Next Article in Journal
Characterization of an Amorphous Titanium Oxide Film Deposited onto a Nano-Textured Fluorination Surface
Previous Article in Journal
Using Central Composite Experimental Design to Optimize the Degradation of Tylosin from Aqueous Solution by Photo-Fenton Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reactions in Electrodeposited Cu/Sn and Cu/Ni/Sn Nanoscale Multilayers for Interconnects

by
Pay Ying Chia
1,
A. S. M. A. Haseeb
1,* and
Samjid Hassan Mannan
2
1
Department of Mechanical Engineering, Faculty of Engineering, University of Malaya, Kuala Lumpur 50603, Malaysia
2
Physics Department, King’s College London, Strand, London WC2R 2LS, UK
*
Author to whom correspondence should be addressed.
Materials 2016, 9(6), 430; https://doi.org/10.3390/ma9060430
Submission received: 17 March 2016 / Revised: 3 May 2016 / Accepted: 19 May 2016 / Published: 31 May 2016

Abstract

:
Miniaturization of electronic devices has led to the development of 3D IC packages which require ultra-small-scale interconnections. Such small interconnects can be completely converted into Cu-Sn based intermetallic compounds (IMCs) after reflow. In an effort to improve IMC based interconnects, an attempt is made to add Ni to Cu-Sn-based IMCs. Multilayer interconnects consisting of stacks of Cu/Sn/Cu/Sn/Cu or Cu/Ni/Sn/Ni/Sn/Cu/Ni/Sn/Ni/Cu with Ni = 35 nm, 70 nm, and 150 nm were electrodeposited sequentially using copper pyrophosphate, tin methanesulfonic, and nickel Watts baths, respectively. These multilayer interconnects were investigated under room temperature aging conditions and for solid-liquid reactions, where the samples were subjected to 250 °C reflow for 60 s and also 300 °C for 3600 s. The progress of the reaction in the multilayers was monitored by using X-ray Diffraction, Scanning Electron Microscope, and Energy dispersive X-ray Spectroscopy. FIB-milled samples were also prepared for investigation under room temperature aging conditions. Results show that by inserting a 70 nanometres thick Ni layer between copper and tin, premature reaction between Cu and Sn at room temperature can be avoided. During short reflow, the addition of Ni suppresses formation of Cu3Sn IMC. With increasing Ni thickness, Cu consumption is decreased and Ni starts acting as a barrier layer. On the other hand, during long reflow, two types of IMC were found in the Cu/Ni/Sn samples which are the (Cu,Ni)6Sn5 and (Cu,Ni)3Sn, respectively. Details of the reaction sequence and mechanisms are discussed.

1. Introduction

As the electronic industry is moving towards miniaturization, three-dimensional (3-D) IC packages are being aggressively pursued. In 3-D packages, IC chips are stacked on top of each other to save space and to improve performance. In miniaturized packages, through-Si-vias (TSVs) and solder micro-bumps are used as interconnects, the size of which ranges from a few micrometres to about 50 micrometres [1,2]. In such small interconnects, Cu metallization and Sn-rich solder alloys react and transform completely into Cu-Sn intermetallic compounds (IMC), e.g., Cu6Sn5 and Cu3Sn after reflow. Intermetallic compounds are generally known to be brittle and can adversely affect the reliability of the interconnection particularly in applications involving shock load. Therefore, there is a need to investigate ways to improve the properties of Cu-Sn IMC.
One way to address this could be to add a third element to Cu-Sn based IMCs. Out of different metallic elements studied, Ni is suggested to be a good candidate for addition into solders [3,4,5,6]. Studies have shown that Ni can suppress ε-Cu3Sn even when added at a small concentration [3,4,5]. Cu3Sn is considered to be more detrimental to reliability [7]. Furthermore, Ni shows marked solubility in η-Cu6Sn5 IMC and can influence its properties. Nogita and Nishimura [6] reported that with the addition of 9% Ni to the (Cu,Ni)6Sn5 IMC, polymorphic phase transformation of the Cu-Sn IMC from the hexagonal structure to the monoclinic structure can be prevented. Cu-Ni-Sn IMC has been reported to have higher hardness and Young’s modulus [8,9,10,11].
Addition of Ni to interconnects has been done in the form of nanoparticles in the recent past. It has been shown [3] that Ni nanoparticles of ~20 nm size undergoes reactive dissolution during reflow leading to in-situ alloying. It is, therefore, expected that if thin nanometric layers of Ni is introduced in Cu/Sn based interconnects in the form of multilayers, the former can get intermixed with Cu and Sn resulting in Ni alloyed IMC. This work investigates the formation of multilayered interconnects containing layers of Cu, Ni and Sn by electrodeposition. The samples are then subjected to intermixing under different conditions.
Considerable amount of work has been done in the past on the Cu/Sn system, both for solid state reactions and solid-liquid reactions. During solid state reactions, it has been found that Cu6Sn5 IMC forms at room temperature. One study reported that it can form at −2 °C [12]. With the use of Auger electron spectroscopy and W markers, it was found revealed that Cu is the main diffusing species during the formation of η-Cu6Sn5 at room temperature. Chopra and co-workers suggested that at the initial stage, the formation of η-Cu6Sn5 took place in the Sn matrix in the form of precipitates through rapid diffusion of Cu into bulk Sn [13]. However, a recent study by Sobiech and co-workers showed that after five days of room temperature aging, Cu atoms diffuse preferentially along Sn grain boundaries where the η-Cu6Sn5 phase formed first [14]. As has been found in the literature, Cu and Sn can mix readily at room temperature. This can lead to premature formation of IMC even before reflow. This may result in difficulties in obtaining a good interconnection, as Cu and Sn can be consumed completely prior to joint formation, particularly when the interconnections are small as in the case of 3D packages. The addition of the nanometric Ni layers between Cu and Sn as proposed above can help to prevent the premature intermixing before reflow. Ni layers should be thin enough though to undergo complete dissolution during reflow. Little information is available in the literature on the intermixing behaviour in Cu/Ni/Sn multilayered system.
In this work, we introduced ultrathin Ni layers of thickness 35 nm to 150 nm between Cu and Sn layers. The motivation is to prevent premature reaction between Cu and Sn prior to reflow. Another motivation is to form, during reflow, Cu-Sn IMCs alloyed with Ni which are expected to lead to better mechanical properties and hence improved reliability of interconnects. Cu/Sn and Cu/Ni/Sn multilayers system were prepared by electrodeposition. Both systems were investigated in terms of IMC formation after 1 day and 24 days of room temperature aging. The effect of Ni in the Cu/Sn system is also studied after both short and long reflow to investigate solid-liquid state reactions.

