Next Article in Journal
Effect of Hot-Pressing Process on Mechanical Properties of UHMWPE Fiber Non-Woven Fabrics
Previous Article in Journal
Mechanical, Durability, and Microstructure Assessment of Wastepaper Fiber-Reinforced Concrete Containing Metakaolin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing Adhesion and Reducing Ohmic Contact through Nickel–Silicon Alloy Seed Layer in Electroplating Ni/Cu/Ag

1
School of Materials Science and Engineering, Jiangsu University, Zhenjiang 212013, China
2
State Key Lab of Photovoltaic Science and Technology, Trina Solar Co., Ltd., Changzhou 213031, China
3
Institute of Technology for Carbon Neutralization, Yangzhou University, Yangzhou 225127, China
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(11), 2610; https://doi.org/10.3390/ma17112610
Submission received: 23 April 2024 / Revised: 13 May 2024 / Accepted: 13 May 2024 / Published: 28 May 2024

Abstract

:
Due to the lower cost compared to screen-printed silver contacts, the Ni/Cu/Ag contacts formed by plating have been continuously studied as a potential metallization technology for solar cells. To address the adhesion issue of backside grid lines in electroplated n-Tunnel Oxide Passivating Contacts (n-TOPCon) solar cells and reduce ohmic contact, we propose a novel approach of adding a Ni/Si alloy seed layer between the Ni and Si layers. The metal nickel layer is deposited on the backside of the solar cells using electron beam evaporation, and excess nickel is removed by H2SO4:H2O2 etchant under annealing conditions of 300–425 °C to form a seed layer. The adhesion strength increased by more than 0.5 N mm−1 and the contact resistance dropped by 0.5 mΩ cm2 in comparison to the traditional direct plating Ni/Cu/Ag method. This is because the resulting Ni/Si alloy has outstanding electrical conductivity, and the produced Ni/Si alloy has higher adhesion over direct contact between the nickel–silicon interface, as well as enhanced surface roughness. The results showed that at an annealing temperature of 375 °C, the main compound formed was NiSi, with a contact resistance of 1 mΩ cm−2 and a maximum gate line adhesion of 2.7 N mm−1. This method proposes a new technical solution for cost reduction and efficiency improvement of n-TOPCon solar cells.

1. Introduction

Benefiting from carbon peaking and carbon neutrality strategies, the development of new energy has become increasingly important. Crystalline silicon solar cells are highly favored due to their high conversion efficiency and good stability [1,2]. Among various types of crystalline silicon solar cells, N-TOPCon solar cells have become the most competitive solar cells [3]. However, this cell type faces the challenge of high silver paste consumption during the metallization process. Screen printing (SP) metallization is a critical step in the production process of crystalline silicon solar cells. However, it is accompanied by the issue of high silver consumption [4]. Reportedly, the highest proportion of costs related to the production of crystalline silicon solar cells (not including silicon wafer prices) are attributed to the use of expensive silver paste. The silver consumption of n-TOPCon solar cells is about 110–120 mg per piece. As a result, it is crucial to explore alternatives to screen-printing processes in order to effectively decrease the cost of metallization [5].
At present, the main metallization methods reported to replace screen printing silver paste are as follows: electroplating (electrochemical deposition metallization) [6,7,8], laser transfer printing [9,10], inkjet printing [11,12,13], etc. Electroplating has emerged as a main research subject among these methods, owing to its significant cost reduction in metallization by employing copper electrodes. This is because copper (Cu 580 S/m) offers similar electrical conductivity to silver (Ag 630 S/m), but at a fraction of the price [14,15]. However, it should be noted that copper has a high diffusion rate in silicon and acts as a deep-level impurity and high recombination center, making it susceptible to the formation of Cu3Si compounds that could adversely affect the carrier lifetime in silicon [16,17] The current solution is to electroplate an Ni intermediate layer between the copper layer and silicon substrate, preventing direct contact between the Cu and underlying Si, thereby avoiding the formation of Cu3Si [18].
Nevertheless, the Si/Ni/Cu gate lines have low adhesion since the Ni layer and Si substrate are merely physically connected [19,20,21]. Fortunately, by performing thermal silicidation at different temperatures, chemical reactions can occur between nickel (Ni) and silicon (Si), forming silicides such as the Ni2Si phase, NiSi phase, and NiSi2 phase [22,23,24,25,26]. Among these, the NiSi phase not only improves the adhesion of gate lines and reduces contact resistance but also enhances the thermal stability of cells. The findings presented by Kluska et al. [27] suggest that the thermal silicidation anneal caused by NiSi alloy improves gate line adhesion. Furthermore, research by Kale et al. suggests that the NiSi/Si interface exhibits better thermal stability compared to the Ni/Si interface [28]. In addition, direct annealing of Si-Ni-Cu can result in the formation of Si-NiSi-Ni-Cu. However, due to the significant difference in the diffusion rates of Cu, Ni, and Si at the same temperature, Kirkendall voids may form at the interfaces, which is detrimental to achieving high thermal stability in the solar cell [29]. In view of this, in this study, the NiSi alloy layer was prepared on the Si substrate using an annealing process, followed by electroplating Ni and Cu successively on its surface, ultimately producing Si-NiSi-Ni-Cu gate line contacts.
In this work, the Ni layer was deposited on the back of the crystalline silicon solar cell and then annealed in the temperature range of 300 °C to 425 °C to form a seed layer (NixSiy). Subsequently, Ni and Cu were sequentially electrodeposited on the seed layer, and, finally, the Si-NiSi-Ni-Cu contact structure was constructed. In order to further explore the phase information and its formation mechanism in the Ni/Si alloy layer, we used X-ray diffraction (XRD), Raman spectroscopy (Raman), X-ray photoelectron spectroscopy (XPS), and other test methods for detailed analysis. In order to evaluate the effect of the seed layer (NixSiy) on the performance of solar cells, we further adapted the Sun-Voc test method. The test results show that the Ni/Si alloy formed by annealing at 375 °C is closer to the NiSi phase, which not only has lower contact resistance but also exhibits an excellent grid-bonding force. This discovery is of great significance for improving the performance of solar cells.

