRef]
  11. Elseman, A.; Sharmoukh, W.; Sajid, S.; Cui, P.; Ji, J.; Dou, S.; Wei, D.; Huang, H.; Xi, W.; Chu, L.; et al. Superior Stability and Efficiency over 20% Perovskite Solar Cells Achieved by a Novel Molecularly Engineered Rutin–AgNPs/Thiophene Copolymer. Adv. Sci. 2018, 5, 1800568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Sajid, S.; Huang, H.; Ji, J.; Jiang, H.; Duan, M.; Liu, X.; Liu, B.; Li, M. Quest for robust electron transporting materials towards efficient, hysteresis-free and stable perovskite solar cells. Renew. Sustain. Energy Rev. 2021, 152, 111689. [Google Scholar] [CrossRef]
  13. Khan, D.; Sajid, S.; Khan, S.; Park, J.; Ullah, I. Identifying the potentials for charge transport layers free n-p homojunction-based perovskite solar cells. Sol. Energy 2022, 238, 69–77. [Google Scholar] [CrossRef]
  14. Elseman, A.; Sajid, S.; Shalan, A.; Mohamed, S.; Rashad, M. Recent progress concerning inorganic holetransport layers for efficient perovskite solar cells. Appl. Phys. A 2019, 125, 476. [Google Scholar] [CrossRef]
  15. Wei, D.; Huang, H.; Cui, P.; Ji, J.; Dou, S.; Jia, E.; Sajid, S.; Cui, M.; Chu, L.; Li, Y.; et al. Moisture-tolerant supermolecule for the stability enhancement of organic–inorganic perovskite solar cells in ambient air. Nanoscale 2019, 11, 1228–1235. [Google Scholar] [CrossRef] [PubMed]
  16. Huang, H.; Yan, H.; Duan, M.; Ji, J.; Liu, X.; Jiang, H.; Liu, B.; Sajid, S.; Cui, P.; Li, Y.; et al. TiO2 surface oxygen vacancy passivation towards mitigated interfacial lattice distortion and efficient perovskite solar cell. Appl. Surf. Sci. 2021, 544, 148583. [Google Scholar] [CrossRef]
  17. Huang, H.; Liu, X.; Duan, M.; Ji, J.; Jiang, H.; Liu, B.; Sajid, S.; Cui, P.; Wei, D.; Li, Y.; et al. Dual Function of Surface Alkali-Gas Erosion on SnO2 for Efficient and Stable Perovskite Solar Cells. ACS Appl. Energy Mater. 2020, 3, 5039–5049. [Google Scholar] [CrossRef]
  18. Stranks, S.; Eperon, G.; Grancini, G.; Menelaou, C.; Alcocer, M.; Leijtens, T.; Herz, L.; Petrozza, A.; Snaith, H. Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber. Science 2013, 342, 341–344. [Google Scholar] [CrossRef] [Green Version]
  19. Dong, Q.; Fang, Y.; Shao, Y.; Mulligan, P.; Qiu, J.; Cao, L.; Huang, J. Electron-hole diffusion lengths >175 μm in solution-grown CH3NH3PbI3 single crystals. Science 2015, 347, 967–970. [Google Scholar] [CrossRef] [Green Version]
  20. Wu, W.-Q.; Wang, Q.; Fang, Y.; Shao, Y.; Tang, S.; Deng, Y.; Lu, H.; Liu, Y.; Li, T.; Yang, Z. Molecular doping enabled scalable blading of efficient hole-transport-layer-free perovskite solar cells. Nat. Commun. 2018, 9, 1–8. [Google Scholar] [CrossRef]
  21. Zhang, J.; Sun, Y.; Yu, H. Reducing energy loss via adjusting the anode work function and perovskite layer morphology for the efficient and stable hole transporting layer-free perovskite solar cells. Chem. Eng. J. 2022, 431, 133948. [Google Scholar] [CrossRef]
  22. Rong, Y.; Hu, Y.; Mei, A.; Tan, H.; Saidaminov, M.; Seok, S.I.; McGehee, M.; Sargent, E.; Han, H. Challenges for commercializing perovskite solar cells. Science 2018, 361, eaat8235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Huang, L.; Ge, Z. Simple, Robust, and Going More Efficient: Recent Advance on Electron Transport Layer-Free Perovskite Solar Cells. Adv. Energy Mater. 2019, 9, 1900248. [Google Scholar] [CrossRef]