2. Results

2.1. Solid State Reactions of Cu/Sn and Cu/Ni/Sn Multilayers

Figure 1 and Figure 2 show FESEM micrographs of FIB milled Cu/Sn and Cu/Ni/Sn samples after room temperature aging for one day and 24 days, respectively. The Cu/Ni/Sn samples used for solid state reactions contain Ni layers with 70 nm thickness. Layers with darker contrast are Sn while that with lighter contrast are Cu (Figure 1a,c). Micrographs at higher magnification (Figure 1b,d) show that the Cu layer contains fine columnar grains with axis perpendicular to the interface. The width of the columnar Cu grain is about 100–200 nm. Each columnar grain contains nanotwins with a twinning plane parallel to the interface. Sn layers (Figure 1b,d) contain more or less equiaxed grains with lateral width of about 1–2 μm. Within the dark grey tin layers, a phase with a lighter grey shade is seen, which corresponds to Cu6Sn5 IMC (Figure 1b).
After the Cu/Sn sample was aged for one day, Cu6Sn5 is seen as thin columns along the Sn grain boundaries (Figure 1a,b). Nanometric voids can be seen between Cu and Cu6Sn5 IMCs layers. In the Cu/Ni/Sn system, IMC was seen to form at the Ni/Sn interface and also along Sn grain boundaries after one day of room temperature aging (Figure 1c,d). However, the extent of IMC formation in the Cu/Ni/Sn sample was smaller when compared to that of the Cu/Sn sample.
After 24 days of room temperature aging, it is observed (Figure 2a) that in the Cu/Sn system, most of the Sn has reacted with Cu to form wider grains of Cu6Sn5 IMC. Some of these grains have merged with neighbouring grains forming blocks of Cu6Sn5 IMC separated from each other by some unreacted Sn. Additionally, a layer of Cu6Sn5 IMC is seen at the Cu/Sn interface (Figure 2a,c). From these observations, it is suggested that Cu6Sn5 IMC grew simultaneously in two directions, depicted by dashed arrow in Figure 2b: (i) it grew along the Sn grain boundary and (ii) grew into bulk Sn parallel to the Cu/Sn interface. Furthermore, voids formed in between Cu and Cu6Sn5 IMC have grown in size (Figure 2b) as compared to those after one day of aging (Figure 1b). In the Cu/Ni/Sn system, it is observed that the extent of formation of IMC is much less. There is still a significant amount of unreacted Sn present in this sample. However, the extent of the growth along the grain boundaries is less in the Cu/Ni/Sn sample as compared with that of the Cu/Sn sample. The diffusion of Cu into Sn through the grain boundary is thought to have been impeded by Ni atoms that may have segregated to the grain boundaries (Figure 2d).
Table 1 shows the number of voids for each range of the void’s diameter for both Cu/Sn and Cu/Ni/Sn samples aged at room temperature. The number of voids was calculated from four micrographs for each condition.
After one day of room temperature aging, the Cu/Sn sample exhibits an average void number of 12 per micrometre along the interface, while the average void number in the Cu/Ni/Sn sample is 11 per micrometre. After 24 days of aging, the average number of voids in the Cu/Sn sample increased to 13.2 per micrometre. However, the number of voids with diameter smaller than 20 nm is found to be 172. In the Cu/Ni/Sn system, the average number of voids is 14.8 per micrometre with the number of voids with diameter smaller than 20 nm is 202. Though there is not much difference in the number of voids formed, fewer voids with larger diameter (voids with diameter bigger than 20 nm) are seen to have formed in the Cu/Ni/Sn sample (Figure 2). The number of voids with diameter smaller than 20 nm is more in the Cu/Ni/Sn sample compared with that in Cu/Sn sample after 24 days of room temperature aging.

2.2. Solid-Liquid State Reactions of Cu/Sn and Cu/Ni/Sn Multilayers

2.2.1. Short Reflow

Figure 3 shows the micrographs for Cu/Sn and Cu/Ni/Sn samples with varying Ni thickness after reflow at 250 °C for 1 min immediately after electrodeposition. Three different Ni thicknesses used in this study are 35 nm, 70 nm and 150 nm. From this point onward, samples with Ni thickness of 35 nm, 70 nm and 150 nm will be designated as Cu/Ni-35/Sn, Cu/Ni-70/Sn, and Cu/Ni-150/Sn, respectively. Figure 3a–d shows cross-sectional FESEM micrographs of Cu/Sn, Cu/Ni-35/Sn, Cu/Ni-70/Sn, and Cu/Ni-150/Sn samples before reflow. The identities of the layers were determined by EDX analysis. The nature of the IMC was found out from the reaction of elements present. In Figure 3a–d, the layers with darker contrast are Cu while that with lighter contrast are Sn. Within the light grey Sn layers, a darker grey phase identified as Cu6Sn5 is seen in Cu/Ni/Sn sample (Figure 3a). As for samples with Ni insertion in between Cu and Sn, no IMC formation is seen. The presence of Ni layers was confirmed with EDX line scans profiles [15]. It may be noted that these samples were stored at room temperature for 5 days prior to taking the micrographs.
Figure 3e–h presents micrographs of Cu/Sn, Cu/Ni-35/Sn, Cu/Ni-70/Sn and Cu/Ni-150/Sn after reflow. In the reflowed Cu/Sn sample (Figure 3e), layers with the darkest contrast correspond to the Cu layer while layers with lightest contrast correspond to Cu6Sn5 phase. The continuous darker grey layers in between Cu and Cu6Sn5 interface and the discontinuous layers in the middle of the Cu6Sn5 IMC correspond to the Cu3Sn phase (Figure 3e). Discontinuous Cu3Sn IMC formed in the middle of the Cu6Sn5 phase due to reaction of the middle layer of copper with its surrounding tin during reflow. As for Figure 3f–h, layers with darker contrast represents Cu while lighter contrast represents the (Cu,Ni)6Sn5 phase. With the addition of Ni, it is seen that (Cu,Ni)6Sn5 IMC was formed regardless of the amount of Ni added. Cu/Ni-35/Sn has been converted completely into (Cu,Ni)6Sn5 (Figure 3f). However, for Cu/Ni-70/Sn and Cu/Ni-150/Sn samples (Figure 3g,h), unreacted copper (darker contrast) in between (Cu,Ni)6Sn5 (lighter contrast), is still left after reflow. The thickness of the unreacted copper increased with increasing Ni thickness. This suggests that Ni starts acting as a barrier layer between Cu and Sn when its thickness increased to 70 nm.

