Paper The following article is Open access

Ohmic contact formation for inkjet-printed nanoparticle copper inks on highly doped GaAs

, , , and

Published 12 March 2021 © 2021 The Author(s). Published by IOP Publishing Ltd
, , Citation Nastaran Hayati-Roodbari et al 2021 Nanotechnology 32 225205 DOI 10.1088/1361-6528/abe902

0957-4484/32/22/225205

Abstract

GaAs compound-based electronics attracted significant interest due to unique properties of GaAs like high electron mobility, high saturated electron velocity and low sensitivity to heat. However, GaAs compound-based electronics demand a significant decrease in their manufacturing costs to be a good competitor in the commercial markets. In this context, copper-based nanoparticle (NP) inks represent one of the most cost-effective metal inks as a proper candidate to be deposited as contact grids on GaAs. In addition, Inkjet-printing, as a low-cost back-end of the line process, is a flexible manufacturing method to deposit copper NP ink on GaAs. These printed copper NP structures need to be uncapped and fused via a sintering method in order to become conductive and form an ohmic contact with low contact resistivity. The main challenge for uncapping a copper-based NP ink is its rapid oxidation potential. Laser sintering, as a fast uncapping method for NPs, reduces the oxidation of uncapped copper. The critical point to combine these two well-known industrial methods of inkjet printing and laser sintering is to adjust the printing features and laser sintering power in a way that as much copper as possible is uncapped resulting in minimum contact resistivity and high conductivity. In this research, copper ink contact grids were deposited on n-doped GaAs by inkjet-printing. The printed copper ink was converted to a copper grid via applying the optimized settings of a picosecond laser. As a result, an ohmic copper on GaAs contact with a low contact resistivity (8 mΩ cm2) was realized successfully.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Among III–V semiconductors, gallium arsenide (GaAs) is widely used in the manufacturing of electronic devices like diodes, integrated circuits, and photovoltaics [1]. The main advantages of GaAs are high electron mobility, high saturated electron velocity and low sensitivity to heat due to its wide bandgap [1]. To render these III–V devices as strong competitors in commercial markets, a reduction of their high manufacturing costs is necessary [2]. Standard contact processes for III–V devices like photolithography, metal evaporation and screen printing are expensive and require complicated processing steps [3].

Significant drawbacks of these traditional front contact methods are given by their high costs resulting from time-consuming and multi-step processes or their specific requirements. Photolithography, for example, uses light to allocate a geometric pattern from a photomask to a photoresist on a substrate and needs several process steps such as coating, curing and stripping. Similarly, metal evaporation frequently needs expensive precursors, and evaporated structures often show poor adhesion of electrodes on the substrate [4]. However, current industrial methods such as screen-printing also require a mask or pattern to realize a printed structure and a higher amount of precursors in comparison to other printing methods.

In order to reduce the front contact processing cost, low-cost back-end of the line (BEOL) processes such as inkjet-printing, combined with novel laser sintering methods for deposition of front electrodes, are promising replacements [5].

Recently, printed electronics and their corresponding technologies have attracted enormous interests in various fields of research and development due to the availability of different conductive nanomaterial inks resulting in the fabrication of low-cost devices via inkjet-printing [69]. Furthermore, inkjet-printing is a low-temperature process and flexible to be utilized on almost any substrate [69].

Essential properties for desirable printing results are good printability, acceptable printing resolution, and the possibility to process the ink according to the substrate surface properties. However, simultaneously the achievement of high electrical conductivity in the range of 108–109 S m−1 is essential for the application of the printed ink as metal contact on semiconductors [10]. Inks suitable for inkjet-printing of front contact grids contain various metals in their nanoparticle (NP) form such as gold (Au), silver (Ag), and copper (Cu). Currently, the most commonly used conductive inks are silver NP-based. Yet, the increasing price of silver acts as a limiting factor for industrial applications [11, 12].

