Open access peer-reviewed chapter

Copper, Iron, and Aluminium Electrochemical Corrosion Rate Dependence on Temperature

Written By

Mykhaylo Viktorovych Yarmolenko

Submitted: 28 March 2021 Reviewed: 03 September 2021 Published: 10 December 2021

DOI: 10.5772/intechopen.100279

From the Edited Volume

Corrosion - Fundamentals and Protection Mechanisms

Edited by Fahmina Zafar, Anujit Ghosal and Eram Sharmin

Chapter metrics overview

401 Chapter Downloads

View Full Metrics

Abstract

Our investigations show that electrochemical corrosion of copper is faster than electrochemical corrosion of aluminium at temperatures below 100°C. Literature data analysis shows that the Al atoms diffuse faster than the Cu atoms at temperatures higher than 475°C, Al-rich intermetallic compounds (IMCs) are formed faster in the Cu-Al system, and the Kirkendall plane shifts towards the Al side. Electrochemical corrosion occurs due to electric current and diffusion. An electronic device working time, for example, depends on the initial copper cover thickness on the aluminium wire, connected to the electronic device, temperature, and volume and dislocation pipe diffusion coefficients, so copper, iron, and aluminium electrochemical corrosion rates are investigated experimentally at room temperature and at temperature 100°C. Intrinsic diffusivities ratios of copper and aluminium at different temperatures and diffusion activation energies in the Cu-Al system are calculated by the proposed methods here using literature experimental data. Dislocation pipe and volume diffusion activation energies of pure iron are calculated separately by earlier proposed methods using literature experimental data. Aluminium dissolved into NaCl solution as the Al3+ ions at room temperature and at temperature 100°C, iron dissolved into NaCl solution as the Fe2+ (not Fe3+) ions at room temperature and at temperature 100°C, copper dissolved into NaCl solution as the Cu+ ions at room temperature, and as the Cu+ and the Cu2+ ions at temperature 100°C. It is found experimentally that copper corrosion is higher than aluminium corrosion, and the ratio of electrochemical corrosion rates, kCu/kAl > 1, decreases with temperature increasing, although iron electrochemical corrosion rate does not depend on temperature below 100°C. It is obvious because the melting point of iron is higher than the melting point of copper or aluminium. It is calculated that copper electrochemical corrosion rate is approximately equal to aluminium electrochemical corrosion at a temperature of about 300°C, so the copper can dissolve into NaCl solution mostly as the Cu2+ ions at a temperature of about 300°C. The ratio of intrinsic diffusivities, DCu/DAl < 1, increases with temperature increasing, and intrinsic diffusivity of aluminium could be approximately equal to intrinsic diffusivity of copper at a temperature of about 460°C.

Keywords

  • electrochemical corrosion
  • metallic coatings
  • electrolysis
  • diffusion
  • intermetallic compounds
  • phases formation kinetics
  • copper
  • aluminium
  • iron
  • Kirkendall-Frenkel porosity
  • Kirkendall shift
  • activation energy

1. Introduction

An Al wire coated with a thin Cu cover (≈15-μm thickness), utilised near an automobile motor, is heated to temperatures about 373–473 K (100–200°C). Intermetallics (IMCs) can be formed at the Cu/Al interface and grow gradually during heating at such temperatures. The IMC layers are brittle and have high resistivity. Therefore, for assurance of the reliability of the product, information on the growth behaviour of the IMC layers during heating is essentially important [1]. Figure 1 shows the problem: an electronic device working time, t0, depends on initial Cu cover thickness, XCu, and temperature. The electric conductivity of copper is higher than the electric conductivity of aluminium in approximately two times, but the formation of intermetallic phases induces a significant increase in contact resistance, which is found to increase linearly with the thickness of the intermetallics formed [2]. The temperature range used to produce the intermetallic phases was from 250 to 515°C. Moreover, the presence of an electrical field greatly accelerated the kinetics of formation of intermetallic phases and altered significantly their morphology, and the impaired mechanical integrity of the Al-Cu bimetallic joints treated by an electrical current was clearly demonstrated by an extensive cracking not only across the whole intermetallic bandwidth but also within different phases and at a neighbouring interface [2]. Three-phase thickness, X123, can be estimated in such a way. The mass conservation law gives

Figure 1.