2.2.2. Long Reflow

In order to further understand the growth of Cu-Sn IMC, longer reflow was done for 60 min at 300 °C. Figure 4 shows the FESEM cross-sectional images of Cu/Sn, Cu/Ni-35/Sn, and Cu/Ni-70/Sn after long reflow. Table 2 shows the EDX composition of the IMCs at different spots designated by A–E in Figure 4. In Figure 4a, it is seen that in the Cu/Sn system, all Cu and Sn layers have been transformed into a uniform layer of Cu3Sn. But, with the insertion of 35 nm of Ni, a lighter grey-coloured ribbon-like layer of (Cu,Ni)6Sn5 IMC was found to be mixed with (Cu,Ni)3Sn IMC which has a darker grey contrast. An increase of Ni thickness to 70 nm results in an increase in the amount of (Cu,Ni)6Sn5 (lighter contrast). This sample (Figure 4b) shows a brick-and-mortar type structure with bricks of (Cu,Ni)6Sn5 surrounded by (Cu,Ni)3Sn. Figure 4d–f shows EDX elemental maps of the Cu/Ni-70/Sn sample. Ni, which is indicated by the colour red is observed to segregate more to the light grey colour “brick” region (Figure 4c,f). EDX results in Table 1 shows that the light grey region is (Cu,Ni)6Sn5. Thus, Ni is seen to have a tendency to segregate to the (Cu,Ni)6Sn5 phase. Higher solubility limit of Ni in Cu6Sn5 compared with Cu3Sn is suggested to be the reason for the segregation [16].
Figure 5a shows the XRD patterns for Cu/Sn, Cu/Ni-35/Sn and Cu/Ni-70/Sn after reflow at 300 °C for 60 min. In all cases, Cu peaks from the substrate are observed at 43.41° and 50.55° which correspond to the Cu (111) and Cu (002) planes. Cu6Sn5 peaks are seen at 30.11° and 35.16° which correspond to the (311) and (40 2 ¯ ) planes. Cu3Sn peaks are seen at 27.99°, 32.66°, 37.89°, 41.99°, and 57.83° which corresponds to the (011), (110), (020), (002), and (022) planes, respectively. With the addition of 35 nm of Ni, it can be seen that there is an increase in the intensity of the Cu6Sn5 IMC peak at the diffraction angle of 30.11°. With a further increase of Ni to 70 nm, XRD patterns show Ni peaks at 44.34° and 51.66° corresponding to the (111) and (200) planes. The existence of these Ni peaks suggests that 70 nm of Ni layers are still not completely consumed after 60 min of reflow at 300 °C.
Figure 5b shows the ratio of the peak intensities for Cu3Sn and Cu6Sn5 as a function of Ni layer thickness. The Cu6Sn5 and Cu3Sn peaks used for the calculation are 30.11° and 37.89°, respectively, which are the most prominent peaks for each phase in the XRD pattern. From Figure 5b, it is seen that the intensity ratio decreases which indicates that the amount of Cu3Sn formed decreases when the Ni thickness increased. With increasing Ni in the Cu/Sn system, more Ni atoms substituted the Cu atoms to form more (Cu,Ni)6Sn5 which are more stable than Cu6Sn5 IMC resulting in less transformation from (Cu,Ni)6Sn5 IMC to (Cu,Ni)3Sn IMC [17,18].
Figure 6 shows zoomed-in section of the XRD patterns of samples shown in Figure 5a recorded 29.5°–31° and 37°–38.5° recorded at a slow scan speed. This range of diffraction angles were selected to investigate possible peak shifts of the most prominent peaks of Cu6Sn5 and Cu3Sn IMC. It is seen in Figure 6a that the peak has shifted to the right when Ni atoms substitute Cu atoms in the (Cu,Ni)6Sn5 IMC. This indicates that the (311) interplanar spacing decreases upon Ni addition. This is in good agreement to previous study where it was found that Ni addition to Cu6Sn5 also decreased the lattice parameter [19]. As Ni atom substitutes Cu sites, the unit cell volume shrinks and the distance between the atoms shortens.
Cu3Sn possesses two features, orthorhombic unit cell, and a long period superstructure. The most prominent peak for the long period superstructure of Cu3Sn is at 37.56° for the (2 10 0) plane and that for the orthorhombic structure at 37.84° for the (021) plane are indicated by vertical lines. In the recorded patterns (Figure 6b), Cu3Sn in the Cu/Sn sample shows a peak which is close to the (2 10 0) long period superlattice peak. For the Cu3Sn IMC in the Cu/Ni/Sn sample, the peak tends to split into multiple sub-peaks and shift to higher angles. The peak for Cu/Ni-35/Sn sample is closer to that for Cu/Sn sample and exhibits two sub-peaks. The peaks for Cu/Ni-70/Sn shift closer to (021) Cu3Sn orthorhombic structures. These changes suggest that Ni atoms enter into the Cu3Sn lattice and exert an influence on the structure.

