Open Access Article
Monika Verma
ab,
Sanjeev Gautam
*b,
Bibek Ranjan Satapathy
c,
Weon Cheol Lim
d,
Ram Charan Meena
e,
Devarani Devi
e,
Suvankar Chakraverty
c and
Keun Hwa Chae
d
aEnergy Research Centre, Panjab University, Chandigarh 160014, India
bAdvanced Functional Materials Lab, Dr S.S. Bhatnagar University Institute of Chemical Engineering & Technology, Panjab University, Chandigarh 160014, India. E-mail: sgautam@pu.ac.in; Tel: +91 97797 13212
cInstitute of Nano Science and Technology, Mohali, Punjab 140306, India
dAdvanced Analysis & Data Center, Korea Institute of Science and Technology, Seoul 02792, South Korea
eInter-University Accelerator Centre, New Delhi 110067, India
First published on 5th December 2025
Silicon solar cells continue to dominate photovoltaic technology, holding a market value of ∼98% with an efficiency of 13–24% at the commercial level, which is limited by the recombination process and generated defects during the fabrication process. This study presents p–n junction fabrication using the ion beam technology, where boron species are implanted at a low energy of 35 keV into n-type Si (100). Doping was confirmed by X-ray photoelectron spectroscopy (XPS), which outperformed other conventional techniques (RBS and XRD) with exceptional elemental detection sensitivity. The shift in binding energy was observed to be 0.24 eV for the main peak in the silicon 2p spectra, resulting from the incorporation of boron into the silicon lattice. The local electronic environment modification was investigated by near-edge X-ray absorption fine structure (NEXAFS) spectroscopy at the O K-edge, which showed local hybridization consistent with boron incorporation and was also validated by FEFF simulations. Moreover, transport measurements exhibited diode-like I–V characteristics obtained using linear sweep voltammetry that were consistent with the Shockley diode model, indicating the formation of a p–n junction and notable suppression of the leakage current to 0.63 µA. Collectively, these findings evidence that the ion beam technology is a viable approach for the fabrication of reduced-defect structures, which are essential for the advancements of photovoltaics.
000 TWh between 2023 and 2030. Crystalline polysilicon-based cells2 account for over 98% of the total market share, surpassing other efficient cell designs like passivated emitter and rear cell (PERC) or TOPCon, heterojunction and back contacts. In spite of their prevalence, silicon solar cells have limited efficiency due to defects in their structure, promoting radiative and non-radiative recombination rates, Auger and surface recombination, Shockley-Read-Hall (SRH) recombination, and low lifetime of separated charge carriers.3 Even a marginal increment in the lifespan of Si solar cells or an improvement in their efficiency can lead to significant global economic benefits by increasing productivity per unit of input.
Commercial procurable silicon solar cells are typically fabricated via thermal diffusion, plasma-enhanced fabrication, or RF sputtering techniques, which often introduce defects in the lattice of the solar cells.4 These defects trap the photo-generated electron–hole pairs after the incidence of light, leading to the recombination of charge carriers and resulting in a decrease in the quantum efficiency of the cells.5,6 To mitigate these effects, solar cell design should be incorporated with defect engineering strategies like (i) reducing the occurrence of defects, (ii) eliminating impurities from the active regions of the device or rearranging them into configurations that are less detrimental, and (iii) implementing defect passivation.7 For defect-free structures, stability, and high-performance silicon solar cells, the ion beam technology offers a promising approach for the fabrication and modification of the silicon lattice.8
Ion beam implantation can be extensively used for the fabrication of p–n junctions by doping the ions into the host atom, providing a novel approach for next-generation solar cells.9 This technology offers repeatability, reliability, and precision engineering for defect-minimized structures with enhanced transport properties. It involves the ions traveling from the source materials through the beam line with defined energy and coming to rest beneath the surface of the substrate depending on their depth profile.10,11 It has been employed for doping various ions in the production of ICs, like boron, phosphorus, arsenic, indium, antimony, nitrogen, and boron difluoride.12 Ion beam-induced damage can be modeled using computational tools, such as SRIM/TRIM software, for the calculation of the range and transport mechanism into matter up to energies of 2 GeV amu−1. It is created by Biersack, Ziegler, Littmark, et al.13 and utilizes a Monte Carlo approach to predict the various parameters of ion–target interactions. These parameters include ion penetration depth, beam straggle, electronic and nuclear energy losses, energy deposition within the target material, phonon generation and its coupling with electrons, sputtering rates, vacancy concentration, displacement per atom (dpa), among others.14,15 Research conducted by Mikoushkin et al.16 investigated the implantation of nitrogen ions (N2+) in n-GaAs. It created a nitride nanolayer and formed a cluster of GaAs1−xNx, which was detected using the X-ray photoelectron spectroscopy (XPS) technique at a specific binding energy scan of N 1s. In a subsequent study, the same group17 implanted the N2+ ion again at an energy of 1.5 keV. However, the detection of the fabricated layer was performed by high resolution XPS with synchrotron radiation. Their findings concluded that the primarily fabricated nitride layer had a wide band-gap (GaN) with the characteristics of quantum dots. Furthermore, the same group conducted another study,18 with different Ar+ ions at an energy of 1.25 keV ion irradiation for the conversion of conductivity from n to p-type GaAs, forming a p–n junction, studied using XPS. Additionally, Qahtan et al.19 studied the production of oxygen vacancies in TiO2 mesoporous films using Ar+ ions at an energy of 500 eV. High resolution XPS confirmed the presence of oxygen, which was linked to lattice oxygen, oxygen vacancies, and absorbed oxygen.