(a) Initial stage (t = 0): An electronic device is working, since the electric current, Ia = ICu + IAl, has optimal value; (b) final stage (t = t0): The electronic device is not working, because the electric current, Ib = I3 + I2 + I1 + IAl, has too small value, since pure Cu cover has disappeared.

XCut=01=99+4X3t0+11+1X2t0+11+2X3t01.526X12330.509X123,andX1232XCu,E1

so three-phase general thickness is approximately greater in two times than the initial Cu cover thickness.

Otherwise, it was proved experimentally that a thin Al pad (≈1-μm thickness) can prevent gold and copper corrosion because the intermetallics formation rate in Au-Al system is much higher than the intermetallic formation rate in Cu-Al system, so it is possible to use Cu instead of Au for wire bonding in microelectronic packaging, and Cu has higher electric conductivity, higher thermal conduction, and lower material cost than Au [3]. Corrosion and intermetallic rate formation in gold and copper wire bonding in microelectronics packaging were investigated in [3] at temperatures T1 = 175°С, T2 = 200°С, and T3 = 225°С during 120, 240, 360, and 480 h. The authors have reported that cross-sectional analysis of the Cu ball on Al pad confirmed that corrosion occurred at temperatures about T = 200°С primarily beneath the Cu balls and did not initiate from the Al pad, formation of CuCl2 did not allow self-passivation of Cu to occur, so the rate of copper corrosion increased, and the rate of Cu-Al intermetallics formation was found to be three to five times slower than Au-Al intermetallic formation at all three annealing temperatures. So, copper dissolved into NaCl solution as Cu2+ ions at temperatures about T = 200°С, as we expected. They did not investigate corrosion rate dependence of copper and aluminium on temperature. Moreover, phase layers general thicknesses for Cu-Al system were calculated [3]:

X1232=K123t+K01=K0eQ RT+K01=3.52104μm2se25500Jmol1RTt+0.44μm2,E2

where R ≈ 8.314 JK−1 is the gas constant, and K01 is the constant related to initial IMC thickness. General reaction rates of IMC formation were calculated: K123(T1) = 3.57·10−7 μm2/s, K123(T2) = 6.26·10−7 μm2/s, and K123(T3) = 7.15·10−7 μm2/s. The pre-exponential factor and IMC formation activation energy was calculated: K0 ≈ 3.52·10−4 μm2/s, Q ≈ 25.5 kJ/mol. We can use these results to calculate the electronic device working time by Eq. (14) in [4] at different temperatures:

t0XCu2C32K123=16981XCu2K1232XCu2K0eQRT5900XCu2μm2e25.5kJmol1RTs;E3
t0T1=175oC=448K5900225e255008.314448s40years;t0T2=200oC28years;
t0T3=225oC21years;t0T4=300oC9years;t0T5=350oC6years.

Other researchers have obtained [5]: K123(T4 = 300°C) = 4.2·10−4 μm2/s, K123(T5 = 350°C) = 3.4·10−3 μm2/s. Eq. (3) gives

t0T4=300oC2XCu2K12312days;t0T5=350oC2XCu2K1231.5days.

We can calculate: Q ≈ 124 kJ/mol; K0 ≈ 8.5·10−5 m2/s = 8.5·107 μm2/s;

t0T=175oC5.3106e124Jmol1448Rs49years;t0T=200oC8.4years;t0T=225oC2XCu2K0eQRT5.3106e124Jmol1498Rs1.7years.

