Robiul
Islam†
ab,
Rahim
Abdur†
b,
Md. Ashraful
Alam
a,
Nadim
Munna
c,
Aninda Nafis
Ahmed
d,
Mosharof
Hossain
b,
Mohammad Shahriar
Bashar
b,
Dipa
Islam
e and
Mohammad Shah
Jamal
*b
aDepartment of Applied Chemistry and Chemical Engineering, Noakhali Science and Technology University, Noakhali 3814, Bangladesh
bInstitute of Energy Research and Development (IERD), Bangladesh Council of Scientific and Industrial Research (BCSIR), Dr Qudrat-E-Khuda Road, Dhanmondi, Dhaka 1205, Bangladesh. E-mail: msjdubd@gmail.com; msjamal@bcsir.gov.bd
cInstitute of Mining, Mineralogy and Metallurgy (IMMM), Bangladesh Council of Scientific and Industrial Research (BCSIR), Science Laboratory Road, Khanjanpur, Joypurhat 5900, Bangladesh
dPilot Plant and Process Development Centre (PP&PDC), Bangladesh Council of Scientific and Industrial Research (BCSIR), Dr Qudrat-E-Khuda Road, Dhanmondi, Dhaka 1205, Bangladesh
eBiomedical and Toxicology Research Institute (BTRI), Bangladesh Council of Scientific and Industrial Research (BCSIR), Dr Qudrat-E-Khuda Road, Dhanmondi, Dhaka 1205, Bangladesh
First published on 31st October 2024
Mn-doped NiO nanoparticles (NPs), denoted as Ni1−xMnxO with x values of 0.00, 0.02, 0.04, 0.06, and 0.08, were synthesized using a chemical precipitation process. These NPs were comprehensively analyzed for their structural, optical, and electrical properties, along with their surface morphology and elemental composition. X-ray Diffraction (XRD) confirmed the single-phase cubic crystal structure and revealed a reduction in crystallite size from 15.26 nm to 10.38 nm as Mn doping increased. Field Emission Scanning Electron microscopy (FE-SEM) determined the average particle sizes ranging from 26.03 nm to 23.30 nm. The optical properties, assessed by UV-visible spectroscopy (UV-vis), revealed a widening of the bandgap from 3.49 eV to 4.10 eV with increasing Mn doping, suggesting tunable optical characteristics. X-ray Photoelectron Spectroscopy (XPS) confirmed the presence of nickel (Ni), oxygen (O), and manganese (Mn) within the NPs. The highest mobility, 1.31 ± 0.03 × 103 cm2 V−1 s−1, was observed in the 6 wt% Mn-doped NiO NPs thin film, as determined by Hall measurements. To assess their practical utility, SCAPS-1D simulation was employed, demonstrating the potential of Mn-doped NiO NPs as a hole transport layer (HTL) in perovskite solar cells (PSCs). The enhanced electrical and optical properties, combined with structural tunability, highlight Mn-doped NiO as a promising material for advanced optoelectronic applications. This study provides valuable insights into the development of efficient and stable solar cells, offering a pathway to optimize material design for improved performance in photovoltaic applications.
Depending on the particle size, shape, and synthesis process, nickel oxide (NiO) nanoparticles (NPs) are highly promising for various electronic and optoelectronic applications due to their wide bandgap and p-type semiconducting properties. In its nanoparticle form, NiO demonstrates enhanced semiconducting behavior, whereas bulk NiO is typically an insulator.8 This transition from an insulating state to a semiconducting one highlights the importance of tuning nanoparticle synthesis to achieve desired functionalities. The characteristics of NiO can be improved by using the right metal dopant that can give an adequate boost to the optical, electrical, and physical properties of NiO.9 Several research groups have attempted to improve the main features of NiO by adding dopants such as copper (Cu),10 iron (Fe),11 cobalt (Co),9 manganese (Mn),12 lithium (Li),13 and aluminum (Al).14 Doping NiO with transition metals such as manganese (Mn) has been shown to dramatically improve its electrical, optical, and structural properties. Mn is a particularly effective dopant due to its similar ionic radius to Ni [Mn4+ (0.53 Å) and Ni2+ (0.69 Å)], allowing for easy substitution into the NiO lattice, resulting in significant enhancements in performance metrics. As a result, incorporating Mn into the NiO lattice may easily replace Ni sites while improving its structural, electrical, and optical characteristics.15 There are several methods to synthesize Mn-doped NiO NPs, such as sol–gel,16 wet chemical,17 homogeneous precipitation,18 pyrolysis,19 and hydrothermal methods.20,21 While each of these techniques has its merits, they are often constrained by complexity, high energy demands, and limited scalability. For instance, the sol–gel process, although capable of producing well-controlled particle sizes, typically requires complex setups and long processing times, which hinder its applicability for large-scale production. Similarly, hydrothermal methods, known for producing high-purity nanoparticles, involve high-pressure systems and prolonged heating cycles, making them impractical for cost-effective manufacturing. In contrast, the chemical precipitation method presents a significant advancement by offering simplicity and scalability, without sacrificing the quality of the nanoparticles produced.