  24. Cui, P.; Wei, D.; Ji, J.; Song, D.; Li, Y.; Liu, X.; Huang, J.; Wang, T.; You, J.; Li, M. Highly efficient electron-selective layer free perovskite solar cells by constructing effective p–n heterojunction. Sol. RRL 2017, 1, 1600027. [Google Scholar] [CrossRef]
  25. Zhao, P.; Han, M.; Yin, W.; Zhao, X.; Kim, S.; Yan, Y.; Kim, M.; Song, Y.; Park, N.; Jung, H. Insulated Interlayer for Efficient and Photostable Electron-Transport-Layer-Free Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2018, 10, 10132–10140. [Google Scholar] [CrossRef] [PubMed]
  26. Zheng, E.; Wang, X.-F.; Song, J.; Yan, L.; Tian, W.; Miyasaka, T. PbI2-Based Dipping-Controlled Material Conversion for Compact Layer Free Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2015, 7, 18156–18162. [Google Scholar] [CrossRef]
  27. Liao, J.-F.; Wu, W.-Q.; Jiang, Y.; Zhong, J.-X.; Wang, L.; Kuang, D.-B. Understanding of carrier dynamics, heterojunction merits and device physics: Towards designing efficient carrier transport layer-free perovskite solar cells. Chem. Soc. Rev. 2020, 49, 354–381. [Google Scholar] [CrossRef]
  28. Wu, W.-Q.; Liao, J.-F.; Zhong, J.-X.; Xu, Y.-F.; Wang, L.; Huang, J. Suppressing Interfacial Charge Recombination in Electron-Transport-Layer-Free Perovskite Solar Cells to Give an Efficiency Exceeding 21%. Angew. Chem. Int. Ed. 2020, 59, 20980–20987. [Google Scholar] [CrossRef]
  29. Li, D.; Chao, L.; Chen, C.; Ran, X.; Wang, Y.; Niu, T.; Lv, S.; Wu, H.; Xia, Y.; Ran, C.; et al. In Situ Interface Engineering for Highly Efficient Electron-Transport-Layer-Free Perovskite Solar Cells. Nano Lett. 2020, 20, 5799–5806. [Google Scholar] [CrossRef]
  30. Huang, L.; Zhang, D.; Bu, S.; Peng, R.; Wei, Q.; Ge, Z. Synergistic Interface Energy Band Alignment Optimization and Defect Passivation toward Efficient and Simple-Structured Perovskite Solar Cell. Adv. Sci. 2020, 7, 1902656. [Google Scholar] [CrossRef]
  31. Cheng, H.; Li, Y.; Zhang, M.; Zhao, K.; Wang, Z.-S. Self-Assembled Ionic Liquid for Highly Efficient Electron Transport Layer-Free Perovskite Solar Cells. ChemSusChem 2020, 13, 2779–2785. [Google Scholar] [CrossRef]
  32. Valles-Pelarda, M.; Hames, B.; García-Benito, I.; Almora, O.; Molina-Ontoria, A.; Sánchez, R.; Garcia-Belmonte, G.; Martín, N.; Mora-Sero, I. Analysis of the hysteresis behavior of perovskite solar cells with interfacial fullerene self-assembled monolayers. J. Phys. Chem. Lett. 2016, 7, 4622–4628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Sadegh, F.; Akman, E.; Prochowicz, D.; Tavakoli, M.; Yadav, P.; Akin, S. Facile NaF Treatment Achieves 20% Efficient ETL-Free Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2022, 14, 34. [Google Scholar] [CrossRef] [PubMed]
  34. Huang, C.; Lin, P.; Fu, N.; Liu, C.; Xu, B.; Sun, K.; Wang, D.; Zeng, X.; Ke, S. Facile fabrication of highly efficient ETL-free perovskite solar cells with 20% efficiency by defect passivation and interface engineering. Chem. Commun. 2019, 55, 2777–2780. [Google Scholar] [CrossRef] [PubMed]
  35. Bai, Y.; Chen, H.; Xiao, S.; Xue, Q.; Zhang, T.; Zhu, Z.; Li, Q.; Hu, C.; Yang, Y.; Hu, Z. Effects of a molecular monolayer modification of NiO nanocrystal layer surfaces on perovskite crystallization and interface contact toward faster hole extraction and higher photovoltaic performance. Adv. Funct. Mater. 2016, 26, 2950–2958. [Google Scholar] [CrossRef]