2. Experiment

2.1. Fabrication of Solar Cells

All experimental wafers use n-type monocrystalline silicon wafers with an area of 158.75 mm × 158.75 mm and a thickness of 170 μm. Firstly, the random pyramid structure was prepared on the surface of the silicon wafer, and then the p-n junction was formed on the front side by the diffusion method. Then, ultra-thin SiO2 and n-type polysilicon (n-poly) were prepared by low-pressure chemical vapor deposition (LPCVD) to form a TOPCon contact structure. The AlOx film with a certain thickness was prepared on the front of the solar cell as the front passivation layer by atomic layer deposition (ALD). Then, the front and back SiNx films were deposited by plasma-enhanced chemical vapor deposition (PECVD), and the front AlOx and the front and back SiNx layers were opened using a UV-ps laser. The width of the front laser contact opening was about 7 μm, and the width of the back laser contact opening was about 40 μm.
The nickel layer was deposited on the back of n-TOPCon using electron beam deposition (0.5 A s−1 deposition rate). The samples were annealed at various temperatures of 300–500 °C in an inert nitrogen (N2) atmosphere. The unreacted Ni was removed with an etching solution (mixed solution of H2SO4 and H2O2). The back plating process sequence includes a single-sided low-concentration HF pretreatment to remove any oxide layer within the laser contact opening (LCO). Subsequently, nickel and copper plating was performed by light-induced plating (LIP). Subsequently, the solar cell was flipped using almost the same process sequence for depositing Ni and Cu. However, this time, the plating does not require light because the pn junction is in forward bias (forward bias (FBP)). The front also appears in the electroless silver bath. The thickness of the Ni layer is about 0.5 μm, the thickness of the Cu layer is about 7 μm, and the thickness of the Ag layer is 0.3 μm (where Cu is used as a conductive layer, Ag is used as an anti-oxidation layer, and it is convenient for subsequent welding to make components). The flow chart of the seed layer crystalline silicon solar cell is shown in Figure 1.

2.2. Characterization Methods

The surface morphology of the seed layer was observed by scanning electron microscopy, and the chemical composition of the sample was determined with an Energy-Dispersive Spectrometer (EDS) (Carl Zeiss AG, Oberkochen, German). The surface roughness of the seed layer was determined with a high-resolution optical profiler (ZetaTM-3.0) (Tokyo, Japan). The X-ray diffraction (XRD) measurements were performed using a Cu Kα radiation diffractometer with a scan rate of 0.01° s−1 in 2θ range of 5–80° to obtain detailed phase formation information. The phase transition for different types of Ni silicide was characterized through X-ray Photoelectron Spectroscopy (XPS, Thermo Fisher Scientific, Waltham, MA, USA) and Raman Spectroscopy (Thermo Fisher Scientific, Waltham, MA, USA). For measuring the contact resistance of the seed layer stack, the Transmission Line Model (TLM) technique was employed. For the adhesion measurement of the seed layer, a peel force test was conducted by using a universal testing machine. Meanwhile, the electrical properties of n-TOPCon solar cells with seed layers with different annealing conditions were investigated by evaluating from Suns-Voc (Sinton FCT650) (Boulder, CO, USA) measurements.