In this study, a p–n junction was fabricated by implanting boron ions onto the upper layer of an n-type Si(100) substrate using ion beam implantations.20 The depth penetration and doping density were precisely controlled by the energy and the fluence of the implantation, respectively,21 enabling the formation of a uniform, defect-minimized layer for efficient electron–hole pair generation. Unlike conventional fabrication methods, here, employment of ion beam technique was particularly done for the effective doping of boron. Optimal doping at a certain fluence and energy is essential for a precise bandgap, and deviation from these parameters can enhance thermal fluctuations and increase the cell temperature.22 Further, an excessive doping concentration can alter charge carrier dynamics by extending their mean free path, ultimately reducing carrier lifetime and negatively impacting quantum efficiency. Prior to evaluating the transport properties using the I–V characteristics of a fabricated p–n junction, it is essential to verify and ensure the doping of boron in Si(100). Moreover, X-ray diffraction (XRD), Rutherford backscattering spectrometry (RBS), and secondary ion mass spectrometry (SIMS) are well-established methods owing to their respective purposes but are unable to detect implanted ions.23,24 However, X-ray photoelectron spectroscopy (XPS) plays an important role in surface analysis and in verifying the boron presence in the Si(100) lattice. It is a powerful technique for the detection of chemical bonds, charge states, and elemental compositions, which also provides some external information, like concentration, core-level lifetimes, multi-electron processes, and other phenomena, like Auger electron emissions, plasmons, and inelastic energy losses.25,26 In addition, near-edge X-ray absorption fine structure (NEXAFS) spectroscopy was performed at the O K-edge to study the hybridization and to examine the changes in the local electronic environment after doping, along with the simulations using FEFF9.6, providing theoretical and physical aspects of hybridization effects.27 The I–V measurements further validated the successful p–n junction formation with a significant reduction in leakage current, which can enhance the performance over conventionally fabricated solar cells.
:
1
:
5 ratio. The final step involved dipping the wafers in 2% diluted hydrofluoric acid for 2 minutes, followed by rinsing with DI and then blow-drying. During characterizations, each sample was chemically etched out for the removal of the surface SiO2 layer with a solution of 5 ml of nitric acid (70%) and 3 ml of hydrofluoric acid (49%) after optimization based on the calculation of the etching rate and depth of deposition.
![]() | ||
| Fig. 1 SRIM/TRIM simulation for boron ions at an energy of 35 keV: (a) stopping range of ions, (b) lateral range of ions, and (c) energy loss via phonons. | ||
| hν = BE + KE + φ, | (1) |
Here, the XPS of the thin films was characterized using a Thermo Fisher Scientific K-Alpha spectrometer instrument at the Institute of Nano Science and Technology (INST), Mohali, India. The instrument consists of a monochromatic light source, which produces X-ray Al Kα, with an energy resolution of 0.5 eV and standardized using pure gold metal. The pass energy of the XPS was 200 eV, with a good balance between the signal intensity and high resolution.