  36. Hu, W.; Zhou, W.; Lei, X.; Zhou, P.; Zhang, M.; Chen, T.; Zeng, H.; Zhu, J.; Dai, S.; Yang, S. Low-temperature in situ amino functionalization of TiO2 nanoparticles sharpens electron management achieving over 21% efficient planar perovskite solar cells. Adv. Mater. 2019, 31, 1806095. [Google Scholar] [CrossRef] [PubMed]
  37. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-light photocatalysis in nitrogen-doped titanium oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef]
  38. Chen, J.; Zhao, X.; Kim, S.; Park, N. Multifunctional chemical linker imidazoleacetic acid hydrochloride for 21% efficient and stable planar perovskite solar cells. Adv. Mater. 2019, 31, 1902902. [Google Scholar] [CrossRef] [PubMed]
  39. Zheng, Y.; Jiang, B.; Gao, Z.; Lin, G.; Sang, N.; Chen, L.; Li, M. Optimization of SnO2-based electron-selective contacts for Si/PEDOT: PSS heterojunction solar cells. Sol. Energy 2019, 193, 502–506. [Google Scholar] [CrossRef]
  40. Peng, Q.; Guo, J.; Zhang, Q.; Xiang, J.; Liu, B.; Zhou, A.; Liu, R.; Tian, Y. Unique lead adsorption behavior of activated hydroxyl group in two-dimensional titanium carbide. J. Am. Chem. Soc. 2014, 136, 4113–4116. [Google Scholar] [CrossRef]
  41. Thirumoorthi, M.; Prakash, J.T.J. Structure, optical and electrical properties of indium tin oxide ultra thin films prepared by jet nebulizer spray pyrolysis technique. J. Asian Ceram. Soc. 2016, 4, 124–132. [Google Scholar] [CrossRef] [Green Version]
  42. Sajid, S.; Alzahmi, S.; Salem, I.; Obaidat, I. Perovskite-Surface-Confined Grain Growth for High-Performance Perovskite Solar Cells. Nanomaterials 2022, 12, 3352. [Google Scholar] [CrossRef] [PubMed]
  43. Akman, E.; Akin, S. Poly(N,N′-bis-4-butylphenyl-N,N′-bisphenyl)benzidine-Based Interfacial Passivation Strategy Promoting Efficiency and Operational Stability of Perovskite Solar Cells in Regular Architecture. Adv. Mater. 2021, 33, 2006087. [Google Scholar] [CrossRef] [PubMed]
  44. Krogmeier, B.; Staub, F.; Grabowski, D.; Rau, U.; Kirchartz, T. Quantitative analysis of the transient photoluminescence of CH3NH3PbI3/PC61BM heterojunctions by numerical simulations. Sustain. Energy Fuels 2018, 2, 1027–1034. [Google Scholar] [CrossRef]
  45. Chen, X.; Lu, H.; Yang, Y.; Beard, M. Excitonic effects in methylammonium lead halide perovskites. J. Phys. Chem. Lett. 2018, 9, 2595–2603. [Google Scholar] [CrossRef] [PubMed]
  46. Wang, T.; Daiber, B.; Frost, J.; Mann, S.; Garnett, E.; Walsh, A.; Ehrler, B. Indirect to direct bandgap transition in methylammonium lead halide perovskite. Energy Environ. Sci. 2017, 10, 509–515. [Google Scholar] [CrossRef] [Green Version]
  47. Kirchartz, T.; Rau, U. Decreasing radiative recombination coefficients via an indirect band gap in lead halide perovskites. J. Phys. Chem. Lett. 2017, 8, 1265–1271. [Google Scholar] [CrossRef]
Figure 1. XPS spectra of ITO substrate and ITO/DEA substrate (a), N 1s spectra of ITO substrate and ITO/DEA substrate (b), analysis of N 1s spectra of ITO/DEA substrate (c), and Sn 3d spectra of ITO substrate and ITO/DEA substrate (d).
Figure 1. XPS spectra of ITO substrate and ITO/DEA substrate (a), N 1s spectra of ITO substrate and ITO/DEA substrate (b), analysis of N 1s spectra of ITO/DEA substrate (c), and Sn 3d spectra of ITO substrate and ITO/DEA substrate (d).
Nanomaterials 13 00250 g001
Figure 2. SEM images of the ITO substrate (a), ITO/DEA substrate (b), AFM images of ITO substrate (c), and ITO/DEA substrate (d).