3. Result and Discussion

In order to evaluate the surface appearance of the seed layer, SEM tests were carried out. Figure 2a is the SEM diagram of the annealing temperature of 375 °C. It can be seen from the figure that the surface has an obviously uneven particle feeling. Figure S1 shows the surface morphology of different annealing temperatures, all of which show rough graininess. The production of nickel–silicon alloy compounds or the ablation state brought on by the laser process could be the origin of this morphology, which would increase the surface roughness. We performed EDS (as shown in Figure 2b) to get more insight into the elemental makeup of the annealed samples. The findings of the research demonstrate that the elements Ni and Si are dispersed equally across the sample’s surface. The oxygen detected on the surface of the sample is caused by the oxidation of the sample surface in the air. To further confirm that the formed compound is a nickel–silicon alloy, we performed XRD, Raman, and XPS tests.
After annealing at different temperatures, Ni and Si form different phases, such as Ni2Si, NiSi, and NiSi2 [30,31]. Among the different nickel silicide phases formed by the seed layer, the NiSi phase is the lowest resistivity phase in the silicide phase. From the XPS diagram of the seed layer annealed at 400 °C shown in Figure 2c, it can be seen from the full peak that both the Ni peak and the Si peak exist, and the Ni2p and Si2p peaks are most suitable at 852.7 and 98.8 eV, respectively. The Ni2p peaks of Ni2Si, NiSi, and NiSi2 are all around 853 eV, and the positions of Si2p peaks are also very close [32,33]. Since the observed Si2p peak is close to the Si2p peaks of NiSi and Ni2Si at 98.85 eV and 98.75 eV [34], the XPS needs to be combined with XRD analysis. In addition, Si2p has a prominent peak at 103.3 eV, corresponding to the Si2p peak of SiO2, which may be caused by oxidation on the surface (Figure S2) [35].
Based on XRD measurements conducted under various annealing conditions, the atomic ratio of the compound nickel silicide was determined (Figure 2d). The Ni-Si alloy formed at a 300 °C annealing temperature is mainly in the Ni2Si phase because it is a thermodynamically favorable phase in the nickel-rich region [25]. The peaks at 26.99°, 36.49°, 47.80°, and 69.02° correspond to the (100), (002), (110), and (202) planes of Ni2Si, respectively, while the peaks at 23.1° and 45.84° belong to the (101) and (112) planes of NiSi. Because it is not conducive to forming silicon-rich silicide phases in a nickel-rich environment, there is an increased thermodynamic force for Si to diffuse into Ni and form NiSi phases with an increasing annealing temperature. The existence of NiSi during annealing at 400 °C is attributed to the preferential diffusion of silicon through the silicide interface and the interaction with Ni2Si, which makes it completely transformed into the NiSi phase [36]. The peaks at 23.1°, 31.67°, 37.82°, 45.84°, and 68.83° are attributed to the (101), (011), (201), (112), and (122) planes of NiSi, respectively, while the peaks at 56.31° and 75.59° correspond to the (311) and (331) planes of NiSi2. Si diffuses further and combines with NiSi to generate NiSi2 when the annealing temperature is raised some more. At the annealing temperature of 500 °C, the peaks at 28.6°, 56.3°, 69.35°, and 76.59° correspond to the (111), (311), (400), and (311) planes of NiSi2, respectively, while the peaks at 23.1°, 37.85°, and 45.84° are attributed to the (101), (201), and (112) planes of NiSi. The Ni is precisely deposited on the back of the n-TOPCon solar cells by electron beam evaporation. The Ni/Si alloy formed by annealing in a nitrogen atmosphere is a mixture, not a single compound.
The composition of the formed compound is further verified by Raman spectroscopy measurements, as shown in Figure 2e. At a 300 °C annealing temperature, the peaks at 100 cm−1 and 140 cm−1 belong to the Ni2Si phase. The peak at 216 cm−1 belongs to the phase of NiSi. At 400 °C, the NiSi phase is present at the major peak position in addition to the Ni2Si phase at 100 cm−1 [37]. It has an obvious peak intensity at 194 cm−1, 216 cm−1, 292 cm−1, and 363 cm−1, and the peak at 396 cm−1 is attributed to the NiSi2 phase. When the annealing temperature is increased to 500 °C, the peak of the Ni2Si phase disappears, while the peaks at 288cm−1 and 396cm−1 belong to the NiSi2 phase, and the remaining peaks are in the NiSi phase. The absence of the Raman-active Si phonon band is found at 521 cm−1 at different annealing temperatures, which confirms that the Ni-Si alloy films formed using electron beam deposition of Ni on Si under different annealing conditions provided a continuous, crack-free film and completely covered the Si surface [38]. The NiSi thin film as the seed layer is crucial for the complete surface coverage of the silicon in the laser opening area because the presence of any defects will cause the contact resistance to be uneven, thereby reducing the solar cell performance. Based on the analysis of XRD and Raman spectra, we conclude that the Ni-Si alloy formed under different annealing conditions is a mixture, which is closest to NiSi at a 375 °C annealing temperature.
The hypothesis of the diffusion formation mechanism of nickel–silicon alloy formed by thermal silicification is proposed (Figure 3) based on the above microscopic characterization tests. At the temperature of 300 °C, Ni2Si would preferentially form due to the diffusion of Ni atoms to the n-poly layer. With the increase in the annealing time, the thickness of Ni2Si increases under the premise of ensuring the thickness of metal nickel. The Ni2Si with a low chemical potential, which acts as a thermodynamic driving force, would have reacted for Formula (1) [39]:
2 Ni + Si Ni 2 Si   ( 183.6   kJ   mol - 1 )
When the annealing temperature is 400 °C, both the remaining Ni and Ni2Si react with n-poly to form NiSi, and Ni2Si will gradually react with Si until it is completely or mostly consumed, which would decrease the overall thermodynamic energy. The chemical potential and reaction formula are shown in (2) and (3) [40]:
Ni + Si NiSi   115.4   kJ   mol - 1
Ni 2 Si + Si 2 NiSi   47   kJ   mol - 1
When the annealing temperature is 500 °C, the Si atoms of the n-poly layer will diffuse into NiSi to form NiSi2. With the increase in annealing time, NiSi will be gradually consumed. The formation rate of NiSi2 is much slower than other silicification processes, following the chemical formula in (4) [41]:
NiSi + Si Ni Si 2 17.4   kJ   mol - 1
From low temperature to high temperature, the transformation order of nickel silicide is Ni2Si, NiSi, and NiSi2, which is related to the participation of Si atoms with high heat energy. The growth of thin films with only the NiSi phase is very important because the mixed film of NiSi and high resistivity the Ni2Si phase may lead to higher resistance-induced loss of the cells.
The thermal silicide samples at different annealing temperatures are evaluated using the TLM test of contact resistance, which is compared with the Ni/Cu/Ag contact directly electroplated without a seed layer. When using the TLM, the total resistance RT between finger and finger is measured and plotted as a function of contact spacing L (shown in Figure 4a). As contact spacing L increases, the effect of the sheet resistance on the total resistance measurement increases, thus creating the slope of the TLM plot ρc [42].
Figure 4b is the contact resistance of the samples under different annealing processes. With the increase in annealing temperature, the contact resistance of the fingers on the back of the seed layer electroplated n-TOPCon is lower than that of the reference sample. In the contact resistance test on the back of the electroplated n-TOPCon solar cell, although the overall value difference is relatively insignificant, it can still be clearly concluded that the nickel–silicon alloy formed by the annealing thermal silicification process has played a positive role in effectively reducing the contact resistance.
Using a high-resolution optical profiler, the surface roughness of several annealed samples was examined to determine how it affected the adherence of the Busbars and Fingers. Figure 4c shows the surface roughness of the seed layer under different annealing processes. From the diagram, it is shown that the surface roughness value of the sample after annealing is higher than that after laser opening the dielectric layer, which indicates that the formed nickel–silicon alloy can increase the surface roughness. The roughness of the sample annealed at 375 °C is better than that of other annealing conditions because NiSi accounts for the largest proportion of the material formed at this temperature. To evaluate the possible role of the formed Ni-Si alloy in promoting the adhesion of the gate wire, the peel force test is conducted at each temperature condition, including directly electroplated Ni/Cu/Ag contacts without a seed layer (Figure 4d). At the annealing temperature from 300 °C to 425 °C, the average adhesion values increase first and then decrease. The highest adhesion of 2.7 N mm−1 is recorded at an annealing temperature of 375 °C. In addition, the adhesion of the samples with different annealing processes shows more than 1 N mm−1 [43]. The average adhesion of the sample without the seed layer is greater than 0.8 N mm−1, but there are still values that do not meet the bonding force standard. The electroplating contact has a multi-layer structure. To have a beneficial gate–wire bonding force, it is necessary to improve the adhesion of the Ni-Si interface. The results show that the bonding force of the seed layer samples is better than that of the direct electroplating Ni/Cu/Ag under different annealing processes.
When UV-ps acts on the SiNx layer on the back of the seed layer n-TOPCon precursor, the Si surface is destroyed, which is manifested as melting, a heat-affected zone, microcracks, and point defects. These destructions lead to enhanced diffusion, resulting in Ni cluster defects [44,45]. The electrical properties of the seed layer n-TOPCon solar cells under different annealing processes are tested with a Sinton FCT650, and the results are shown in Table 1. For reliable data, 10 cells are measured under different annealing conditions in each group. With the increase in the annealing temperature, The Fill Factor (FF) increases first and then decreases. The average FF of the cells annealed at 375 °C is the highest, which is 82.02%, which is 1.24% higher than that of the cell annealed at 425 °C. During the annealing process, the efficiency shows a similar trend, that is, at the annealing temperature of 425 °C, the efficiency reaches the lowest level. In view of the obvious loss of FF observed at 425 °C annealing conditions, we decided not to manufacture batteries with higher annealing temperatures (Table 2).
From the results of Figure 5a, it is shown that with the increase in annealing temperature, Rs decreases first and then increases, but it is better than that of Ni/Cu/Ag directly electroplated without seed layer, which is consistent with the results of the contact resistance test. Under the condition of an annealing temperature lower than 400 °C, the pFF does not change much, so the reduction from pFF to the measured FF equals the FF loss due to Rs. Under the premise of other unchanged parameters, the relationship between Rs and FF is the opposite, and the decrease in Rs will lead to an increase in FF, so the change in FF is mainly affected by Rs (Figure 5b). It can be seen from Figure 5c that the pFF has no downward trend at 375 °C, while the pFF at 400 °C decreases by 0.6%. When the annealing temperature reaches 425 °C, the pFF reaches 81.69%, a decrease of 1%. Raval et al. [46]. showed that the pFF was not severely degraded until it reached the 400 °C annealing temperature. The composite current density (J02) and the space charge area shunt resistance have an impact on the pFF, making it a useful metric for assessing the junction’s quality. If the metal atoms from the contacts form a trap site in the space charge region, the amount of shunt path and J02 will be increased and finally decrease the pFF. Since the edges are protected during plating, they are not affected by any ohmic shunt. Therefore, the shunting does not adversely affect the pFF, and the loss can be safely attributed to the recombination (J02) loss. The higher standard deviation of the average pFF value at 400 °C annealing is reflected in the recombination-based FF loss. The larger standard deviation of the average pFF value upon annealing at 400 °C represents the FF loss resulting from recombination.