Furthermore, for a low-resistivity ohmic contact and good adhesion of Ag on GaAs, in most cases, the deposition of intermediate metallic barrier or adhesion layers is necessary [13, 14]. As a consequence, less expensive metals which are able to make direct contact on GaAs with low contact resistivity are becoming more and more attractive and demanded alternatives [15]. For the application as contact electrodes, the most promising low-cost candidates to replace Ag inks are copper-based conductive ink formulations with the capability for direct ohmic contact on GaAs [1620].

Since in these inks NPs are coated with polymers or chelated with organic groups in order to prevent agglomeration, a sintering step after printing is necessary to remove the organic fractions. This sintering step uncaps the NPs, induces a merging of the NPs and as a result creates the required conductivity in contact grids. In this context, sintering methods for different NP inks include thermal annealing [21], plasma [22], rapid electrical sintering [23], intense pulsed light sintering [24] and laser sintering [7]. In recent years, intensive research has been performed to synthesize NP inks being able to be reduced and uncapped in situ and by thermal annealing [2527].

For Au and Ag, sintering can be performed via annealing in a conventional convection oven at a temperature range from 180 °C to 300 °C [21]. In contrast to Ag, in the case of Cu, due to its high reactivity, extensive oxidation might happen during sintering processes such as thermal annealing. A high amount of oxidized Cu, however, reduces the conductivity of uncapped Cu NPs. Therefore, in these cases, annealing under an inert atmosphere with a reducing agent is necessary [20, 21].

In this research, the contact of a Cu NP ink is studied on n-doped GaAs surfaces. Thus, the application of a reducing atmosphere (e.g. an acidic vapor atmosphere) with a high probability will alter the GaAs surface, thereby leading to an unfavorable behavior of the contact. As a result, thermal annealing of the Cu NP ink is not applicable, and a laser sintering process is used for the work presented in this study.

Laser sintering is one of the most efficient and fastest sintering methods for Cu NP inks [7]. In this method, NPs absorb energy from a continuous or pulsed laser source. More specific, laser sintering provides high amounts of concentrated energy in a relatively short time (10−12–10−9 s). After transformation of the absorbed laser energy to heat, the released heat causes the NPs to be uncapped and molten. As a consequence Cu NPs do not have enough time to oxidate and merge with their adjacent particles, eventually forming a conductive Cu structure [7].

In this work, we present the application of industrially scalable and low-cost BEOL methods to develop contact electrodes of Cu and Ag NP inks to n-doped GaAs via inkjet-printing combined with respective sintering methods for uncapping the NP ink. The Cu NP and Ag NP inks that we applied for our research are commercially available. For each ink, a suitable sintering method was chosen. In case of the Ag NP ink, thermal annealing in a convection oven is the sintering method recommended by the ink supplier [28]. In contrast to that, a picosecond laser was used to sinter the Cu NP ink and to simultaneously prevent the oxidation of Cu.

2. Results and discussion

2.1. Metal/GaAs contact

One of the main parameters that needs to be evaluated for a metal/semiconductor contact is its contact resistivity. For high performance and consistency of integrated circuits, a low contact resistivity and a small Schottky barrier height at the metal/semiconductor interface are essential. Additionally, these need to be combined with stable contact behavior in terms of chemical structure, adhesion and resistivity. When the metal and the semiconductor are not in contact, their Fermi level energies might differ, which leads to different work functions (Φ) (figure 1(a)). As soon as they are in contact, Fermi levels of metal and semiconductor align at equilibrium. Depending on the work function of metal and n-doped semiconductor, either a Schottky or an ohmic contact can develop. An ideal contact condition for a metal with an n-doped semiconductor occurs when the work function of the metal (ΦM) is smaller than the work function of the semiconductor (ΦS), resulting in an ohmic junction. At equilibrium, electrons move barrier-free between semiconductor and metal (figure 1(b)). For an n-doped semiconductor, when the Fermi level of the semiconductor is higher than the metal and ΦM > ΦS, a Schottky junction is formed. In a Schottky junction, a barrier for electrons to move from metal to semiconductor is formed, which is called the Schottky barrier (φB). By decreasing the work function differences between metal and semiconductor, the depletion region at the interface is narrowed, and the barrier height is reduced, which leads to the tunneling of charge carriers between semiconductor and metal. Therefore, in this case, a semi or quasi ohmic behavior of metal on semiconductor might be observed.