3. Discussion

3.1. IMC Formation during Solid State Reactions

During room temperature aging, copper and tin react to form Cu6Sn5 IMC even after one day of aging. In this present work, it is found that Cu6Sn5 IMC forms simultaneously at two places; (i) at the interface between Cu/Sn, possibly through rapid interstitial diffusion of Cu into bulk Sn and (ii) along the Sn grain boundaries.
Cu6Sn5 formation at the Cu/Sn interface was not observed in the earlier study done by Sobiech and co-workers [14]. One of the reasons for this might be related to the use of a lower resolution microscopy in that study. The formation of this thin layer of Cu6Sn5 at the entire Cu/Sn interface found in this study might be related to the fine copper microstructure. In this study, the copper layer was electrodeposited from alkaline pyrophosphate bath which is known to yield fine grains [20]. This is confirmed in the present study as the copper layers have grains in the range of 100–200 nm. These tiny grains also contain nano-twins as seen in Figure 2b. These finer microstructures can result in favourable nucleation sites for Cu6Sn5. In previous studies [12,13], deposition of Cu and Sn was done by thermal evaporation or electron beam evaporation which might yield a different structure and, thus, might provide fewer nucleation points.
Apart from the formation of Cu6Sn5 at the Cu/Sn interface, Cu also diffuses along Sn grain boundaries resulting in a slender Cu6Sn5 morphology at the grain boundaries (Figure 1a). Similar observations involving Cu diffusing along Sn grain boundaries were reported by Sobiech and co-workers [14,21]. After a complete coverage of Cu6Sn5 at Sn grain boundaries and the Cu/Sn interfaces, it grows by a volume diffusion of Cu into Sn in a direction perpendicular to Sn grain boundaries. This is how the Cu6Sn5 IMC grows wider and thicker after 24 days of room temperature aging.
A schematic of the stages in the formation of Cu6Sn5 IMC in the Cu/Sn system during room temperature aging is shown in Figure 7. The stages involved are (i) formation of a thin layer of Cu6Sn5 in between Cu and Sn and along Sn grain boundaries; (ii) growth of IMC along the Sn grain boundaries and (iii) volume diffusion of Cu into Sn. The unique “hourglass” shape of the Cu6Sn5 IMC (Figure 2b) formed at the grain boundaries is due to the diffusion of Cu from two direction as shown in Figure 7e. The corners situated at the junctions between substrate and grain boundaries receive Cu flux from two directions: (i) from Cu substrate side and (ii) from grain boundary side (Figure 7e). As a result, Cu6Sn5 grows faster here leading to the hourglass shape.
In the case when Ni is introduced between Cu and Sn (Figure 1 and Figure 2), the extent of IMC formation is less. It is seen in Figure 1 that Cu diffusion through the grain boundaries is restricted. Bader et al. has reported on a slower dissolution rate of Ni into Sn solders in comparison to Cu, and thus this can results in less IMC formation in the Ni-containing system [22]. It has been reported that the Cu diffusivity in Sn is 2.5 × 10−7 cm2/s [23] while Ni diffusivity in Sn is 5.4 × 10−9 cm2/s at a temperature above 160 °C [24]. Thus, it is expected that due to the slower diffusivity of Ni, Ni tends to block the diffusion path of Cu at the Sn grain boundaries which leads to less IMC formation.
Under room temperature aging conditions, it is observed that nanometric voids formed even after one day aging. With the increase of aging time to 24 days, the voids at the interface between Cu and Cu6Sn5 become prominent (Figure 2b). It has been reported that the voids can form in the Cu/Sn system due to Kirkendall effects. Kirkendall voids are reported to form at the Cu/Cu3Sn interface during solid state reactions because of the large difference in diffusion fluxes of Cu and Sn [12,25]. In this study, voids formed at the Cu/Cu6Sn5 interface due to the faster Cu diffusion leaving Cu vacancies at the Cu/Cu6Sn5 interface. However, when Ni is introduced into the Cu/Sn system, it is observed that the number of voids is less and the size of the voids is smaller. One reason for this might be the higher Ni diffusion flux in the system which can lead to a lesser difference in diffusion flux of Cu and Ni versus Sn [26].

3.2. IMC Formation during Solid Liquid State Reactions

3.2.1. Dissolution of Ni during Reflow

From Figure 3, it is shown that only (Cu,Ni)6Sn5 IMC was formed after reflow regardless of the thickness of the Ni layer inserted between Cu and Sn layers. This is expected as the amount of Ni inserted into the Cu/Sn system is too low, 2 at %–8 at %, which is calculated from the deposited thickness of the layers. With this amount of Ni in the system, Ni3Sn4, and Ni3Sn2 IMC do not form [27]. The amount of Ni introduced is only sufficient for Ni to substitute Cu in the (Cu,Ni)6Sn5 IMC. However, when the thickness of Ni increases, the amount of unreacted copper increased (Figure 3f–h). It has been reported earlier [3,28] that some metallic nanoparticles including Ni undergo reactive dissolution during reflow. Thus, at smaller thickness, e.g., 35 nm, most of the Ni is expected to dissolve and react to form IMC. At higher thickness e.g., 70 nm, only a part of Ni layer is thought to have been dissolved. The remaining portion of the Ni layer acts as a barrier film. It has been reported that the diffusivity of Cu into Ni is very slow, 2.2 × 10−18 cm2/s at 300 °C [29]. Thus, when the thickness of Ni increased, the diffusion of Cu passing through Ni layers to reach the Sn is slower. This results in more unreacted copper when the thickness of Ni in the system increased.

3.2.2. Effect of Ni Addition in Cu6Sn5 IMC

From the elemental maps in Figure 4, it is observed that Ni preferentially segregates to the (Cu,Ni)6Sn5 phase. EDX results for the (Cu,Ni)6Sn5 formed in the Cu/Ni-70/Sn sample shows that the Ni content in the IMC is 17.85 at %. This is in agreement with results published previously where it was reported that the solubility of Ni in (Cu,Ni)6Sn5 was around 29 at % [4]. The brick-wall morphology structure seen in Figure 4 is suggested to result from the growth of the Cu3Sn phase happening along the grain boundaries of the (Cu,Ni)6Sn5 IMC. Schematic in Figure 8 suggests steps in the growth of (Cu,Ni)3Sn during long reflow. It will be worthwhile to investigate the effect of the brick-wall morphology of the composite, (Cu,Ni)6Sn5 + (Cu,Ni)3Sn on mechanical properties and crack propagation mechanisms.
Peak intensity from XRD patterns (Figure 6) shows an increase for Cu6Sn5 and a decrease for Cu3Sn IMC as the thickness of Ni in the sample increases from 35 nm to 70 nm. This shows that Ni addition to the (Cu,Ni)6Sn5 IMC stabilises the phase and, thus, there is less transformation to (Cu,Ni)3Sn IMC. This is in good agreement to studies done by Yu [18] and Nogita [19]. Nogita reported that 9 at % of Ni in the (Cu,Ni)6Sn5 IMC is sufficient to stabilise the IMC in the hexagonal structure. Yu et al. reported that when Ni substituted Cu atoms to form (Cu,Ni)6Sn5, volume shrinkage happens and the distance between Cu and Sn atoms is shortened. This resulted in much stronger bonds and, thus, a more stable phase.