4. Conclusions

In this study, a metal nickel layer is deposited on the backside of solar cells using electron beam evaporation, and excess nickel is removed by H2SO4:H2O2 etchant under annealing conditions of 300–425 °C to form a seed layer. XRD and Raman spectroscopy results indicate that optimizing the annealing temperature plays a crucial role in controlling the phases of the seed layer. TLM measurements showed that the specific contact resistivity of the Si/Ni stack followed the phase transition sequence with the changing annealing temperature. By annealing at 375 °C, the contact with NiSi as the primary phase was found with the lowest contact resistance. Peel strength tests along the Busbars indicated that the Si-NiSi-Ni-Cu gate line contact, prepared through annealing within the temperature range of 375 °C, exhibited superior adhesion with a peak value of 2.7 N mm−1. The IV measurement data indicated that the cells prepared at an annealing temperature of 375 °C exhibited the best performance, characterized by the highest FF of 82.02% and efficiency Eff of 23.67%. After the stability test, the efficiency of the solar cell still remains above 90%, indicating that the prepared crystalline silicon solar cell exhibits certain stability under specific test conditions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma17112610/s1, Figure S1: SEM images of samples at different annealing temperatures; Figure S2: XPS diagram of samples at different annealing temperatures (a) Ni peaks; (b) Si peaks.