It was reported in [1] that the growth of layer 1 is controlled predominantly by boundary diffusion, but that of layers 2 and 3 are governed mainly by volume diffusion at temperatures T = 483–543 K (210–270°C) for various periods up to 3.456Ms (960 h). The authors obtained: K01 ≈ 5.3·10−7 m2/s, Q1 ≈ 86 kJ/mol, K023 ≈ 4.2·10−5 m2/s, Q23 ≈ 146 kJ/mol. We can calculate: K123(T6 = 210°C) = 1.5·10−6 μm2/s,

t0T=210oC2XCu2K1239.6years.

The less temperature is, the higher contribution of grain-boundary diffusion and dislocation pipe diffusion to the layers growth is, so models of grain-boundary diffusion and dislocation pipe diffusion involving outflow into volume should be taken into account [6, 7, 8].

Diffusion activation energy of Al is less than diffusion activation energy of Cu (QAl < QCu) at temperatures from 160–250°C for mutual diffusion in copper-aluminium thin film double layers, but the pre-exponential factors are different in 10 times [9]:

DAl=4105e121kJmol1/RTm2/s,DCu=9.5104e135kJmol1/RTm2/s,E4

in θ-phase (phase 1) CuAl2, CAl = 2/3 ≈ 0.67, CCu = 1/3 ≈ 0.33;

DAl=1.51011e68kJmol1/RTm2/s,DCu=1106e106kJmol1/RTm2/s,E5

in η2-phase (phase 2) CuAl, CAl = CCu = 1/2 = 0.5;

DAl=1.7107e116kJmol1/RTm2/s,DCu=2.4106e125kJmol1/RTm2/s,E6

in γ2-phase (phase 3) Cu9Al4, CAl = 4/13 ≈ 0.31, CCu = 9/13 ≈ 0.69.

We can calculate the mutual diffusion coefficient for each phase at temperature 160°C by the Darken equation [8, 10] and taking into account Eqs. (4)(6):

Di=CAlDCu+CCuDAl;i=1,2,3;D1=6.641020m2/s;D2=1.31019m2/s;D3=1.81021m2/s.

We can calculate using the methods described in [4, 11]: K123(T7 = 160°C) ≈ 2.8·10−6 μm2/s, t0T=160oC2XCu2K1235years, so the problem remains unsolved.

It was founded experimentally, that copper electrochemical corrosion is higher than aluminium electrochemical corrosion in approximately two times at room temperature [4, 11], so a thin Al layer can prevent copper electrochemical corrosion. It was reported also about the influence of hydrogen and the absence of a passive layer on the corrosive properties of aluminium alloys [12].

Besides, the soldered copper/tin-based contacts are the weakest part of the chip that can be related to intermetallics and the Kirkendall-Frenkel porosity formation in the contact zone [13]. One of the most common reasons for chip failure is the soldered. The typical range of packaging and operation of the integrated circuits is from room temperature to 250°C [14].

Hydrostatic pressure of Argon gas (≈10 MPa) can decrease Kirkendal-Frenkel porosity formation, but practically cannot decrease mutual diffusion coefficients, but hot isostatic pressing (p ≈ 100 MPa, Argon) removes porosity due to homogenisation heat treatment in alloy CMSX4 and superalloy CMSX10 [15].

It was clarified also that carbon steel-stainless steel with the environment of flowing sodium chloride does indeed produce synergetic corrosion instead of antagonistic corrosion [16].

Electric current can destruct wire bonding in microelectronics packaging, so we planned to investigate copper, iron, and aluminium electrochemical corrosion at room temperature and temperature 100°C. Direct current can dissolve metal anode into electrolyte, and we planned to do experiments under the same conditions: initial radii of Cu, Fe, and Al anodes should be approximately equal, electrolyte concentration should be the same, anodes lengths immersed into electrolyte should be equal, graphite cathodes should be the same, direct electric current value should be practically the same. Aluminium can dissolve into electrolyte only as of the Al3+ ions, so the charge of aluminium ions should be exactly equal to 3, but copper can dissolve into electrolyte as the Cu+ ions and the Cu2+ ions, and the charge of copper ions could be equal to 1 or 2, and iron can dissolve into electrolyte as the Fe2+ ions and the Fe3+ ions, and the charge of iron ions could be equal to 2 or 3. We need to find appropriate mathematical equations to calculate the charges of copper, iron, and aluminium ions dissolved into NaCl solution.