3.2.3. Substitution of Ni Atoms in Cu Lattice in Cu3Sn IMC

(Cu,Ni)3Sn formed in the Cu/Ni-35/Sn and Cu/Ni-70/Sn samples after long reflow. Though this IMC is rarely reported in solder joints, literature shows that Ni can substitute Cu in Cu3Sn to become (Cu,Ni)3Sn in bulk sample prepared by melting and term aging [16,18,30,31]. Table 3 shows the effect of concentration and heating parameters on Ni concentration in (Cu,Ni)3Sn IMC. Oberndoff [31] has reported on the formation of (Cu,Ni)3Sn after samples containing 40 at % Sn, 35–50 at % Cu and 10–25 at % Ni after annealing at 235 °C for 1100–1700 h. Oberndoff also reported that the solubility of Ni in (Cu,Ni)3Sn can be as high as 3 at %. Lin and co-workers [16] also observed the formation of (Cu,Ni)3Sn when annealing was done at 240 °C for 1695 h on two groups of samples, with compositions (i) 25 at % Sn, 5–70 at % Cu, 5–70 at % Ni and (ii) 40 at % Sn, 30–50 at % Cu, 10–30 at % Ni. Lin et al. reported that when the concentration of Sn was low (25 at %), the amount of Ni in (Cu,Ni)3Sn varied from 5 at % to 70.8 at %. However, when the concentration of Sn was high (40 at %), the amount of Ni in (Cu,Ni)3Sn only ranged from 1.5 at % to 4.3 at %. In the present study, EDX results show that the amount of Ni in the (Cu,Ni)3Sn is 1.87 at % for the Cu/Ni-35/Sn sample and 4.56 at % for the Cu/Ni-70/Sn sample. This is in good agreement with previous studies as the concentration of Sn used in this study is around 40–50 at %. The formation of (Cu,Ni)3Sn IMC found in this study occurs because of the limited Sn supply in the samples used. Past studies used solder balls, solder pastes, or thick layers of Sn solders which can provide an effectively unlimited supply of Sn for the reactions of Cu-Sn IMCs. When the supply of Sn is limited, the amount of Cu6Sn5 IMC formed is limited as well. Since Ni only has a solubility of around 27 at % in (Cu,Ni)6Sn5, Ni atoms will also substitute Cu in the (Cu,Ni)3Sn IMC as the reaction is thermodynamically possible [18].
From the XRD patterns (Y 6), it is seen that peaks for the Cu3Sn phase has shifted to higher angles when Ni substitutes Cu. Substitution of Cu by Zn was studied by van Sande et al. [32]. Addition of Zn to Cu3Sn was found to reduce the long period in the superstructure. The periodicity was also reduced when Ni was added [33]. Further studies are necessary to understand in detail the effect of Ni addition on the structure of Cu3Sn.

4. Materials and Methods

In this study, the multilayer interconnect systems were prepared on commercially pure copper substrates by electrodeposition. Sequential deposition was carried out to prepare both Cu/Sn and Cu/Ni/Sn multilayer interconnects. Cu/Sn multilayers were prepared in the sequence of Cu/Sn/Cu/Sn/Cu while Cu/Ni/Sn multilayers were prepared in the sequence of Cu/Ni/Sn/Ni/Cu/Ni/Sn/Ni/Cu. For the sake of simplicity, Cu/Sn/Cu/Sn/Cu and Cu/Ni/Sn/Ni/Cu/Ni/Sn/Ni/Cu will be designated by Cu/Sn and Cu/Ni/Sn, respectively. Copper deposition in this study, was performed using the alkaline pyrophosphate copper bath (CuSO4·5H2O, 30 g/L; Na4PO2O7, 120 g/L; NH3, 1 mL/L). The bath pH was adjusted to 8.5 by using sulphuric acid for the pyrophosphate copper bath. Watts bath (NiSO4, 150 g/L; NiCl2, 60 g/L; H3BO3, 37.5 g/L) was used for the deposition of nickel. Methanesulfonic acid (MSA)-based bath (SnSO4, 30 g/L; gelatin, 1 g/L; hydroquinone, 5 g/L; MSA, 120 mL/L) was used for the deposition of tin [15,34]. Stirring condition was set at 80 rpm for all baths. Deposition current density was 10 mA/cm2 for copper deposition and 20 mA/cm2 for tin and nickel deposition. The deposition time required to achieve required thickness was estimated using Faraday’s law. Copper was plated to achieve a thickness of 1000 nm while tin was plated to a thickness of 2000 nm. Nickel was plated to a thickness of 35 nm, 70 nm, and 150 nm to study the effect of varying Ni concentration in the Cu/Sn system.
The multi-layered samples were subjected to different processing conditions to study their intermixing and phase transformation characteristics: (i) room temperature aging; (ii) short reflow and (iii) long reflow. Room temperature aging studies were done at 25 °C. Short reflow was done in a FT-02 convection reflow oven at 250 °C while long reflow was done in a tube furnace in 2% H2/N2 atmosphere at 300 °C. FEI Quanta 450 (Houston, TX, USA) Field-Emission Scanning Electron Microscope (FESEM) was used to examine the cross-section of the electrodeposits. The number and size of the voids formed in the Cu/Sn and Cu/Ni/Sn samples were calculated from four micrographs for each condition. Energy dispersive X-ray spectroscopy (EDX, Oxford Instruments, Oxfordshire, UK) was used to determine the composition of the multilayer system. FIB milling was done in a FEI Helios Nanolab 650 Dual Beam FESEM. The phases of the multilayer sample after long reflow were investigated by X-ray Diffraction (PANalytical, Almelo, The Netherlands) using a PanAnalytical diffractometer with Cu Kα radiation which has a wavelength of 1.5418 Å. The step size used was 0.26° and the scan step time was 2.11 s. For slower scans, a step size of 0.01° and scan step time of 8.80 s were used. Peaks shown in the XRD pattern were identified by using the Powder Diffraction File card (JCPDS).