Figure 1.

Figure 1. Junction characteristics of a metal and n-doped semiconductor, (a) before contact (b) after contact with ΦM < ΦS and ΦM > ΦS.

Standard image High-resolution image

Among all metals, Cu according to its respective work function, plus its lower cost compared to Au and Ag, is a proper candidate for contacts on GaAs [15]. In contrast to Cu, up to now Au and Ag are usually applied on GaAs in an alloy form or need an additional interface layer to achieve a low contact resistivity [13, 14]. However, the doping level of GaAs might also lead to a low contact resistivity of Ag on GaAs [29]. N-doping in semiconductors provides free electrons in the outer layer of the valence band region. These electrons need less energy to be lifted from the valence band into the conduction band, than the electrons which cause the semiconductivity in semiconductor. Because of the high Fermi-level pinning in GaAs caused by its surface state density, formation of band bending leads to the formation of a depletion region on the surface of the semiconductor similar to a pn junction. A high doping level (n = 1 × 1018–1 × 1019 cm−3) in GaAs decreases this depletion region and facilitates the movements of charge carriers at the semiconductor-metal interface [30]. In addition to Cu, in this work Ag was depsotited as well as contact material on GaAs in order to investigate the effect of the semiconductor doping level on the electrical contact with Ag.

The energy level alignment of the metal/semiconductor contacts was assessed based on the work function derived from ultraviolet photoelectron spectroscopy (UPS) and the valence region of metals and semiconductor derived from x-ray photoelectron spectroscopy (XPS). Cu and Ag NP inks were deposited on GaAs by inkjet-printing and sintered with required sintering methods. For the proof of concept, both metals were also deposited by electron beam physical vapor deposition (EBPVD), in order to compare the electrical contact of the printed ink to the pure evaporated metals on GaAs.

Figure 2 shows UPS results of the secondary electron cut-off of all samples, which reveals the work function of the metal and the semiconductor. UPS mirrors the broad distribution of states of the conduction band in metal and the valence band of a semiconductor. Therefore, the work functions of metal and semiconductor as well as the energy level alignment at the contact interface can be derived from UPS. XPS of the valence band region of the n-doped GaAs, is shown in figure 2 and illustrates the valence band maximum of the semiconductor. The region of low binding energy in the XPS spectrum is attributed to electrons stemming from the sample surface and thus represents the GaAs valence band in high detail. XPS in that context was selected instead of UPS since a higher penetration depth in the sample can be expected for x-ray photons due to their higher energy compared to UV radiation. Hence, the XPS signal characterizes the actual sample surface better based on a strong signal from the semiconductor with a negligible contribution of residual molecules that were adsorbed on the sample surface during sample preparation under ambient conditions.

Figure 2.

Figure 2. Secondary electron cut-off from UPS spectra (left) and valence region from XPS (right).

Standard image High-resolution image

A highly n-doped GaAs wafer with a doping level of n = 1 × 1019 cm−3 (from Azur SPACE) was cleaned by a standard solvent cleaning method [31] to provide a dust-free and clean surface. Solvent cleaned GaAs shows a work function of Φ = 3.8 eV which is lower than the work function for GaAs reported in literature of Φ = 4.2–4.7 eV [32]. This shift in the work function is predictable due to adsorbed solvent molecules on the surface. However, after printing the ink on GaAs, the printed structures are annealed or sintered, which leads to uncapping of the ink and contact formation at the interface. The applied high temperature required for sintering (higher than 250 °C) might cause partial evaporation of the adsorbed solvent on the GaAs surface. Therefore, due to the high temperature of the sintering process, UPS and XPS studies of an annealed surface of GaAs were necessary. A cleaned and annealed GaAs surface (15 min at 250 °C) shows a Φ of 4.2 eV. Evaporated Cu, and Cu ink feature a work function of 4.4 eV. In comparison, for evaporated Ag Φm is 4.2 eV and Φm is 4.3 eV for Ag ink.