Advertisement

2. Experimental results of copper, iron, and aluminium electrochemical corrosion

2.1 Investigation at room temperature

Cylindrical anodes (99.99% Cu, 99.96% Fe, and 99.99% Al) were used for copper and aluminium [4, 11], and also iron electrochemical corrosion investigation. Sodium chloride (NaCl) solution was used as an electrolyte (Figure 2).

Figure 2.

Scheme of experimental equipment at room temperature. Cu, Fe, and Al anodes dissolve into NaCl solution as Cu+, Fe2+, and Al3+ ions.

Direct electric current and anode mass decreasing were measured. First of all, we need to be assured that the Cu+ ions (or the Cu2+), the Fe2+ (or the Fe3+), and the Al3+ were present in NaCl solution. The rate of anode dissolving into electrolyte can be calculated using Faraday’s law of electrolysis:

dmdt=MIzF,dm=ρLπdR2t.E7

Here, m is anode mass dissolved into the electrolyte, t is a time of the experiment, M is molar mass, I is the direct electric current value, F is the Faraday constant (F ≈ 96,500 C mol−1), z is a charge of ions, R is anode radius, L is anode length immersed into the electrolyte. Electric current value did not change, so one can calculate:

z=MItFπρLR2t=0R2t,E8

where ρ is anode density. Charges of copper, iron, and aluminium ions were calculated:

zCu=63.55103kgmol2.8A1.2103sFπ8.9103kgm3LCuRCu2t=0RCu2t40.9951,E9
zAl=27103kgmol3.1A1.2103sFπ2.7103kgm3LAlRAl2t=0RAl2t42.9543,E10
zFe=55.847103kgmol3.15A1.2103sFπ7.86103kgm3LFeRFe2t=0RFe2t42.032,E11

where LCu ≈ LAl ≈ LFe: LCu = LFe = 5·10−2 m, LAl = 4.5·10−2 m; R0Cu = R0Al = 2.8 mm, R0Fe = 2.98 mm; IAl ≈ ICu ≈ IFe: IFe = 3.15A, IAl = 3.1A, ICu = 2.8A, so copper dissolved into NaCl solution as the Cu+ ions, iron dissolved into NaCl solution as the Fe2+ ions, and aluminium dissolved into NaCl solution as the Al3+ ions. Anodes radii-decreasing kinetics is shown in Figure 3. Experiments were carried during t1 = 5 min, t2 = 10 min, t3 = 15 min, and t4 = 20 min. Experimental results are as follows: R1Cu = 2.74 mm, R2Cu = 2.67 mm, R3Cu = 2.59 mm, R4Cu = 2.5 mm; R1Al = 2.77 mm, R2Al = 2.73 mm, R3Al = 2.68 mm, R4Al = 2.62 mm, R1Fe = 2.95 mm, R2Fe = 2.92 mm, R3Fe = 2.88 mm, R4Fe = 2.83 mm. Measurement precision was 0.01 mm or 10 micrometres.

Figure 3.

Cu, Fe, and Al anodes radii decreasing kinetics at room temperature.

Chemical reactions took place near the positive electrode (anode):

Cu0e=Cu+,Al03e=Al3+,Fe02e=Fe2+;
Cu++Cl=CuCl,Al3++3Cl=AlCl3,Fe2++2Cl=FeCl2;E12
Cle=Cl0,Cl0+Cl0=Cl2.

Chlorine gas was formed near the anode.

Chemical reactions took place near the negative electrode (cathode):

Na+e=Na0,2Na+2H2O=2NaOH+H2.E13

Hydrogen gas was formed near the cathode.