5. Conclusions

  • In the electrodeposited Cu/Sn system, Cu6Sn5 IMC is seen to grow even after one day of room temperature aging, while less Cu-Sn IMC is found in Cu/Ni/Sn system under the same conditions. This finding is important to prevent premature intermixing between the multilayers prior to reflowing during manufacturing process.
  • During solid state reaction in the Cu/Sn system, the growth of the Cu6Sn5 first started through rapid reaction at the Cu/Sn interface followed by grain boundary diffusion of Cu atoms into the Sn grain boundaries. With the addition of Ni in between Cu and Sn, Ni atoms are suggested to block the diffusion path of Cu atoms into Sn grain boundaries and slow down IMC formation.
  • During liquid solid reaction, at 70 nm thickness Ni dissolution is incomplete in the Cu/Ni/Sn multilayers. Thus, to achieve homogeneous IMC layer during short reflow, Ni thickness less than 70 nm is recommended. The addition of Ni suppresses formation of Cu3Sn IMC, regardless of the thickness of the Ni layers.
  • After 60 min of long reflow, Cu/Sn multilayers have been transformed totally into Cu3Sn. In the Cu/Ni/Sn system, Ni atoms take part in the formation of (Cu,Ni)6Sn5 and, thus, stabilizes the IMC and retards transformation into (Cu,Ni)3Sn. Formation of (Cu,Ni)3Sn is suspected to be due to the limited Sn supply in the system.
  • The mechanical properties of the unique “brick-wall” morphology of the [(Cu,Ni)6Sn5 + (Cu,Ni)3Sn] composite formed in the Cu/Ni-70/Sn samples after long reflow would be interesting, as it may influence fracture propagation.

Acknowledgments

The authors would like to acknowledge the financial support from High Impact Research (HIR) Grant, University of Malaya (Project No. UM.C/625/1/HIR/MOHE/ENG/26) and Postgraduate Research Grant, University of Malaya Grant Number (Project No. PG031-2012B).