Figure 2. SEM images of the ITO substrate (a), ITO/DEA substrate (b), AFM images of ITO substrate (c), and ITO/DEA substrate (d).
Nanomaterials 13 00250 g002
Figure 3. UPS spectra of ITO substrate and ITO/DEA substrate (a,b), and the contact angle photographs of water droplets on ITO substrate and ITO/DEA substrate (c).
Figure 3. UPS spectra of ITO substrate and ITO/DEA substrate (a,b), and the contact angle photographs of water droplets on ITO substrate and ITO/DEA substrate (c).
Nanomaterials 13 00250 g003
Figure 4. SEM images of the surface and cross-section of the perovskite (Psk.) layers prepared on ITO substrates (ac) and ITO/DEA substrates (df), respectively.
Figure 4. SEM images of the surface and cross-section of the perovskite (Psk.) layers prepared on ITO substrates (ac) and ITO/DEA substrates (df), respectively.
Nanomaterials 13 00250 g004
Figure 5. XRD patterns (a), UV-Vis spectra (b), PL-spectra (c), and TRPL-spectra (d) of the ITO/perovskite and ITO/DEA/perovskite, respectively.
Figure 5. XRD patterns (a), UV-Vis spectra (b), PL-spectra (c), and TRPL-spectra (d) of the ITO/perovskite and ITO/DEA/perovskite, respectively.
Nanomaterials 13 00250 g005
Figure 6. Cross-sectional SEM images of as-fabricated devices with ITO substrate (a) and ITO/DEA substrate (b). The J-V characteristic curves of devices fabricated with ITO substrate (c) and ITO/DEA substrate (d).
Figure 6. Cross-sectional SEM images of as-fabricated devices with ITO substrate (a) and ITO/DEA substrate (b). The J-V characteristic curves of devices fabricated with ITO substrate (c) and ITO/DEA substrate (d).
Nanomaterials 13 00250 g006
Figure 7. Dark J-V curves of as-prepared devices (a) and electron-only devices (b), EIS measurement (c), and EQEs with integrated Jsc of the PSCs prepared using ITO substrates and ITO/DEA substrates (d).
Figure 7. Dark J-V curves of as-prepared devices (a) and electron-only devices (b), EIS measurement (c), and EQEs with integrated Jsc of the PSCs prepared using ITO substrates and ITO/DEA substrates (d).
Nanomaterials 13 00250 g007
Figure 8. Stability analyses of the PSCs prepared with ITO substrates and ITO/DEA substrates. PCE at maximum power point tracking under continuous illumination of AM 1.5 G irradiation (a) and Normalized PCE for 550 h under AM 1.5 G irradiation (b).
Figure 8. Stability analyses of the PSCs prepared with ITO substrates and ITO/DEA substrates. PCE at maximum power point tracking under continuous illumination of AM 1.5 G irradiation (a) and Normalized PCE for 550 h under AM 1.5 G irradiation (b).
Nanomaterials 13 00250 g008
Table 1. Photovoltaic parameters of the PSCs under forward bias and reverse bias.
Table 1. Photovoltaic parameters of the PSCs under forward bias and reverse bias.
DeviceScanVoc (V)Jsc (mA cm−2)FF (%)PCE (%)
ITO/Psk./spiro/AuReverse1.09622.9474.1918.65
Forward1.08320.8374.2416.75
ITO/DEA/Psk./spiro/AuReverse1.10924.4576.6020.77
Forward1.09624.0277.9420.52
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

Sajid, S.; Alzahmi, S.; Wei, D.; Salem, I.B.; Park, J.; Obaidat, I.M. Diethanolamine Modified Perovskite-Substrate Interface for Realizing Efficient ESL-Free PSCs. Nanomaterials 2023, 13, 250. https://doi.org/10.3390/nano13020250

AMA Style

Sajid S, Alzahmi S, Wei D, Salem IB, Park J, Obaidat IM. Diethanolamine Modified Perovskite-Substrate Interface for Realizing Efficient ESL-Free PSCs. Nanomaterials. 2023; 13(2):250. https://doi.org/10.3390/nano13020250

Chicago/Turabian Style

Sajid, Sajid, Salem Alzahmi, Dong Wei, Imen Ben Salem, Jongee Park, and Ihab M. Obaidat. 2023. "Diethanolamine Modified Perovskite-Substrate Interface for Realizing Efficient ESL-Free PSCs" Nanomaterials 13, no. 2: 250. https://doi.org/10.3390/nano13020250

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