Anodes radii decreasing rate constants can be calculated as the average value of four experiments to increase calculation precise:

kCu=4R02i=14Ri2i=14ti1.25109m2s,kAl=4R02i=14Ri2i=14ti7.291010m2s,E14
kFe=4R0Fe2i=14RiFe2i=14ti7.261010m2s,kCu≈ 1.71kAl;kAl≈ kFe,

so copper electrochemical corrosion is much higher than aluminium and iron electrochemical corrosion, despite IFe ≈ IAl ≥ ICu: IFe ≈ IAl ≈ 1.1ICu. It needs to point out that kCu, kFe, and kAl have dimensionalities as diffusion coefficients [m2/s], because electrochemical corrosion occurs through anodes’ surface.

2.2 Investigation at temperature 100°C

Experiments were carried also at temperature 100°C. Cylindrical anodes (99.99% Cu, 99.99% Al, and 99.96% Fe) were used for copper and aluminium [17] and also iron electric corrosion investigation. Sodium chloride (NaCl) solution was used as an electrolyte (Figure 4). Direct electric current and anodes’ mass decreasing rate were measured (Figure 5).

Figure 4.

Scheme of experimental equipment at T = 100°C. Cu, Fe, and Al anodes dissolved into NaCl solution as Cu+, cu2+, Fe2+, and Al3+ ions.

Figure 5.

Cu, Fe, and Al anodes radii decreasing kinetics at T = 100°C.

Electric current value did not change, so one can calculate the following:

zCu=63.55103kgmol3.05A1.2103sFπ8.9103kgm3LCuRCu2t=0RCu2t4≈ 1.471+22,E15
zAl=27103kgmol3.15A1.2103sFπ2.7103kgm3LAlRAl2t=0RAl2t4≈ 2.853,E16
zFe=55.847103kgmol3.15A1.2103sFπ7.86103kgm3LFeRFe2t=0RFe2t4≈ 2.012,E17

where LCu = LAl = 4·10−2 m, LFe = 5·10−2 m, R0Cu = 2.27 mm, R0Al = 2.6 mm, R0Fe = 2.83 mm, IAl = 3.15 A, IFe = 3.13 A, ICu = 3.05 A, so copper dissolved into NaCl solution as Cu+ and Cu2+ ions (copper dissolved into NaCl solution as Cu+ ions at room temperature), iron dissolved into NaCl solution as the Fe2+ ions (as at room temperature), and aluminium dissolved into NaCl solution as Al3+ ions (as at room temperature). Anode radii-decreasing kinetics is shown in Figure 5. Experiments were carried during t1 = 5 min, t2 = 10 min, t3 = 15 min, and t4 = 20 min. Experimental results are as follows: R1Cu = 2.2 mm, R2Cu = 2.12 mm, R3Cu = 2.03 mm, R4Cu = 1.92 mm; R1Al = 2.56 mm, R2Al = 2.51 mm, R3Al = 2.45 mm, R4Al = 2.38 mm; R1Fe = 2.80 mm, R2Fe = 2.76 mm, R3Fe = 2.72 mm, R4Fe = 2.67 mm. Measurement precision was 0.01 mm or 10 micrometres. We carried additional experiments, but the result was the same.

Chemical reactions are more complicated at 100°C than at room temperature near positive electrodes (anodes):

Cu0e=Cu+,Cu02e=Cu2+,Al03e=Al3+,Fe02e=Fe2+;
Cu++Cl=CuCl,Al3++3Cl=AlCl3,Fe2++2Cl=FeCl2;E18
Cu2++2Cl=CuCl2,Cle=Cl0,Cl0+Cl0=Cl2.

Chlorine gas and boiling water were formed near anodes.

Chemical reactions took place near negative electrodes (cathodes):

Na+e=Na0,2Na+2H2O=2NaOH+H2.E19

Hydrogen gas and boiling water were formed near cathodes.