Near-edge X-ray absorption fine structure (NEXAFS) spectroscopy was performed at the 10D (XAS KIST) beamline of the Pohang Accelerator Laboratory (Pohang, South Korea). NEXAFS data were recorded in total electron yield (TEY, surface mode),32 and calibration of energy was performed through standards in the grating energy range with an energy resolution of 0.04 eV. Further, the data were normalized by subtracting the background and fitting the pre- and post-edge tails using Athena.33 I–V characteristics were measured using PSTrace software from PalmSens instruments at IIT Jodhpur, India.
Fitting of XPS spectra involves the proper background function for the modeling of inelastic electrons, followed by a peak fitting model using a suitable function to deconvolute the core-level spectrum into sub-peaks.38 The backgrounds of Si 2p and O 1s were removed using Shirley39 and B 1s using the Tougaard baseline treatment. In addition, the Voigt function (Gaussian + Lorentzian) was used for boron's peak fitting, and the Gaussian function was used for Si 2p and O 1s, keeping the FWHM fixed. An atomic concentration of boron was estimated40 using eqn (2) considering the relative sensitivity factor (RSF) for boron 0.376, silicon 0.9, and oxygen 2.881, which is instrument dependent.
![]() | (2) |
is for all n elements presented in the sample. The surface atomic concentration of boron was evaluated as 6.23%, 18.40%, and 19.33% for BN1, BN2, and BN3, respectively. The abrupt changes in the atomic concentration for the sample BN2 having fluence of 10 times higher than BN1 is caused by the non-linear relationship between ion fluence and atomic concentrations. At lower fluence (BN1), the point defect has limited interaction, and above (5 × 1014 ions per cm2), defect clusters start developing, which is attributed to sputtering effects and cascade overlap.41 However, for BN3, the lattice reached a steady state of damage formation, leading to the formation of the cluster instead of the substitutional sites, as explained using the modified Kinchin–Pease equation.42
Fig. 3 illustrates the XPS spectra of doped samples BN1, BN2, and BN3 for the elements Si 2p, B 1s, and O 1s, respectively. As depicted in Fig. 3(a), the Si 2p spectrum assured the doping of boron with a shift towards higher binding energy at 102.75 eV with an oxidation state of Si+4. As depicted in Fig. 3(b), B 1s spectra showed the doping of boron exhibiting different oxidation states B0, B+0.33, and B+3 at 186.98 eV,43 189.42 eV,44 and 192.06 eV45 binding energy, respectively. The chemical states of boron corresponding to these binding energies were metallic, B6O, and B2O3, making a composition with oxygen presented in the lattice of silicon24,46 (and reference data taken from the NIST XPS database). The O 1s spectrum in Fig. 3(c) had the main peak at 531.95 eV, representing the presence of C
O, due to the contamination from carbon tape, and the deconvoluted peak at 532.07 eV demonstrates that the low binding energy component arises due to the absorption of OH−.47
Besides, Fig. 3(d)–(f) depict Si 2p, B 1s, and O 1s scans of the BN2 sample (1 × 1015 at 35 keV), respectively. The Si 2p spectrum confirmed the doping of boron with a shift towards higher binding energy with oxidation state Si+4 at 102.88 eV and verified the presence of B15.9Si19.0O65.1.48 After implantation, boron interacted with silicon in the matrix chemically and formed a bond with the oxidation states of B0, B+0.33, and B+3. The deconvoluted peaks at binding energies43 of 186.28 eV, 189.76 eV, and 192.9 eV were related to their existing composition in metal form, composition with oxygen,44 silicon and boron,48 respectively. Similarly, Fig. 3(f) shows the O 1s spectrum with binding energies of 532.14 eV and 532.72 eV.49
Fig. 3(g)–(i) illustrate the Si 2p, B 1s, and O 1s spectra of the BN3 sample (1 × 1016 at 35 keV), respectively. The implantation of boron into the lattice of silicon was very prominent that associates it, with different oxidation states of B0 at binding energy43 of 186.88 eV corresponds to metallic boron, B+0.33 state at binding energy of 190.8 eV for B6O,44 and B+3 at binding of energy 103.8 eV making a composition B15.9Si19.0O65.1.48 Fig. 3(g) and (i) confirm the doping of boron with shifting towards higher binding energy with oxidation state Si+4 at 102.99 eV and verified the presence of B15.9Si19.0O65.1.48 As the fluence of boron increased, the binding energy shifted towards the higher ranges.50 All the implanted samples had deviated binding energies, considering that the shielding effect (addition or removal of valence electrons), matrix interactions, and charging effects resulting from charge accumulation on the surface led to transient enhanced diffusion (TED). Interstitial atoms produced during implantation are responsible for this effect. Further, defects in the lattice, interstitial cluster, etc., could trap or absorb these interstitially produced atoms.51 The elements present in the silicon matrix interacted with each other chemically and electronically, changing their environment and causing a shift in the binding energy.