Moreover, the hydrothermal method is one of the simplest techniques to synthesize Mn-doped NiO NPs providing homogeneous mixing, better crystallinity, uniform particle size dissemination, and smaller particle size with high purity. As a member of the metal oxide family, NiO is a p-type material with a wide bandgap and has extraordinary structural, and optical properties, with excellent chemical stability.22,23 The amazing unique properties of NiO have made this compound a promising candidate for photovoltaic applications,24–26 gas sensors,27 UV photodetectors,28 and electrochromic devices.29 Recently scientists have been looking at NiO NPs a lot to make perovskite solar cells (PSCs) work better and last longer. People have tried putting these NPs into PSCs as a hole-transporting material (HTM) to help move charges, improve contact between layers, line up energy levels, and cut down on energy loss and recombination.30–32 Strategies such as pre-doping NiO films with silver (Ag) ions to form a p/p+ homojunction have shown significant improvements in charge separation, energy level alignment, and overall efficiency of PSCs.33 NiO is tough, moves holes well, and doesn't cost much, so it could be great for the hole transport layer (HTL) in PSCs. But making it and treating it after still cause some headaches.34 Overall, the incorporation of Mn-doped NiO nanoparticles into PSCs presents a promising strategy for significantly enhancing device performance. The tailored structural, optical, and electrical properties of these doped nanoparticles improve hole transport efficiency, which could lead to better stability and higher power conversion efficiencies. This advancement not only holds potential for improving the commercial viability of PSCs but also paves the way for broader market adoption.
In this study, we have synthesized undoped NiO and Mn-doped NiO NPs by a simple chemical precipitation process. The concentrations of Mn-dopant solution were varied in a wide range of 0, 2, 4, 6, and 8 weight percent (wt%) of NiO precursor solution. This wide range of doping concentrations were carefully controlled, allowing fine-tuning of the structural, optical, and electrical properties of the NiO NPs. Such precision in doping control is critical for optimizing the performance of these materials in different device applications. The prepared samples' structural, morphological, optical, elemental composition and valence state and electrical characteristics have been investigated to understand the Mn doping effect on NiO NPs. Also, the optical and electrical properties obtained from the investigation was used in SCAPS-1D simulation to simulate NiO NPs based thin film HTL layer in PSCs. In this case an inverted structure of PSCs with Glass\ITO\NiO\MAPbI3\PCBM\Al layers.
The key innovation in this work lies in the method's ability to overcome the traditional challenges associated with NiO synthesis. By employing a straightforward chemical precipitation process, we have employed a technique that is not only more sustainable and scalable but also produces nanoparticles with enhanced functional properties tailored for advanced applications. In particular, Mn-doped NiO nanoparticles synthesized by this method exhibit superior hole transport properties, making them highly effective as a hole-transporting material (HTM) in PSCs. This improvement is critical because it leads to enhanced charge mobility, better energy alignment, and a reduction in energy loss and recombination, all of which contribute to significantly higher power conversion efficiencies (PCE) and improved stability of PSC devices. Previous research has demonstrated the potential of NiO NPs in enhancing the performance of PSCs, particularly as an HTM to facilitate charge transport and improve interfacial contact between active layers. Our Mn-doped NiO nanoparticles, synthesized via this low-temperature method, offer an even greater advantage by further improving the hole transport efficiency and tuning the electronic properties of the NiO layer. This results in better energy level alignment and more effective charge separation, leading to overall improved device performance. Additionally, the chemical stability, low cost, and ease of production of these doped NiO nanoparticles make them a highly attractive option for the commercial viability of PSCs.