Anode radii-decreasing rate constants can be calculated as average value of four experiments to increase calculation precise:

kCu=4R0Cu2i=14RiCu2i=14ti1.154109m2/s,1.25.109atroom temperature,
kAl=4R0Al2i=14RiAl2i=14ti8.421010m2/s,7.29.1010atroom temperature,E20
kFe=4R0Fe2i=14RiFe2i=14ti6.831010m2s,7.23.1010atroom temperature,
kCu1.37kAl,1.72atroom temperatureT127°C,

so copper electrochemical corrosion is higher at room temperature T1 ≈ 27°C, aluminium electrochemical corrosion is higher at temperature T2 = 100°C, and the ratio of electrochemical corrosion rates, kCu/kAl, decreases with temperature increasing, although iron electrochemical corrosion rate practically does not depend on temperature below 100°C. It is obvious, because of the higher melting point of iron than the melting point of copper or aluminium. We can conclude that the Cu2+ ions are less mobile than Cu+ ions. It needs to point out that kCu, kAl, and kFe have dimensionalities as diffusion coefficients, D*Cu, D*Al, D*Fe [m2/s], because electrochemical corrosion occurs through anodes’ surface.

Dislocation pipe and volume diffusion activation energies can be calculated in such a way. The Arrhenius law is valid for dislocation pipe diffusion and volume diffusion in ultra-high-purity samples [8, 18]:

Dd=D0deQdRTorDd=D0deEdkBT,andDV=D0VeQVRTorDV=D0VeEVkBT,E21
QJ/mol=FEeV.E22

Here, R ≈ 8.314 JK−1 is the gas constant, kB is the Boltzmann constant, Qd(Ed) is the dislocation pipe diffusion activation energy (Qd = FEd), F ≈ 96,500 Cmol−1 is the Faraday constant, QV(EV) is the volume diffusion activation energy (QV = FEV), D0d and D0V are the pre-exponential factors, T is the absolute temperature.

Our experimental results allow us to calculate:

kCukAlTDCuDAlT=D0CuD0AleQAlQCuRT;lnD0CuD0Al=0.6;QAlQCu=2.9kJ/mol,E23
D0CuD0Al=0.55,DCuDAlT3=1T3=2900J/mol0.6R583K310oC,E24

so diffusion activation energy of Al, QAl, is higher than the diffusion activation energy of Cu, QCu, (QAl > QCu, QAl-QCu = 2.9 kJ/mol), at temperatures from 20 to 100°C, because the Cu+ ions have higher mobilities than the Al3+ ions, and copper electrochemical corrosion rate can be approximately equal to aluminium electrochemical corrosion at temperature about T3 ≈ 300°C due to the Cu2+ ions are less mobile than the Cu+ ions. Moreover, the pre-exponential factors are approximately the same: D*0Al ≈ 2D*0Cu.

Advertisement

3. Intrinsic diffusivities ratio and diffusion activation energy calculations

3.1 Intrinsic diffusivities ratio of Cu and Al analysis

We can analyse described the experimental results in the Al-Cu system for bulk samples [19] since the ratio D*Al/D*Cu was not calculated in [19]:

DCuDAlj=1NXjXK1Ciπj=1NXj+CiXKπ<1,Ci=CAl,E25

where N is formed phases quantity, Xj is phase j’s thickness, Ci is the average concentration of aluminium in phase i, and XK is the Kirkendall shift length.