Figure 3 shows the energy level diagram derived from UPS measurements (figure 2) for differently treated GaAs samples, evaporated Cu, laser-sintered Cu ink, annealed Ag ink and evaporated Ag. Cu ink shows a barrier height ΦB of 0.74 eV on solvent-treated and annealed GaAs. On the same surface, Ag ink showed a barrier height of 0.64 eV. Reported values for barrier heights of pure Cu and pure Ag on GaAs range between 0.88 and 0.99 eV [33]. Therefore, achieving an ohmic contact between NP inks and n-doped GaAs surfaces with low contact resistivity can be possible because of a low barrier height of 0.74 or 0.64 eV. Such ohmic behavior is governed by tunneling of charge carriers at narrow depletion regions due to the impurities in NP inks and a high doping level in GaAs [34].

Figure 3.

Figure 3. Energy level diagram for, annealed GaAs, pure evaporated Cu, laser-sintered Cu ink, annealed Ag ink and pure evaporated Ag. VL is vacuum level, Φ is work function, Ef is energy of Fermi level, VBM is valence band maximum, CBM is conduction band minimum.

Standard image High-resolution image

As NP inks are uncapped, residual organic groups may still alter their work function and energy levels. For example, Wang et al investigated the tunability of NP inks' work function by varying the capping groups. Thus, capping molecules which are present in NP inks change the uncapped metal work function leading to a reduction of the barrier height at the metal/semiconductor interface. The literature reported values of 0.88–0.99 eV are presented for pure deposited metals on GaAs and therefore are higher than the values reported for the NP inks [35].

2.2. Selection of conductive inks-electrical behavior

As it is mentioned in section 2.1, a low contact resistivity of the ink on GaAs is a critical factor to provide an ohmic contact. transmission line method (TLM) was used to assess the resistivity and the electrical behavior of the ink on GaAs. Therefore, a series of evaporated and inkjet-printed interdigitated structures of Ag and Cu with channel lengths of 82, 144, 205, 271, 329 and 405 μm were deposited on GaAs in order to measure and interpret their electrical characteristics (current–voltage (IV) measurements) and calculate the corresponding contact resistivity.

As depicted in figures 4(a) and (b), IV characteristics obtained by TLM, for Cu deposited via evaporation as well as Cu ink on GaAs show a close to ohmic contact characteristic. This ohmic behavior shown by evaporated Cu indicates the in principle possibility of an ohmic contact on GaAs for pure Cu. The ohmic behavior of the Cu NP layer can be attributed to a good fusion of Cu NPs at the interface to the semiconductor, induced by the application of the laser sintering process for uncapping of the metal, which enables the NPs to merge well [16].

Figure 4.

Figure 4.  IV characteristics of TLM structures with distances of 82–405 μm for (a) evaporated Cu, (b) sintered Cu layer, (c) evaporated Ag, (d) sintered Ag layer on n-GaAs.

Standard image High-resolution image

In figures 4(c) and (d), printed Ag ink and Ag deposited via EBPVD on GaAs show rectifying behavior of the current flow which indicates Schottky behavior. Since Ag forms an ohmic contact on GaAs usually when deposited as an alloy or in the presence of intermediate layer, this Schottky behavior of sintered Ag and evaporated on GaAs is expected [13, 14].

According to the results derived from UPS, XPS and electrical behavior evaluations, the Cu ink showed a low barrier contact of 0.74 eV on GaAs with a contact resistivity of 10 mΩ cm2 leading to a desired ohmic metal contact on GaAs. Further processing optimizations were investigated to reduce the contact resistivity.