Author Contributions

P.Y.C. and A.S.M.A.H. conceived and designed the experiments; P.Y.C. performed the experiments; P.Y.C., A.S.M.A.H. and S.H.M. analyzed the data; A.S.M.A.H. contributed reagents/materials/analysis tools; P.Y.C. and A.S.M.A.H. wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Agarwal, R.; Zhang, W.; Limaye, P.; Labie, R.; Dimcic, B.; Phommahaxay, A.; Soussan, P. Cu/Sn Microbumps Interconnect for 3D TSV Chip Stacking. In Proceedings of the Electronic Components and Technology Conference, Las Vegas, NV, USA, 1–4 June 2010.
  2. Yoon, K.S.; Yoo, S.; Shin, Y.S.; Han, M.K.; Lee, C.W. Joint Properties of Solder Capped Copper Pillars for 3D Packaging. In Proceedings of the Electronic Components and Technology Conference, Las Vegas, NV, USA, 1–4 June 2010.
  3. Tay, S.L.; Haseeb, A.S.M.A.; Johan, M.R.; Munroe, P.R.; Quadir, M.Z. Influence of Ni nanoparticle on the morphology and growth of interfacial intermetallic compounds between Sn3.8Ag0.7Cu lead-free solder and copper substrate. Intermetallics 2013, 33, 8–15. [Google Scholar] [CrossRef]
  4. Wang, Y.W.; Chang, C.C.; Kao, C.R. Minimum effective Ni addition to SnAgCu solders for retarding Cu3Sn growth. J. Alloys Compd. 2009, 478, L1–L4. [Google Scholar] [CrossRef]
  5. Gao, F.; Takemoto, T.; Nishikawa, H. Effects of Co and Ni addition on reactive diffusion between Sn–3.5Ag solder and Cu during soldering and annealing. Mater. Sci. Eng. A 2006, 420, 39–46. [Google Scholar] [CrossRef]
  6. Nogita, K.; Nishimura, T. Nickel-stabilized hexagonal (Cu,Ni)6Sn5 in Sn–Cu–Ni lead-free solder alloys. Scr. Mater. 2008, 59, 191–194. [Google Scholar] [CrossRef]
  7. Subramaniam, K.N. Reliability of lead-free electronic solder interconnects: Roles of materials and service parameters. In Lead-Free Solders: Materials Reliability for Electronics; Subramaniam, K.N., Ed.; John Wiley and Sons Ltd.: Chichester, UK, 2012; pp. 281–304. [Google Scholar]
  8. Tsukamoto, H.; Dong, Z.; Huang, H.; Nishimura, T.; Nogita, K. Nanoindentation characterization of intermetallic compounds formed between Sn–Cu (–Ni) ball grid arrays and Cu substrates. Mater. Sci. Eng. B 2009, 164, 44–50. [Google Scholar] [CrossRef]
  9. Xu, L.; Pang, J.H.L. Nano-indentation characterization of Ni–Cu–Sn IMC layer subject to isothermal aging. Thin Solid Films 2006, 504, 362–366. [Google Scholar] [CrossRef]
  10. Jang, G.Y.; Lee, J.W.; Duh, J.G. The nanoindentation characteristics of Cu6Sn5, Cu3Sn, and Ni3Sn4 intermetallic compounds in the solder bump. J. Electron. Mater. 2004, 33, 1103–1110. [Google Scholar] [CrossRef]
  11. Mu, D.; Huang, H.; Mcdonald, S.D.; Nogita, K. Creep and mechanical properties of Cu6Sn5 and (Cu,Ni)6Sn5 at Elevated Temperatures. J. Electron. Mater. 2013, 42, 304–311. [Google Scholar] [CrossRef]
  12. Tu, K.N. Interdiffusion and reaction in bimetallic Cu-Sn thin films. Acta Metall. 1973, 21, 347–354. [Google Scholar] [CrossRef]
  13. Chopra, R.; Ohring, M.; Oswald, R.S. Observation of a marker shift in a Cu-Sn thin film diffusion couple by Auger Electron Spectroscopy methods. Thin Solid Films 1981, 86, 43–47. [Google Scholar]
  14. Sobiech, M.; Krüger, C.; Welzel, U.; Wang, J.Y.; Mittemeijer, E.J.; Hügel, W. Phase formation at the Sn/Cu interface during room temperature aging: Microstructural evolution, whiskering, and interface thermodynamics. J. Mater. Res. 2011, 26, 1482–1493. [Google Scholar] [CrossRef]
  15. Chia, P.Y.; Haseeb, A.S.M.A. Intermixing reactions in electrodeposited Cu/Sn and Cu/Ni/Sn multilayer interconnects during room temperature and high temperature aging. J. Mater. Sci. Mater. Electron. 2015, 26, 294–299. [Google Scholar] [CrossRef]
  16. Lin, C.H.; Chen, S.W.; Wang, C.H. Phase equilibria and solidification properties of Sn-Cu-Ni Alloys. J. Electron. Mater. 2002, 31, 907–915. [Google Scholar] [CrossRef]
  17. Nogita, K. Stabilisation of Cu6Sn5 by Ni in Sn–0.7Cu–0.05Ni lead-free solder alloys. Intermetallics 2010, 18, 145–149. [Google Scholar] [CrossRef]
  18. Yu, C.; Liu, J.; Lu, H.; Li, P.; Chen, J. First-principles investigation of the structural and electronic properties of Cu6–xNixSn5 (x = ¼, 0, 1, 2) intermetallic compounds. Intermetallics 2007, 15, 1471–1478. [Google Scholar] [CrossRef]
  19. Nogita, K.; Mu, D.; McDonald, S.D.; Read, J.; Wu, Y.Q. Effect of Ni on phase stability and thermal expansion of Cu6–xNixSn5 (X = 0, 0.5, 1, 1.5 and 2). Intermetallics 2012, 26, 78–85. [Google Scholar] [CrossRef]
  20. Din, J.W.; Snyder, D.D. Electrodeposition of Cu. In Modern Electroplating; Schlesinger, M., Paunovic, M., Eds.; John Wiley and Sons: Hoboken, NJ, USA, 2010; pp. 33–78. [Google Scholar]
  21. Sobiech, M.; Teufel, J.; Welzel, U.; Mittemeijer, E.J.; Hügel, W. Stress relaxation mechanisms of Sn and SnPb coatings electrodeposited on Cu: Avoidance of whiskering. J. Electron. Mater. 2011, 40, 2300–2313. [Google Scholar] [CrossRef]
  22. Bader, S.; Gust, W.; Hieber, H. Rapid formation of intermetallic compounds by interdiffusion in the Cu-Sn and Ni-Sn systems. Acta Metall. 1995, 43, 329–337. [Google Scholar]
  23. Dyson, B.F.; Anthony, T.R.; Turnbull, D. Interstitial diffusion of copper in tin. J. Appl. Phys. 1967, 38, 3408. [Google Scholar] [CrossRef]
  24. Guo, F.; Lee, J.; Choi, S.; Lucas, J.P.; Bieler, T.R.; Subramaniam, K.N. Processing and aging characteristics of eutectic Sn–3.5Ag solder reinforced with mechanically incorporated Ni particles. J. Electron. Mater. 2001, 30, 1073–1082. [Google Scholar] [CrossRef]
  25. Huang, T.S.; Tseng, H.W.; Hsiao, Y.H.; Cheng, C.H.; Lu, C.T.; Liu, C.Y. Abnormal Cu3Sn Growth and Kirkendall Formation Between Sn and (111) and (220) Preferred-Orientation Cu Substrates. Electrochem. Solid State Lett. 2011, 14, H393–H396. [Google Scholar] [CrossRef]
  26. Yu, H.; Vuorinen, V.; Kivilahti, J.K. Solder/Substrate Interfacial Reactions in the Sn-Cu-Ni Interconnection System. J. Electron. Mater. 2007, 36, 136–146. [Google Scholar] [CrossRef]
  27. Vuorinen, V.; Yu, H.; Laurila, T.; Kivilahti, J.K. Formation of Intermetallic Compounds between Liquid Sn and Various CuNix Metallizations. J. Electron. Mater. 2008, 37, 792–805. [Google Scholar] [CrossRef]
  28. Haseeb, A.S.M.A.; Leong, Y.M.; Arafat, M.M. In-situ alloying of Sn–3.5Ag solder during reflow through Zn nanoparticle addition and its effects on interfacial intermetallic layers. Intermetallics 2014, 54, 86–94. [Google Scholar] [CrossRef]