Author Contributions

Conceptualization, Z.W. (Zhao Wang), H.L. and J.G.; methodology, Z.W. (Zhao Wang), Z.W. (Zigang Wang), G.C. and J.D.; validation, Z.W. (Zhao Wang), H.L. and Y.D.; writing—original draft, Z.W. (Zhao Wang) and Z.Z.; formal analysis, D.C., Z.W. (Zigang Wang) and K.W.; investigation, D.C. and Y.C.; data curation, Z.W. (Zigang Wang); resources, Z.Z.; funding acquisition, J.D.; supervision, J.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Special Fund for Science and Technology Innovation of Jiangsu Province (grant No. BE2022022-4), the Natural Science Foundation of Jiangsu Province (grant No. BK20230204), the Qing Lan Project, and the 333 Project of Jiangsu Province.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

Author Zhao Wang, Daming Chen, Zigang Wang, Kuiyi Wu, Zhuohan Zhang, Yifeng Chen and Jifan Gao were employed by the company Trina Solar Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Wang, X.L.; Huang, X.L.; Lv, J.; Wang, J.B.; Cao, M.; Wang, T.; He, J.P. Mechanical and Electrical Properties of Ni-Ag Metallized TOPCon Silicon Solar Cells Structured by Green Picosecond Laser. J. Electron. Mater. 2023, 52, 3859–3867. [Google Scholar] [CrossRef]
  2. Aleem, M.; Vishnuraj, R.; Krishnan, B.; Ashok, A.; Pullithadathil, B. Narrow Line Width Ni–Cu–Sn Front Contact Metallization Patterns for Low-Cost High-Efficiency Crystalline Silicon Solar Cells using Nano-Imprint Lithography. Energy Technol. 2023, 11, 2300090–2300103. [Google Scholar] [CrossRef]
  3. Pospischil, M.; Kuchler, M.; Klawitter, M.; Rodríguez, C.; Padilla, M.; Efinger, R.; Linse, M.; Padilla, A.; Gentischer, H.; König, M.; et al. Dispensing technology on the route to an industrial metallization process. Energy Procedia 2015, 67, 138–146. [Google Scholar] [CrossRef]
  4. Chen, X.; Church, K.; Yang, H.; Cooper, I.B.; Rohatgi, A. Improved front side metallization for silicon solar cells by direct printing. In Proceedings of the 37th IEEE Photovoltaic Specialists Conference (PVSC), Seattle, WA, USA, 19–24 June 2011; pp. 3667–3671. [Google Scholar]
  5. Huang, Y.W.; Zhou, Z.Q.; Yang, H.X.; Wei, C.X.; Wan, Y.Y.; Wang, X.J.; Bai, J.G. Glucose application protects chloroplast ultrastructure in heat-stressed cucumber leaves through modifying antioxidant enzyme activity. Biol. Plant. 2015, 59, 131–138. [Google Scholar] [CrossRef]
  6. Shen, X.W.; Hsiao, P.C.; Wang, Z.M.; Benjamin, P.; Sean, L.; Alison, L. Refined nickel nucleation and plated metal adhesion induced by pulsed light-induced plating on picosecond laser-ablated silicon solar cells. Sol. Energy Mater. Sol. Cells 2022, 240, 111638–111648. [Google Scholar] [CrossRef]
  7. Chang, Y.C.; Wang, S.S.; Deng, R.; Li, S.Y.; Ji, J.J.; Chong, C.M. Investigation of laser doping and plating process for cost-effective PV metallization. Sol. Energy Mater. Sol. Cells 2022, 235, 111445. [Google Scholar] [CrossRef]
  8. Yu, C.; Gao, K.; Peng, C.W.; He, C.R.; Wang, S.B.; Shi, W.; Allen, V.; Zhang, J.T.; Wang, D.Z.; Tian, G.Y.; et al. Industrial-scale deposition of nanocrystalline silicon oxide for 26.4%-efficient silicon heterojunction solar cells with copper electrodes. Nat. Energy 2023, 8, 1375–1385. [Google Scholar] [CrossRef]
  9. Adrian, A.; Rudolph, D.; Willenbacher, N.; Lossen, J. Finger Metallization Using Pattern Transfer Printing Technology for c-Si Solar Cell. IEEE J. Photovolt. 2020, 10, 1290–1298. [Google Scholar] [CrossRef]
  10. Rodofili, A.; Wolke, W.; Kroely, L.; Bivour, M.; Cimiotti, G.; Bartsch, J.; Glatthaar, M.; Nekarda, J. Laser Transfer and Firing of NiV Seed Layer for the Metallization of Silicon Heterojunction Solar Cells by Cu-Plating. Sol. RRL 2017, 8, 1700085–1700091. [Google Scholar] [CrossRef]
  11. Abolmasov, S.N.; Abramov, A.S.; Ivanov, G.A.; Terukov, E.I.; Emtsev, K.V.; Nyapshaev, A.; Bazeley, A.A.; Gubin, S.P.; Kornilov, D.Y.; Tkachev, S.V.; et al. Heterojunction solar cells based on single-crystal silicon with an inkjet-printed contact grid. Tech. Phys. Lett. 2017, 43, 78–80. [Google Scholar] [CrossRef]
  12. Devenport, S.; Morrison, D.; Dunnill, S.; Baistow, I.; Dixon, R.; Yu, K.; Bell, R.; Clarke, E. Inkjet printed nickel silicide applied as front contact seed-layer for electrolessly copper plated silicon solar cells. Energy Mater. 2014, 9, 490–493. [Google Scholar] [CrossRef]
  13. Hatt, T.; Morawietz, T.; Bartsch, J.; Tutsch, L.; Glatthaar, M. Multifunctional Titanium Oxide Layers in Silicon Heterojunction Solar Cells Formed via Selective Anodization. Sol. RRL 2023, 7, 2300335–2300342. [Google Scholar] [CrossRef]
  14. Rehman, A.; Lee, S. Review of the Potential of the Ni/Cu Plating Technique for Crystalline Silicon Solar Cells. Materials 2014, 7, 1318–1341. [Google Scholar] [CrossRef]
  15. Kamp, M.; Bartsch, J.; Nold, S.; Retzlaff, M.; Hörteis, M.; Glunz, S.W. Economic evaluation of two-step metallization processes for silicon solar cells. Energy Procedia 2011, 8, 558–564. [Google Scholar] [CrossRef]