Five phases are formed in the Al-Cu system at temperatures from 400 to 535°C in bulk samples [19]: θ-phase (phase 1) CuAl2 (C1 = 2/3), η2-phase (phase 2) CuAl (C2 = 1/2), ζ2-phase (phase 3b) Cu4Al3 (C3b = 3/7), δ-phase (phase 3a) Cu3Al2 (C3a = 2/5), and γ2-phase (phase 3) Cu9Al4 (C3 = 4/13 ≈ 0.31 ≈ 1/3, C = CAl). Inert markers were in δ-phase (phase 3a) Cu3Al2 (C3a = 2/5 = 0.4) and moved to Al side during mutual diffusion. In general, inert markers move to the faster diffusivity component side. We can calculate (C3a = 0.4):

DСuDAlT1=535oCX1+X2+X3b+X3a+X3XK0.6πX1+X2+X3b+X3a+X3+0.4XKπ0.814;y1=DAlDCu=10.8141.228,E26

T1 = 535°C = 808 K, t = 40 h, XK ≈ 20.5 μm, X1 + X2 + X3b + X3a + X3 ≈ 180 μm;

DСuDAlT2=515oCX1+X2+X3b+X3a+X3XK1C3aπX1+X2+X3b+X3a+X3+C3aXKπ0.856;y2=DAlDCu1.168,E27

T2 = 515°C = 788 K, t = 40 h, XK ≈ 11 μm, X1 + X2 + X3b + X3a + X3 ≈ 127 μm);

DСuDAlT3=495oCX1+X2+X3b+X3a+X3XK0.6πX1+X2+X3b+X3a+X3+0.4XKπ0.916;y3=10.9161.092,E28

T3 = 495°C = 768 K, t = 40 h, XK ≈ 5 μm, X1 + X2 + X3b + X3a + X3 ≈ 101 μm;

DСuDAlT4=475oCX1+X2+X3b+X3a+X3XK10.4πX1+X2+X3b+X3a+X3+0.4CXKπ≈ 0.969;y4=DAlDCu≈ 1.032,E29

T4 = 475°C = 748 K, t = 90 h, XK ≈ 2 μm, X1 + X2 + X3b + X3a + X3  113 μm.

We can use these four points to calculate by the least square method to increase calculation precise:

ΔQ=QAlQCu=4j=141000RTjlnyjj=14lnyjj=141000RTj4j=141000RTj2j=141000RTj213.4kJ/mol,E30
y0=expj=141000RTj2j=14lnyjj=141000RTjj=141000RTjlnyj4j=141000RTj2j=141000RTj2exp2.29,E31
DAlDCuT=D0AlD0CueQAlQCuRT9e13.4kJmol1RT,E32
DAlDCuT5=y5=1T5=13400J/molRln9733K460oC,E33

so QAl < QCu (QAl-QCu = −13.4 kJ/mol) because the Cu2+ ions have less mobilities than the Al3+ ions, and we can conclude that the Kirkendall displacement changes sign at a temperature about T5 ≈ 460°C for bulk samples. The pre-exponential factors are different in nine times: D*0Al ≈ 9D*0Cu.

Diffusion activation energy of Al is less than the diffusion activation energy of Cu (QAl < QCu) at temperatures from 160–250°C for mutual diffusion in copper-aluminium thin film double layers, but the pre-exponential factors are different in 10 times [9]. Isolated W islands, 150 Å in diameter, have been deposited between Cu and Al thin film double layers to serve as inert diffusion markers. Marker displacements have been measured. We can calculate the ratio D*Al/D*Cu for each phase at different temperatures:

DCuDAlT=D0CuD0AleQAlQCuRT=24e14kJmol1RTinθ-phasephase1CuAl2,CAl=2/30.67,
D1CuD1AlT=250oC=523K=24e14000Jmol18.314x52324e3.220.96,
D1CuD1AlT=160oC=433K24e3.890.49,

so the Al atoms diffuse faster than the Cu atoms in θ-phase at temperatures from 160 to 250°C;

DCuDAlT=D0CuD0AleQAlQCuRT=7104e38kJmol1RTinη2-phasephase2CuAl,CAl=1/2=0.5,
D2CuD2AlT=250oC=7104e38000Jmol1RT11.2,D2CuD2AlT=160oC1.8,

so the Cu atoms diffuse faster than the Al atoms in η2-phase at temperatures from 160 to 250°C;

DCuDAlT=D0CuD0AleQAlQCuRT=14e9kJmol1RTinγ2-phasephase3Cu9Al4,CAl=4/130.31,
D3CuD3AlT=250oC=14e9000Jmol1RT1.77,D3CuD3AlT=160oC=14e9kJmol1RT1.15,

so the Cu atoms diffuse faster than the Al atoms in γ2-phase at temperatures from 160 to 250°C.