2.3. Sintering method evaluation

The quality of the sintered metal contacts depends on the quality of the printed layer, the sintering conditions and laser fluence applied [7]. The small size of NPs in the range of 20–100 nm enhances the printability of these inks [36]. Additionally, it decreases their melting point to a lower temperature compared to their bulk forms due to melting temperature depression in NPs [37]. Lower melting temperature facilitates the sintering process at a moderate working temperature being less than 300 °C, which very likely reduces related manufacturing costs [38].

In addition to the principle electrical contact formation, the metal/semiconductor contact quality depends on the adhesion and merging of grain boundaries of metals near to the interface to a semiconductor. In that context, Niittynen et al investigated the laser sintering method for the uncapping of Cu NP inks. They discovered that the laser fluence level affects the quality of Cu particles near to the interface of metal and substrate [16]. A similar behavior exists for Ag ink and was investigated by Xiao et al [39]. They showed that different curing temperatures influence the Ag grain boundary formation and that organic NP capping species improve the adhesion of the Ag ink on the substrate. Moreover, ink components and sintering conditions might improve the electrical quality of the sintered metal and reduce the contact resistivity of metal on GaAs.

Thus, in order to evaluate the efficiency of the sintering process, Cu NP structures with defined form and thickness were inkjet-printed on GaAs. Then laser sintering was performed with a picosecond laser to convert the Cu ink into the final desired metallic Cu layer. Laser fluence was varied in order to realize a Cu layer with an optimally low contact resistivity.

To achieve a low contact resistivity at the metal/semiconductor interface and to realize a highly conductive metal layer, an optimal level for the applied fluence needs to be found. Application of this optimized fluence level means exposing the sample to a laser power, which leads to a high amount of uncapped NPs and simultaneously avoids ablation of the printed structure. This amount of uncapped NPs improves the conductivity of the contact grids and reduces the contact resistivity. However, fluence values higher than this optimized fluence value lead to ablation of the ink resulting in a non-ohmic electrical contact.

In order to find the optimal level for the laser fluence, in a first step, the sheet resistance of the sintered Cu layer at different fluences was measured via a four-point probe method to evaluate the conductivity of sintered Cu. The structures were printed on glass slides to omit the influence of the metal/semiconductor contact resistance on sheet resistance. Figure 5 shows the results for the respective sheet resistances of these structures sintered at different fluence levels. It can be seen that with increasing laser fluence the sheet resistance decreases from 0.45 Ω/□ at 1.80 mJ cm−2 to 0.23 Ω/□ at 2.46 mJ cm−2. With further increase of the laser intensity, the sheet resistance increases again (0.26 Ω/□ at 2.83 mJ cm−2).

Figure 5.

Figure 5. Sheet resistance of copper nanoparticles sintered by laser beam irradiation with different fluence values ranging from 1.80 to 2.83 mJ cm−2.

Standard image High-resolution image

The minimum sheet resistance achieved defines the optimum fluence level for the sintering process.

After the printing on glass, in a second step, a series of Cu ink TLM structures, with finger distances of 34, 102, 170, 238, 306 and 408 μm (figure 6), were inkjet-printed and laser-sintered on GaAs with a thickness of 400 nm. To evaluate the contact resistivity of the Cu layer on GaAs, the resistances of these laser-sintered TLM structures were calculated from IV measurements and plotted versus the finger distances (figure 6).

Figure 6.

Figure 6. (a) Interdigitated TLM structures inkjet-printed and laser-sintered on GaAs at fluence of 8.3 mJ cm−2 with a finger distance of 170 μm, (b) resistance versus Distance of TLM structures at same fluence.

Standard image High-resolution image

The contact resistivity then was calculated based on the respective formulae [40] and resulted in a value of 8 mΩ cm2 at a fluence of 8.3 mJ cm−2. This low contact resistivity provides a low-barrier electron transfer between metal and semiconductor. The processing window for fluence values resulting in an effective and efficient laser sintering for 400 nm Cu on GaAs ranges from 3.9 to 8.3 mJ cm−2, resulting in a decrease of contact resistivity from 10 mΩ cm2 for 6.3 mJ cm−2 to 8 mΩ cm2 for 8.3 mJ cm−2. It has to be noted that the optimum fluence level of the NP ink is different on GaAs compared to glass, which is explained by the different heat capacity of these two materials.