Properties | PCBM (ETL)37,38 | Perovskite (MAPbI3)39 | NiO (HTL)27,28 | ||||
---|---|---|---|---|---|---|---|
a | b | c | d | e | |||
a a – Pure NiO, b – 2, c – 4, d – 6, and e – 8 wt% Mn doped NiO HTL layer. | |||||||
Thickness (nm) | 70 | 320 | 50 (exp.) | 50 (exp.) | 50 (exp.) | 50 (exp.) | 50 (exp.) |
Bandgap, Eg (eV) | 2.1 | 1.56 | 3.49 (exp.) | 3.69 (exp.) | 3.74 (exp.) | 3.79 (exp.) | 3.91 (exp.) |
Electron affinity, xe (eV) | 4.1 | 3.90 | 1.8 | 1.8 | 1.8 | 1.8 | 1.8 |
Dielectric permittivity, ∈r (relative) | 4 | 10 | 11.75 | 11.75 | 11.75 | 11.75 | 11.75 |
CB effective density of states, NC (cm−3) | 2.5 × 1021 | 2.76 × 1018 | 2.0 × 1018 | 2.0 × 1018 | 2.0 × 1018 | 2.0 × 1018 | 2.0 × 1018 |
VB effective density of states, NV (cm−3) | 2.5 × 1021 | 3.90 × 1018 | 2.0 × 1018 | 2.0 × 1018 | 2.0 × 1018 | 2.0 × 1018 | 2.0 × 1018 |
Electron thermal velocity (cm s−1) | 1 × 107 | 1 × 107 | 107 | 107 | 107 | 107 | 107 |
Hole thermal velocity | 1 × 107 | 1 × 107 | 107 | 107 | 107 | 107 | 107 |
Electron mobility (cm2 V−1 s−1) | 0.01 | 15 | 8 | 8 | 8 | 8 | 8 |
Hole mobility (cm2 V−1 s−1) | 0.01 | 15 | 211 (exp.) | 249 (exp.) | 469 (exp.) | 1310 (exp.) | 62.60 (exp.) |
Shallow uniform acceptor density, NA (cm−3) | 1 × 1011 | 6.32 × 10 10 (exp.) | 7.30 × 10 10 (exp.) | 1.16 × 10 11 (exp.) | 2.14 × 10 11 (exp.) | 3.89 × 10 11 (exp.) | |
Shallow uniform donor density, ND (cm−3) | 5 × 1017 | 1 × 1011 | — | — | — | — |
Sample Ni1−xMnxO (x =) | Structural properties derived from (200) plane | ||||||
---|---|---|---|---|---|---|---|
2θ (degree) | β (degree) | a (Å) | D 200 (nm) | ε × 10−3 (%) | δ × 10−3 (nm−2) | σ s (GPa) | |
0.00 | 43.32 | 0.59 | 4.17 | 15.26 | 6.48 | 4.37 | 0.40 |
0.02 | 43.33 | 0.77 | 4.17 | 11.58 | 8.46 | 7.44 | 0.45 |
0.04 | 43.39 | 0.78 | 4.17 | 11.45 | 8.55 | 7.63 | −0.07 |
0.06 | 43.41 | 0.80 | 4.17 | 11.14 | 8.79 | 8.06 | −0.19 |
0.08 | 43.36 | 0.86 | 4.17 | 10.38 | 9.44 | 9.28 | 0.26 |
The average crystallite size (Dhkl) was appraised by means of the Debye–Scherrer's equation:40
![]() | (1) |
The wavelength of the X-ray, denoted by λ, is accompanied by the Full Width at Half Maxima (FWHM) intensity of the primary peak detected at 2θ in radian, referred to as β. Additionally, θ represents Bragg's angle of diffraction, while k signifies a constant. In addition the lattice parameter (a), dislocation density (δ), microstrain (ε), and stress (σs) of the NiO crystals were calculated using the following equations:40
![]() | (2) |
![]() | (3) |
![]() | (4) |
![]() | (5) |
![]() | ||
Fig. 2 FE-SEM morphological images of the synthesized Ni1−xMnxO NPs: x = (a) 0, (b) 0.02, (c) 0.04, (d) 0.06, (e) 0.08 and variation of average particle size of NiO NPs with Mn doping percentage (f). |
EDS measurements demonstrate that the generated NPs had Ni and O peaks, which is compatible with the XRD results. The EDS spectrum of Mn-doped NiO nanoparticles (Mn = 0, 2, 4, 6, and 8 wt%) is shown in Fig. 3(a)–(e) shows just Ni, O, and Mn peaks (neglecting the peaks for platinum, as it was coated during the FE-SEM measurement process to form conductive layer), indicating the samples' high purity.