  16. Bartsch, J.; Mondon, A.; Schetter, C.; Hörteis, M.; Glunz, S.W. Copper as conducting layer in advanced front side metallization processes for crystalline silicon solar cells, exceeding 20% on printed seed layers. In Proceedings of the 35th IEEE Photovoltaic Specialists Conference, Honolulu, HI, USA, 20–25 June 2010; IEEE: Piscataway, NJ, USA, 2010; Volume 157, p. H942. [Google Scholar]
  17. Flynn, S.; Lennon, A. Copper penetration in laser-doped selective-emitter silicon solar cells with plated nickel barrier layers. Sol. Energy Mater. Sol. Cells 2014, 130, 309–316. [Google Scholar] [CrossRef]
  18. Wu, H.H.; Chen, W.J. Failure Behavior of Nickel Silicide Diffusion Barrier between Electroplating Cu and Textured Si Substrate. Int. J. Electrochem. Sci. 2020, 15, 5277–5286. [Google Scholar] [CrossRef]
  19. Lee, E.J.; Kim, D.S.; Lee, S.H. Ni/Cu metallization for low-cost highefficiency PERC cells. Sol. Energy Mater. Sol. Cells 2022, 74, 65–70. [Google Scholar] [CrossRef]
  20. Tous, L.; Dorp, D.H.; Russell, R.; Das, J.; Aleman, M.; Bender, H.; Meersschaut, J.; Opsomer, K.; Poortmans, J.; Mertens, R. Electroless nickel deposition and silicide formation for advanced front side metallization of industrial silicon solar cells. Energy Procedia 2012, 21, 39–46. [Google Scholar] [CrossRef]
  21. Bhan, S.; Kudielka, H. Ordered bcc-phases at high temperatures in alloys of transition metals and B-subgroup elements. Int. J. Mater. Res. 1978, 69, 333–336. [Google Scholar] [CrossRef]
  22. Tinani, M.; Mueller, A.; Gao, Y.; Irene, E.A. In situ real-time studies of nickel silicide phase formation. J. Vac. Sci. Technol. 2001, 19, 376–383. [Google Scholar] [CrossRef]
  23. Lauwers, A.; Kittl, J.A.; Van Dal, M.; Chamirian, O.; Lindsay, R.; Potter, M.; Demeurisse, C.; Vrancken, C.; Maex, K.; Pagès, X.; et al. Low temperature spike anneal for Ni-silicide formation. Microelectron. Eng. 2004, 76, 303–310. [Google Scholar] [CrossRef]
  24. Geenen, F.; Solano, E.; Jordan-Sweet, J.; Lavoie, C.; Mocuta, C.; Detavernier, C. The influence of alloying on the phase formation sequence of ultra-thin nickel silicide films and on the inheritance of texture. J. Appl. Phys. 2018, 123, 185302–185316. [Google Scholar] [CrossRef]
  25. Christensen, M.; Eyert, V.; Freeman, C.; Wimmer, E.; Jain, A.; Blatchford, J.; Riley, D.; Shaw, J. Formation of nickel-platinum silicides on a silicon substrate: Structure, phase stability, and diffusion from ab initio computations. J. Appl. Phys. 2013, 114, 915–922. [Google Scholar] [CrossRef]
  26. Tran, T.T.; Lavoie, C.; Zhang, Z.; Primetzhofer, D. In-situ nanoscale characterizationof composition and structure during formation of ultrathin nickel silicide. Appl. Surf. Sci. 2021, 536, 147781–147788. [Google Scholar] [CrossRef]
  27. Kluska, S.; Bartsch, J.; Büchler, A.; Cimiotti, G.; Brand, A.A.; Hopman, S.; Glatthaar, M. Electrical and mechanical properties of plated Ni/Cu contacts for Si solar cells. Energy Procedia 2015, 77, 733–743. [Google Scholar] [CrossRef]
  28. Kale, A.S.; Nemeth, W.; Perkins, C.L.; Young, D.; Marshall, A.; Florent, K.; Kurinec, S.K.; Stradins, P. Thermal Stability of Copper-Nickel and Copper-Nickel Silicide Contacts for Crystalline Silicon. ACS Appl. Energy Mater. 2018, 1, 2841–2848. [Google Scholar] [CrossRef]
  29. Rehman, A.; Lee, S.H.; Shin, E.G.; Lee, S.H. Improved adhesion of Ni/Cu/Ag plated contacts with thermally formed nickel silicon interface for C-Si solar cells. Mater. Lett. 2015, 161, 181–184. [Google Scholar] [CrossRef]
  30. Liu, C.M.; Liu, W.L.; Hsieh, S.H.; Tsai, T.K.; Chen, W.J. Interfacial reactions of electroless nickel thin films on silicon. Appl. Surf. Sci. 2005, 243, 259–264. [Google Scholar] [CrossRef]
  31. Houa, A.Y.; Tinga, Y.H.; Taia, K.L.; Huanga, C.Y.; Lub, K.C.; Wu, W.W. Atomic-scale silicidation of low resistivity Ni-Si system through in-situ TEM investigation. Appl. Surf. Sci. 2021, 538, 148129–148137. [Google Scholar] [CrossRef]
  32. Rahman, M.K.; Nemouchi, F.; Chevolleau, T.; Gergaud, P.; Yckache, K. Ni and Ti silicide oxidation for CMOS applications investigated by XRD, XPS and FPP. Mater. Sci. Semicond. Process. 2017, 71, 470–476. [Google Scholar] [CrossRef]
  33. Grunthaner, P.J.; Grunthaner, F.J.; Mayer, J.W. XPS study of the chemical structure of the nickel/silicon interface. J. Vac. Sci. Technol. 1980, 17, 924–929. [Google Scholar] [CrossRef]
  34. Yu, L.; Zhou, H.Q.; Sun, J.Y.; Mishra, I.K.; Luo, D.; Yu, F.; Yu, Y.; Chen, S.; Ren, Z.F. Amorphous NiFe layered double hydroxide nanosheets decorated on 3D nickel phosphide nanoarrays: A hierarchical core–shell electrocatalyst for efficient oxygen evolution. J. Mater. Chem. A 2018, 6, 13619–13623. [Google Scholar] [CrossRef]
  35. Sun, Y.N.; Feldman, A.; Farabaugh, E.N. X-Ray photoelectron spectroscopy of O1s and Si2p lines in films of SiOx formed by electron beam evaporation. Thin Solid Films 1988, 157, 351–360. [Google Scholar] [CrossRef]
  36. Kittl, J.A.; Pawlak, M.A.; Torregiani, C.; Lauwers, A.; Demeurisse, C.; Vrancken, C.; Absil, P.P.; Biesemans, S.; Detavernier, C.; Jordan-Sweet, J.; et al. Kinetics of Ni3Si2 formation in the Ni2Si–NiSi thin film reaction from in situ measurements. Appl. Phys. Lett. 2007, 91, 232102. [Google Scholar] [CrossRef]