SEM images (figure 7) reveal that the increase of the fluence values for sintering the Cu NP ink deposited on GaAs up to the optimal fluence level (below 8.3 mJ cm−2) increases the fusion of NPs. As it is shown in figure 7(d), a fluence level higher than 8.3 mJ cm−2 leads to ablation and deformation of Cu NPs. Fluence levels less than 3.9 mJ cm−2 lead to a Cu layer with less conductivity, high contact resistivity and bad adhesion of the metal on the semiconductor.

Figure 7.

Figure 7. SEM images of laser-sintered Cu on GaAs at different fluence levels, (a) 3.9, (b) 6.3, (c) 8.3, (d) 9.1 mJ cm−2.

Standard image High-resolution image

Apart from that, the thickness of the printed layer obviously strongly influences the electrical behavior of the sintered Cu NP layer. From the IV characteristics of TLM structures shown in figures 8(a) and (b), it can be seen that a 250 nm thick layer of Cu of leads to a Schottky contact behavior compared to an ohmic contact behavior achieved for a layer with a thickness of 400 nm. The main cause for this differing behavior can be explained by SEM images of the surface of the two differently thick contacts (figures 8(c) and 9(d)). It can be seen clearly that at a thickness of 250 nm (figure 8(c)), distinct cracks form in the copper layer. In contrast to that, the copper layer shows a continuous coverage in the case of a thickness of 400 nm (figure 8(d)). Furthermore, the fusion of the NPs seems to be much more pronounced in the case of 400 nm. Thus, for the Cu NP ink presented in this work, a minimum layer thickness of 400 nm can be defined as a requirement for the deposition of a conductive contact grid on GaAs.

Figure 8.

Figure 8.  IV characteristics of TLM structures with distances of 82–405 μm for of (a) 250 nm Cu ink, (b) 400 nm Cu ink, SEM images of laser-sintered Cu on GaAs (c) 250 nm, (c) 400 nm layer thickness.

Standard image High-resolution image
Figure 9.

Figure 9. XPS patterns of (a) Cu 2P3/2 of Cu layer before laser sintering, (b) Cu 2P3/2 of Cu layer before laser sintering, (b) Cu 2P3/2 of Cu layer after laser sintering, (c) C1s of carbon content before laser sintering, (d) C1s of carbon contents after laser sintering.

Standard image High-resolution image

As a final step of evaluation, XPS measurements were performed on the Cu NP layer before and after laser sintering (figure 9). Peak fitting of the corresponding Cu 2p3/2 peak before sintering revealed the contribution of seven different components (figure 9(a)). These contributions can be assigned to metallic Cu bonds as well as to Cu–O and Cu–C bonds in the binding energy region between 933 and 940 eV. Associated satellite peaks were observed in the region between 940 and 945 eV, resulting from different bonds between Cu as well as from NP capping groups in the sample. A metallic Cu peak and Cu2O peaks appear between 933 and 934 eV and are not easily distinguishable due to many different groups in the ink formulation (figure 9(a)). The fitted peak at 935 eV is attributed to CuO. From 935 to 940 eV, a broad peak can be attributed to the bond between Cu and carbon groups.

After laser sintering, XPS shows only Cu metallic bonds in the fitted Cu2p3/2 peak at a binding energy of 933 eV (figure 9(b)) and all of the satellites have disappeared.

The presence of only Cu metallic bonds indicates an efficient uncapping and subsequent merging of the Cu NPs resulting in the conversion to metallic Cu.

The XPS signal that can be attributed to carbon based components before sintering shows the contribution of C–C at 284 eV, Cu–C between 289 and 290 eV as well as C–O peaks at 286 eV for single- and 288 eV for double bonds (figure 9(c)).