In addition to XPS, near-edge X-ray absorption fine structure (NEXAFS) spectroscopy at the O K-edge of pristine and BN3 was recorded, as depicted in Fig. 4(a). Since the surface of silicon is highly reactive to oxygen, it readily forms a native SiO2 layer, which is removed using chemical etching. However, interfused oxygen (production of wafers) within the lattice of the fabricated samples provides valuable information about the hybridization of oxygen and the electronic modification of the respective doped elements in the implanted sample. The spectral features labeled a1 and a2 in Fig. 4(a) show the hybridization of the s orbital of oxygen with silicon atoms, where subsequent differences could be observed in the doped sample. The intense edge spectral features a3 and a4 were assigned to the hybridization between O 2p and Si 3p,52 in which doping definitely disturbed the arrangements and respective hybridization. The distributions of energy in the implanted samples are clearly distinguishable from the pristine ones, indicating that the modification in the electronic structure resulted from the doping of boron. This suggests that boron implantation perturbed the Si–O bonds in the fabricated samples compared to pristine samples, affecting the spectral feature, which was in agreement with the XPS findings.
The obtained experimental spectra were also simulated using FEFF9.6, considering the exchange potential Hedin Lundqvist (HL) (−2 for the real part) and Dirac Hara (DH) (0.2 for the imaginary part) to account for the inelastic behavior of emitted photoelectrons and complex electron–electron interactions, respectively. The calculations were performed using 6 full multiple scattering (FMS) paths to determine the cluster size. However, larger FMS values could be used but showed negligible improvement over the chosen value. The reduction amplitude factor (SO2) that affected the amplitude of the absorption signal was restricted by many-body effects, which was set to 0.9, meaning a reduction of 10% signal, yielded the best fit to the experimental data, providing information about the local electronic environment. The self-consistent field was set to 3, with the local density of states varying from −20 (minimum energy) to 10 (maximum energy). The exchange potential for the fitting was HL, which was −2.0 in real parts, representing the increment in conduction area with relative molecular orbital ratios,27 and DH was 0.2, as illustrated in Fig. 4(b).
Moreover, Fig. 5 depicts the diode-like I–V characteristics obtained using PSTrace software from PalmSens instruments. All the measurements were conducted at room temperature and under normal daylight without any particular light illumination using a tungsten probe. The data were recorded through linear sweep voltammetry, having a potential ranging from −2 V to 2 V, with a scan rate of 0.05 V s−1 and step intervals of 0.05 V.
Fig. 5(a) illustrates the pristine sample, which exhibits a relatively high leakage current (Il) of 1.5 µA compared to the implanted samples. Fig. 5(b) and (c) correspond to boron-implanted samples BN1 and BN2, showing the formation of a p–n junction,53 where the relationship between current–voltage and the Shockley–diode equation is expressed as follows:
| I = I0(exp(qV/ηKT) − 1), |
The leakage currents 0.85 µA and 0.63 µA of BN1 and BN2, respectively, at −2 V are reduced in comparison to the pristine sample.54 Interestingly, Fig. 5(d) reveals a distinct behavior for the sample implanted with a high fluence of 1 × 1016 ions per cm2, showing the reduction in knee voltage (Kknee). This demonstrated a sharp elevation in current and a leakage current of 1.57 µA, suggesting the possible formation of intermediate energy states within the material due to damage to the lattice. Another contributing factor could be TED caused by the supersaturation of silicon interstitial due to point defect.55 Furthermore, this excess silicon interstitial generated, combined with boron doping, occupies the electrically inactive, metastable clusters, forming the mobile boron-interstitial pairs and subsequently large immobile clusters.56,57 These findings demonstrate that precise and controlled doping via implantation significantly improves the I–V characteristics of the diode, making it a promising candidate for photovoltaic applications.
| This journal is © The Royal Society of Chemistry 2025 |