![]() | ||
Fig. 3 EDS spectrum of the synthesized Ni1−xMnxO NPs: (a) x = 0, (b) x = 0.02, (c) x = 0.04, (d) x = 0.06, (e) x = 0.08. |
The elemental composition of Mn-doped NiO (Mn = 0, 2, 4, 6, and 8 wt%) nanoparticles are given in Table 3 which was found from EDS spectrum analysis.
Sample | Atomic percentage (%) | ||
---|---|---|---|
Ni | O | Mn | |
Pure NiO | 77.48 | 22.52 | 0.00 |
2% Mn doped NiO | 82.95 | 15.23 | 1.82 |
4% Mn doped NiO | 79.30 | 17.05 | 3.65 |
6% Mn doped NiO | 77.11 | 18.23 | 4.66 |
8% Mn doped NiO | 74.10 | 18.85 | 7.05 |
In pure NiO, the stoichiometric atomic percentages of Ni and O are 77.48% and 22.52% by weight, respectively. In all the samples, the oxygen weight percentage is lower, and the Ni to O and Mn to O demonstrates the non-stoichiometric nature of NiO. The NiO ratio deviates from 1:
1, indicating non-stoichiometry. This non-stoichiometry leads to a color shift in NiO, with stoichiometric NiO appearing green and non-stoichiometric NiO appearing black. Nickel acetate tetrahydrate [C4H6NiO4·4H2O] was employed as a precursor in the synthesis of NiO NPs. Before synthesis, it was green, but after the production of NiO NPs, it turned black, clearly demonstrating the non-stoichiometric nature of NiO and the increased presence of oxygen vacancies. The presence of Mn atomic percent in the doped samples indicate Mn was successfully doped into the NiO lattice.
The deconvoluted Ni 2p XPS spectra of pure NiO NPs revealed (Fig. 4(a) bottom side) distinguishing 2p3/2 spin–orbit peaks at the binding energy of 853.8, 855.8, and 862 eV, 2p1/2 spin–orbit peaks at the binding energy of 871.5, 873.5, and 880 eV for Ni2+, Ni3+, and corresponding satellite peaks respectively. The O 1s XPS spectra of undoped NiO NPs show (Fig. 4(b) bottom side) a prominent peak at 529.6 eV, corresponding to NiO and relatively lower intensity peak at 531.7 eV for Ni2O3.47,48 No other additional peak was observed.
![]() | ||
Fig. 4 The XPS spectra of (a) Ni 2p, (b) O 1s (bottom side pure NiO and top side 6% Mn-doped NiO NPs), NiO, and (c) Mn 2p of 6% Mn-doped NiO NPs. |
In the case of the 6 wt% Mn-doped NiO NPs sample, the deconvoluted Ni 2p XPS spectra similarly showed (Fig. 4(a) top side) distinguishing 2p3/2 spin–orbit peaks at the binding energy of 853.9, 855.9, and 861.7 eV, 2p1/2 spin–orbit peaks at the binding energy of 872.1, 873.7, and 880 eV for Ni2+, Ni3+, and conforming satellite peaks in that order. On the other hand, the O 1s XPS spectra show (Fig. 4(b) top side) relatively low intensity peak at 529.6 eV for NiO, a prominent peak at 531.7 eV, corresponding to Ni2O3 and an additional peak at 529.8 eV for MnO2. The deconvoluted Mn 2p XPS spectra depicted (Fig. 4(c)) characteristic 2p3/2 spin–orbit peak at the binding energy of 641.6 eV and 2p1/2 spin–orbit peak at the binding energy of 652.9 eV for MnO2.47,48
The XPS results confirmed that synthesized pure NiO with preferred oxidation state of Ni2+ but 6 wt% Mn-doped NiO NPs e preferred oxidation state of Ni3+. This higher oxidation state of Ni3+ in the Mn doped nanoparticles could be attributed to the influence of the Mn dopant on the electronic structure of the NiO matrix, leading to changes in the oxidation states of the nickel ions.49 In the case of 6 wt% Mn-doped NiO NPs, Mn ions oxidation state is Mn4+. This result helps understand the structural and morphological data.