After laser sintering, the contribution of C–C bonds at 284 eV is clearly reduced, and no Cu–C peak can be detected. The remaining carbon contribution, stemming from single- and double carbon–oxygen bonds at 286 and 288 eV, is assigned to residual uncapped organic groups on the sample surface (figure 9(d)).

A comparison of the achieved resistance values to those in literature further supports our results on the efficiency of the presented printing and sintering process for the Cu NP ink. With respect to the sheet resistance, for example, Soltani et al applied inkjet printing to deposit a similar commercial Cu NP ink on the surface of silicon wafer [7] and, in contrast to our work, used a continuous wave laser with a wavelength of λ = 808 nm for annealing. The lowest sheet resistance achieved was 1.2 Ω/□ which is five times higher than the optimum sheet resistance of 0.23 Ω/□ presented above. As this commercial ink features a composition similar to the ink presented herein, it is possible to compare these two laser sintering methods directly [7]. Jun et al recently published inkjet printing results on a synthetic NP ink which was uncapped via thermal annealing. The lowest sheet resistance reported was 1 Ω/□ with a resistivity of 1.2 × 10−6 Ω m. By adding some copper complexes to that ink, a lower sheet resistance of 1 × 10−3 Ω/□ with a resistivity of 1.2 10−9 Ω m was realized [27]. In that context, also Kim et al showed that addition of a reducing agent to the NP ink improves the quality of printed copper layer after sintering [26].

In that context, it should be noted that the thickness of deposited conductive thin films considerably influences their resistivity. In particular, layers with a thickness in the nm regime in principle show a higher resistivity than bulk Cu [41, 42], due to a strong dependence of the charge carrier mobility on the crystallinity and nanostructure of polycrystalline materials [41, 42]. With respect to that Kang et al [25], for instance, reported a resistivity of 3.6 × 10−8 Ω m for an inkjet-printed and thermally-sintered reactive Cu ink layer on glass with a thickness of 3.2 μm. In our study, we achieved a resistivity of 1.1 × 10−7 Ω m for a laser-sintered Cu layer with a thickness of 400 nm. For the sake of completeness, bulk Cu shows a resistivity of 1.7 × 10−8 Ω m [25].

Compared to literature at first glance, the value of 8 mΩ cm2 for the achieved contact resistivity seems to be high, in particular when comparing it to reported values for Cu in the range of 10–2–10–4 mΩ cm2 [4347]. However, these low resistivity values have not been realized with Cu NP inks but were found for Cu alloys combined with an additional barrier layer, the latter of which is needed due to the high diffusion rate of Cu in GaAs [15]. In addition, it needs to be noted that being in the nanometer regime the thickness of the printed Cu NP layer combined with impurities stemming from remaining organic groups in the NP Cu ink can increase its contact resistivity [7, 48]. Thus to the best of our knowledge, the presented contact resistivity values of Cu NPs on GaAs belong to the lowest so far reported in the literature for Cu NP inks.

3. Conclusion

In conclusion, a Cu NP ink as a cost-effective contact grid was deposited on n-doped GaAs via a combination of the two well-known BEOL methods Inkjet-printing and laser sintering.

According to the results derived from UPS and XPS, the Cu ink showed a contact barrier on GaAs of 0.74 eV which resulted in a narrow depletion region at the interface and a semi ohmic behavior of Cu on GaAs.

IV characteristics of sintered Cu on n-doped GaAs showed a desired ohmic metal contact on semiconductor, with a low contact resistivity caused by an efficient fusion of laser-sintered Cu at the metal/semiconductor interface.

An optimization of the laser fluence level resulted in a low sheet resistance of 0.23 Ω/□ for sintered Cu layer. With an optimized sintering process, a well-sintered Cu layer featuring the desired low contact resistivity of 8 mΩ cm2 on n-doped GaAs was achieved successfully.