  37. Bhaskaran, M.; Sriram, S.; Perova, T.S.; Ermakov, V.; Thorogood, G.J.; Short, K.T.; Holland, A.S. In situ Micro-Raman Analysis and X-ray Diffraction of Nickel Silicide Thin Films on Silicon. Micron 2009, 40, 89–93. [Google Scholar] [CrossRef] [PubMed]
  38. Kale, A.; Beese, E.; Saenz, T.; Warren, E.; Nemeth, W.; Young, D.; Marshall, A.; Florent, K.; Kurinec, S.K.; Agarwal, S.; et al. Study of Nickel Silicide as a Copper Diffusion Barrier in Monocrystalline Silicon Solar Cells. In Proceedings of the 43rd IEEE PVSC, Portland, OR, USA, 5–10 June 2016; pp. 2913–2916. [Google Scholar]
  39. Acker, J.; Bohmhammel, K. Optimization of thermodynamic data of the Ni–Si system. Thermochim. Acta 1997, 337, 187–193. [Google Scholar] [CrossRef]
  40. Shatynski, S.R. The thermochemistry of transition metal carbides. Oxid. Met. 1979, 13, 105–118. [Google Scholar] [CrossRef]
  41. Schlesinger, M.E. Thermodynamics of solid transitionmetal silicides. Chem. Rev. 1990, 90, 607–628. [Google Scholar] [CrossRef]
  42. Lee, S.H.; Lee, D.W.; Rehman, A.U.; Baik, J.W.; Lee, S.H. Study of Annealing Temperature for Ni/Cu/Ag Plated Front Contact Single Crystalline Solar Cells. IEEE J. Photovolt. 2016, 6, 1090–1093. [Google Scholar] [CrossRef]
  43. Guo, S.; Gregory, G.; Gabor, A.M.; Schoenfeld, W.V.; Davis, K.O. Detailed investigation of TLM contact resistance measurements on crystalline silicon solar cells. Sol. Energy 2017, 151, 163–172. [Google Scholar] [CrossRef]
  44. Tous, L.; Dorp, D.H.; Hernandez, J.L.; Allebe, C.; Bender, H.; Meersschaut, J.; Aleman, M.; Russell, R.; Poortmans, J.; Mertens, R. Minimizing junction damage associated with nickel silicide formation for the front side metalization of silicon solar cells. In Proceedings of the 26th European Photovoltaic Solar Energy Conference and Exhibition, Hamburg, Germany, 5–6 September 2011; pp. 1210–1215. [Google Scholar]
  45. Buchler, A.; Kluska, S.; Brand, A.; Geisler, C.; Hopman, S.; Glatthaar, M. Micro characterization and imaging of spikes in nickel plated solar cells. Energy Procedia 2014, 55, 624–632. [Google Scholar] [CrossRef]
  46. Saseendran, S.S.; Raval, M.C.; Solanki, C.S.; Saravanan, S.; Suckow, S.; Kottantharayil, A.; Joshi, A.P. Study of nickel silicide formation and associated fill-factor loss analysis for silicon solar cells with plated Ni-Cu based metallization. IEEE J. Photovolt. 2015, 5, 1554–1562. [Google Scholar]
Figure 1. Flow chart of seed layer electroplating n-TOPCon solar cell.
Figure 1. Flow chart of seed layer electroplating n-TOPCon solar cell.
Materials 17 02610 g001
Figure 2. (a) SEM image and (b) EDS elements of seed layer at annealing temperature of 375 °C; the samples annealed at different temperatures of (c) XPS; (d) XRD; and (e) Raman.
Figure 2. (a) SEM image and (b) EDS elements of seed layer at annealing temperature of 375 °C; the samples annealed at different temperatures of (c) XPS; (d) XRD; and (e) Raman.
Materials 17 02610 g002
Figure 3. The structure diagram of Ni-Si alloy was obtained by different annealing processes. The green ball represents the Ni atom, and the blue ball represents the Si atom.
Figure 3. The structure diagram of Ni-Si alloy was obtained by different annealing processes. The green ball represents the Ni atom, and the blue ball represents the Si atom.
Materials 17 02610 g003
Figure 4. (a) Contact resistance model; under different annealing conditions; the samples under different annealing temperatures (b) contact resistance values (c) surface roughness of sample; (d) the adhesion of busbars.
Figure 4. (a) Contact resistance model; under different annealing conditions; the samples under different annealing temperatures (b) contact resistance values (c) surface roughness of sample; (d) the adhesion of busbars.
Materials 17 02610 g004
Figure 5. Electrical properties under different annealing conditions (a) Rs; (b) FF; and (c) pFF.
Figure 5. Electrical properties under different annealing conditions (a) Rs; (b) FF; and (c) pFF.
Materials 17 02610 g005
Table 1. The electrical properties of the cells under different annealing processes.
Table 1. The electrical properties of the cells under different annealing processes.
SampleJsc (mA cm−2)Voc (mV)FF (%)Eff (%)J02 (A cm−2)
BSL40.77 ± 0.14706.53 ± 1.781.67 ± 0.2123.52 ± 0.07(6.17 ± 0.03) × 10−9
300 °C40.78 ± 0.22706.9 ± 0.781.8 ± 0.2923.58 ± 0.11(6.26 ± 0.02) × 10−9
350 °C40.78 ± 0.32706.95 ± 1.381.92 ± 0.2723.62 ± 0.15(6.08 ± 0.025) × 10−9
375 °C40.82 ± 0.13706.98 ± 1.582.02 ± 0.1823.67 ± 0.12(5.27 ± 0.015) × 10−9
400 °C40.42 ± 0.35705.67 ± 1.481.34 ± 0.1323.43 ± 0.17(10.71 ± 0.02) × 10−9
425 °C39.95 ± 0.46702.67 ± 1.880.78 ± 0.1323.27 ± 0.07(13.74 ± 0.03) × 10−9
Table 2. Comparison of electrical properties before and after stability test.
Table 2. Comparison of electrical properties before and after stability test.
SampleJsc (mA cm−2)Voc (mV)FF (%)Eff (%)Rs (ohm cm2)
BSLBefore40.77706.5381.6723.520.20586
After (4 h)40.72704.8579.8422.910.47449
375 °C annealing temperatureBefore40.81706.8781.9623.650.18636
After (4 h)40.80705.7880.2123.090.44138
After (8 h)40.44702.1278.5622.310.81981
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