![]() | ||
Fig. 5 Optical absorption spectra of the NiO and Mn-doped NiO thin films (a) full visible range, and (b) enlarged within 300 to 400 nm wavelengths. |
The optical bandgap, Eg, of Mn-doped NiO (Mn = 0, 2, 4, 6, and 8 wt%) NPs thin film was calculated using the Tauc plot. The direct band gap values for the samples are obtained by the Tauc method as follows:
(αhν)n = A(hν − Eg) | (6) |
In eqn (6), α is the absorption coefficient of the NPs, A is a constant, Eg is the energy bandgap, hν is the incident photon energy and n is an index and can have values i.e. 2, 3, 1/2, 1/3 which depends on the band-to-band transition. During the band gap computation, it is observed that for n = 2, eqn (6) provides the best linear fit on the projection of the Tauc plot as shown in Fig. 6(a)–(e). The plots are drawn as αhν2versus hν. The straight linear portion of this curve is extrapolated onto the horizontal axes, and energy band gap values are achieved for different Mn/NiO ratios.
![]() | ||
Fig. 6 Tauc plot of the Ni1−xMnxO NPs: (a) x = 0, (b) x = 0.02, (c) x = 0.04, (d) x = 0.06, (e) x = 0.08. And variation of NiO NPs optical bandgap with different Mn doping concentrations (f). |
The values of the bandgap of undoped NiO, 2, 4, 6, and 8 wt% Mn-doped NiO NPs were confirmed to be 3.49, 3.69, 3.74, 3.79, and 4.10 eV respectively, as depicted in Fig. 6(a)–(e). Fig. 6(f) shows the synthesized NiO NPs' optical bandgap variation against the Mn doping concentration. The optical bandgap values increase with Mn doping. Mn incorporation into NiO NPs decreases the average crystallite size, indicating a size-reduction effect.17,51,52 Additionally, the optical band gap of the Mn-doped NiO NPs increases due to the quantum size effect, with a shift towards higher energy levels observed in the UV-vis absorption spectra.17,51,52 These changes in size and band gap are crucial for tuning the properties of the nanoparticles, making them potentially useful in applications such as photocatalysis and magnetic materials with tailored optical characteristics.
The mobility of charge carriers in NiO nanoparticles doped with varying concentrations of Mn (2, 4, 6, and 8 wt%) reveals a distinguished tendency: an initial increase, hit the highest point at 6 wt%, followed by a decrease at 8 wt% (Fig. 7(b)). This behavior can be explained by considering the impact of Mn doping on the NiO lattice structure and the resulting charge carrier dynamics.
The trend in mobility with increasing Mn doping reflects a balance between enhancing carrier concentration and minimizing structural defects and scattering centers. The resistivity of NiO nanoparticles doped with varying concentrations of Mn (2, 4, 6, and 8 wt%) shows (Fig. 7(c)) a trend opposite to that of mobility: it decreases initially, reaching a minimum at 6 wt%, and then increases at 8 wt%. This behavior can be explained by examining the interplay between carrier concentration and mobility, as resistivity is inversely proportional to the bulk concentration and mobility. The observed trends in resistivity and mobility with varying Mn doping levels reflect the delicate balance between carrier concentration and mobility. The increase in resistivity at higher doping levels (8 wt%) is primarily due to the significant reduction in mobility, which dominates over the effect of the high carrier concentration.
The electrical parameters of various layers employed during the simulation are reported in Table 1. Fig. 8(a) depicts the current density Vs applied voltage curve for the devices with various Mn-doped NiO HTL layers. All the devices had identical short circuit current (JSC) and open circuit voltage (VOC) but due to their efficiency (η) differed based on their fill factor (FF) (Fig. 8(b)). The device with pure NiO HTL layer had the η of 14.66%. Devices using 2, 4, 6, and 8 wt% Mn-doped NiO HTL layers exhibited η of 15.68, 16.35, 17.07, and 13.71% in that order. 6 wt% Mn-doped NiO thin film showed the highest mobility with the lowest resistivity, leading to best performance device.
The enhancement in carrier mobility, peaking at 6 wt% Mn doping, further underscores the potential of Mn-doped NiO for electronic applications, as confirmed by Hall measurements. SCAPS-1D simulations demonstrated the efficacy of the synthesized NPs in perovskite solar cells (PSCs), achieving a peak power conversion efficiency (PCE) of 17.07% with 6 wt% Mn-doped NiO NPs as the hole transport layer (HTL). These findings pave the way for further exploration of Mn-doped NiO in photovoltaic devices, offering a viable route toward the development of high-efficiency, stable solar cell technologies. Future research may explore optimizing doping concentrations and exploring other transition metals to further enhance the performance of such devices.
Footnote |
† Authors contributed equally. |
This journal is © The Royal Society of Chemistry 2025 |