4. Experimental

4.1. Inkjet printing process

In order to deposit the selected Cu Dycotec ink (DM-CUI-5002) Ag Sigma Aldrich ink (Sigma Aldrich 736473) for this experiment and achieve the required results, the following method was determined and carried out in a cleanroom environment (ISO class 6) as follow. 2 × 2 cm pieces of GaAs wafer with a doping level of, were first cleaned by standard wafer cleaning methods with different polar and non-polar solvents [31]. For Cu printed sample, the GaAs wafers were subsequently dip-coated for 60 s in a 5% acidic solution composed of acetic acid + hydrogen peroxide 1:1 and de-ionized H2O and then blown optically dry with an N2 gun. They were then placed on a hot plate for 5 min at 120 °C. Immediately prior to printing, each substrate was again placed on a hot plate for 120 s to ensure minimal H2O on the substrates. For Ag printed samples, only standard cleaning was applied for GaAs wafers. Using a PiXDRO LP50 printing platform, in combination with a FUJIFILM Spectra SE 128 print head, Cu ink and Ag ink were deposited and fully covered each substrate. The print head, as well as printing platform parameters, are presented in table 1. After printing, the substrates were dried in a vacuum oven at 800 mbar below ambient pressure, and at 50 °C for 2 h.

Table 1. Printing platform parameters.

InkCu ink- DM-CUI-5002Ag ink—Sigma Aldrich 736473
Print headSpectra SE 128Spectra SE 128
Waveform (rise/dwell/fall)4/7/4 μs 85 V4/6/4 μs 85 V
Print head temperature (°C)3525
Meniscus vacuum (mbar)−29−26
Resolution (dpi)250400
Number of nozzles used11
Quality factor11
Step size11
Print directrionY (stage)Y (stage)
Stage temperature (°C)4025
Print speed (mm s−1)5050

4.2. Laser process for Cu ink

For laser sintering, a diode-pumped, picosecond laser with a wavelength of 1064 nm was used. The used EdgeWave PX200-P1-GF laser is based on a Nd:YVO4 crystal, has a maximum pulse repetition rate of 50 MHz and reaches a power of up to 100 W. Series of TLM structures were laser-sintered with an adjusted fluence on squares of 500 μ× 500 μm on interdigitated structures with different distances. Table 2 shows the laser sintering parameters.

Table 2. Laser sintering parameters.

ParametersValues
Repetition rate (MHz)50 MHz
Amount of used laser power30–40 W
Galvo Scanner speed (m s−1)0.01
Jump speed (m s−1)0.05
Lateral distance (μm)5
Focus (over optic table) (mm)0

4.3. Annealing conditions for Ag ink

Ag printed samples were dried for 2.5 h at 50 °C and then annealed for one hour at 250 °C in a convection oven.

4.4. XPS and UPS measurements

The XPS and UPS data were determined by using a 'Multiprobe' ultra-high vacuum surface analysis system from Omicron Nanotechnology with a base pressure of around 5 × 10−10 mbar. The excitation source for the XPS was a DAR400 x-ray tube with Al anode and an XM500 quartz crystal monochromator providing an excitation energy of 1486.6 eV (Al Kα-line) with an energy broadening of around 0.1 eV. The photoelectron electron energies were measured with an EA125 5 channel puls-counting channeltron hemispherical electron analyzer with a pass energy of 80 eV for the survey scans and 20 eV for the detailed spectra. The total energy resolution for the detailed scans (excitation source, electron analyzer) was 0.5 eV. For the peak analysis we used the Unifit2017 spectra processing software. The fit procedure was a convolution of Gaussian and Lorentzian peak profile (Voigt profile). The valance band edge was fitted using the square root model with Gaussian broadening. The UPS-spectra were measured in the same vacuum system with the EA125 electron analyzer (1 eV pass energy) and a HIS13 UV-discharge-lamp with an excitation energy of 21.2 eV (He I line).

Acknowledgments

The research leading to these results has received funding by the European Union's Horizon 2020 research and innovation programme under grant agreement no. 727497. Authors acknowledge Azur SPACE for providing the highly n-doped GaAs wafer for this research.

Data availability statement

All data that support the findings of this study are included within the article (and any supplementary files).

Please wait… references are loading.
10.1088/1361-6528/abe902