Wang, Z.; Liu, H.; Chen, D.; Wang, Z.; Wu, K.; Cheng, G.; Ding, Y.; Zhang, Z.; Chen, Y.; Gao, J.; et al. Enhancing Adhesion and Reducing Ohmic Contact through Nickel–Silicon Alloy Seed Layer in Electroplating Ni/Cu/Ag. Materials 2024, 17, 2610. https://doi.org/10.3390/ma17112610

AMA Style

Wang Z, Liu H, Chen D, Wang Z, Wu K, Cheng G, Ding Y, Zhang Z, Chen Y, Gao J, et al. Enhancing Adhesion and Reducing Ohmic Contact through Nickel–Silicon Alloy Seed Layer in Electroplating Ni/Cu/Ag. Materials. 2024; 17(11):2610. https://doi.org/10.3390/ma17112610

Chicago/Turabian Style

Wang, Zhao, Haixia Liu, Daming Chen, Zigang Wang, Kuiyi Wu, Guanggui Cheng, Yu Ding, Zhuohan Zhang, Yifeng Chen, Jifan Gao, and et al. 2024. "Enhancing Adhesion and Reducing Ohmic Contact through Nickel–Silicon Alloy Seed Layer in Electroplating Ni/Cu/Ag" Materials 17, no. 11: 2610. https://doi.org/10.3390/ma17112610

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