  29. Abdul-Lettif, A.M. Investigation of interdiffusion in copper-nickel bilayer thin films. Phys. B Condens. Matter 2007, 388, 107–111. [Google Scholar] [CrossRef]
  30. Chen, S.W.; Wu, S.H.; Lee, S.W. Interfacial reactions in the Sn-(Cu)/Ni, Sn-(Ni)/Cu, and Sn/(Cu,Ni) Systems. J. Electron. Mater. 2003, 32, 1188–1194. [Google Scholar] [CrossRef]
  31. Oberndoff, P.J.T.L. Lead-Free Solder System: Phase Relations and Microstructures. Ph.D. Thesis, Eindhoven University of Technology, Eindhoven, The Netherlands, 27 June 2001. [Google Scholar]
  32. Van Sande, M.; De Ridder, R.; van Tendeloo, G.; van Landuyt, J.; Amelinckx, S. A Study by means of Electron Microscopy and Electron Diffraction of Vacancy Ordering in Ternary Alloys of the system AlCuNi. Phys. Status Solidi A 1978, 50, 587–599. [Google Scholar] [CrossRef]
  33. Sang, X.; Du, K.; Ye, H. An ordered structure of Cu3Sn in Cu–Sn alloy investigated by transmission electron microscopy. J. Alloys Compd. 2009, 469, 129–136. [Google Scholar] [CrossRef]
  34. Goh, Y.; Haseeb, A.S.M.A.; Sabri, M.F.M. Effects of hydroquinone and gelatin on the electrodeposition of Sn–Bi low temperature Pb-free solder. Electrochim. Acta 2013, 90, 265–273. [Google Scholar] [CrossRef]
Figure 1. Field Emission Scanning Electron Microscope (FESEM) micrographs showing FIB milled cross sectional images of Cu/Sn (a,b) and Cu/Ni/Sn (c,d) aged at room temperature for one day at 20,000× and 80,000× magnifications, respectively.
Figure 1. Field Emission Scanning Electron Microscope (FESEM) micrographs showing FIB milled cross sectional images of Cu/Sn (a,b) and Cu/Ni/Sn (c,d) aged at room temperature for one day at 20,000× and 80,000× magnifications, respectively.
Materials 09 00430 g001
Figure 2. FESEM micrographs showing FIB milled cross sectional images of Cu/Sn (a,b) and Cu/Ni/Sn (c,d) aged at room temperature for 24 days at 20,000× and 80,000× magnifications, respectively.
Figure 2. FESEM micrographs showing FIB milled cross sectional images of Cu/Sn (a,b) and Cu/Ni/Sn (c,d) aged at room temperature for 24 days at 20,000× and 80,000× magnifications, respectively.
Materials 09 00430 g002
Figure 3. FESEM micrographs showing cross sectional images of (a,e) Cu/Sn; (b,f) Cu/Ni-35/Ni; (c,g) Cu/Ni-70/Ni; and (d,h) Cu/Ni-150/Ni before and after reflow at 250 °C for 60 s, respectively, at a magnification of 10,000×.
Figure 3. FESEM micrographs showing cross sectional images of (a,e) Cu/Sn; (b,f) Cu/Ni-35/Ni; (c,g) Cu/Ni-70/Ni; and (d,h) Cu/Ni-150/Ni before and after reflow at 250 °C for 60 s, respectively, at a magnification of 10,000×.
Materials 09 00430 g003
Figure 4. FESEM micrographs showing cross sectional images of (a) Cu/Sn; (b) Cu/Ni-35/Sn and (c) Cu/Ni-70/Sn and EDX elemental maps of (d) Cu; (e) Sn; and (f) Ni, of Cu/Ni-70/Sn after reflow at 300 °C for 60 min.
Figure 4. FESEM micrographs showing cross sectional images of (a) Cu/Sn; (b) Cu/Ni-35/Sn and (c) Cu/Ni-70/Sn and EDX elemental maps of (d) Cu; (e) Sn; and (f) Ni, of Cu/Ni-70/Sn after reflow at 300 °C for 60 min.
Materials 09 00430 g004
Figure 5. (a) XRD patterns for (i) Cu/Sn; (ii) Cu/Ni-35/Sn and (iii) Cu/Ni-70/Sn after reflowing at 300 °C for 60 min and (b) ratio of Cu3Sn peak intensity to that of Cu6Sn5 peak intensity for Cu/Sn, Cu/Ni-35/Sn, and Cu/Ni-70/Sn after reflow at 300 °C for 60 min.
Figure 5. (a) XRD patterns for (i) Cu/Sn; (ii) Cu/Ni-35/Sn and (iii) Cu/Ni-70/Sn after reflowing at 300 °C for 60 min and (b) ratio of Cu3Sn peak intensity to that of Cu6Sn5 peak intensity for Cu/Sn, Cu/Ni-35/Sn, and Cu/Ni-70/Sn after reflow at 300 °C for 60 min.
Materials 09 00430 g005
Figure 6. XRD patterns for Cu/Sn, Cu/Ni-35/Sn and Cu/Ni-70/Sn with diffraction angle of (a) 2θ = 29.5°–31° for the Cu6Sn5 (311) plane and (b) 2θ = 37°–38.5° for the Cu3Sn phase.
Figure 6. XRD patterns for Cu/Sn, Cu/Ni-35/Sn and Cu/Ni-70/Sn with diffraction angle of (a) 2θ = 29.5°–31° for the Cu6Sn5 (311) plane and (b) 2θ = 37°–38.5° for the Cu3Sn phase.
Materials 09 00430 g006
Figure 7. (ad) Schematic diagram of Cu6Sn5 IMC formation in the Cu/Sn system during room temperature aging and (e) diffusion of Cu atoms into Sn to form Cu6Sn5 IMC.
Figure 7. (ad) Schematic diagram of Cu6Sn5 IMC formation in the Cu/Sn system during room temperature aging and (e) diffusion of Cu atoms into Sn to form Cu6Sn5 IMC.
Materials 09 00430 g007
Figure 8. Schematic of the stage of growth of IMCs after long reflow at 300 °C for 60 min.
Figure 8. Schematic of the stage of growth of IMCs after long reflow at 300 °C for 60 min.
Materials 09 00430 g008
Table 1. Number of voids vs. diameter of voids in Cu/Sn and Cu/Ni/Sn aged at room temperature for 24 days.
Table 1. Number of voids vs. diameter of voids in Cu/Sn and Cu/Ni/Sn aged at room temperature for 24 days.
Diameter of Voids (nm)Number of Voids
CuSnCuNiSn
1 Day24 Days1 Day24 Days
<1035353188
11–20112137110114
21–303122116
31–4091755
41–5002746
>50126510
Table 2. Number of voids vs. diameter of voids in Cu/Sn and Cu/Ni/Sn aged at room temperature for 24 days.
Table 2. Number of voids vs. diameter of voids in Cu/Sn and Cu/Ni/Sn aged at room temperature for 24 days.
SpotComposition (at %)RatioPhase Identified
CuSnNi
A76.2723.73Cu:Sn 3.21:1.00Cu3Sn
B55.2735.129.61(Cu + Ni):Sn 9.28:5.00(Cu,Ni)6Sn5
C72.3525.781.87(Cu + Ni):Sn 2.88:1.00(Cu,Ni)3Sn
D37.7444.4117.85(Cu + Ni):Sn 6.26:5.00(Cu,Ni)6Sn5
E70.0625.384.56(Cu + Ni):Sn 2.94:1.00(Cu,Ni)3Sn
Table 3. Effects of Concentration and Heating Parameters on Ni concentration in (Cu,Ni)3Sn IMC.
Table 3. Effects of Concentration and Heating Parameters on Ni concentration in (Cu,Ni)3Sn IMC.
Concentration (at %)Temperature (°C)Duration (h)Ni Concentration in (Cu,Ni)3Sn (at %)Reference
SnCuNi
4035–5010–252351100–17003[31]
255–705–7024016955–70.8[16]
4030–5010–3024016951.5–4.3[16]

Share and Cite

MDPI and ACS Style

Chia, P.Y.; Haseeb, A.S.M.A.; Mannan, S.H. Reactions in Electrodeposited Cu/Sn and Cu/Ni/Sn Nanoscale Multilayers for Interconnects. Materials 2016, 9, 430. https://doi.org/10.3390/ma9060430

AMA Style

Chia PY, Haseeb ASMA, Mannan SH. Reactions in Electrodeposited Cu/Sn and Cu/Ni/Sn Nanoscale Multilayers for Interconnects. Materials. 2016; 9(6):430. https://doi.org/10.3390/ma9060430

Chicago/Turabian Style

Chia, Pay Ying, A. S. M. A. Haseeb, and Samjid Hassan Mannan. 2016. "Reactions in Electrodeposited Cu/Sn and Cu/Ni/Sn Nanoscale Multilayers for Interconnects" Materials 9, no. 6: 430. https://doi.org/10.3390/ma9060430

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