The Cu-rich phases can be formed faster than the Al-rich phases at temperatures from 160 to 250°C, and the Cu atoms can diffuse faster than the Al atoms in the Al-Cu system at temperatures from 160 to 250°C. The Al-rich phases can be formed faster than the Cu-rich phases at temperatures from 400 to 535°C, and the Al atoms can diffuse faster than the Cu atoms in the Al-Cu system at temperatures from 400 to 535°C. It depends on the crystal structure of each phase, but, in general, it could depends on conclusions that the Cu2+ ions are less mobile than the Cu+ ions, and the ratio D*Al/D*Cu depends on temperature.

3.2 Diffusion activation energy calculation

3.2.1 Diffusion activation energy calculation in the Cu-Al system

Mutual diffusion coefficients were calculated for all five phases [19]:

D˜1=5.6105e127.6kJmol1/RTm2/s;D˜2=2.2104e148.5kJmol1/RTm2/s;E34
D˜3b=1.6102e230.5kJmol1/RTm2/s,D˜3a=2.1104e138.1kJmol1/RTm2/s,D˜3=8.5105e136kJmol1/RTm2/s.

We can see that Q1 < Q2, Q1 < Q3, and Q3 < Q2 because of K1 > K2, K1 > K3, and K3 > K2, and D01 ≈ D02 ≈ D03. Phase j’s rate formation is Kj. Three phases are formed in the Al-Cu system at temperatures 300 and 350°C [5]: CuAl2, CuAl, and Cu9Al4. Phases formation rates were experimentally measured: K1 = 860x10−18 m2/s, K2 = 100x10−18 m2/s, and K3 = 360x10−18 m2/s at temperature 350°C; K1 = 77x10−18 m2/s, K2 = 18x10−18 m2/s, and K3 = 35x10−18 m2/s at temperature 300°C, so K1 > K2, K1 > K3, and K3 > K2. We can calculate assuming C1 = 2/3, C2 = 1/2, C3 = 1/3, C = CAl [11]:

D112C11C1K1+C21C1K1K2+C31C1K1K3;E35
D212C21C2K2+C21C1K1K2+C31C2K2K3;E36
D312C31C3K3+C31C1K1K3+C31C2K2K3;E37
D1T2=3500C19K1+112K1K2+118K1K31501018m2/s,D1T1=300oC151018m2/s;E38
D2T2=350oC18K2+112K1K2+112K2K350x1018m2/s,D2T1=300oC81018m2/s;E39
D3T2=350oС19K3+118K1K3+112K2K3901018m2/s,D3T1=300oC121018m2/s.E40

Moisy et al. [5] did not calculate diffusion activation energies and the pre-exponential factors, so we can do it:

Qi=RT1T2T2T1lnDiT2DiT1,D0i=DiT1eQiRT1=DT2ieQiRT2;E41
D˜1=4.3105e136.7kJmol1/RTm2/s,D˜2=6.6108e108.8kJmol1/RTm2/s,E42
D˜3=9.6107e119.6kJmol1/RTm2/s.

Eq. (42) correspond to Eq. (34). We can use several, N, points to calculate by the least square method to increase calculation precise:

Qi=Nj=1N1000RTjlnDiTjj=1NlnDiTjj=1N1000RTjNj=1N1000RTj2j=1N1000RTj2kJ/mol,E43
D0i=expj=151000RTj2j=15lnDiTjj=151000RTjj=141000RTjlnDiTj5j=151000RTj2j=151000RTj2m2/s.E44

Eqs. (43) and (44) give Eq. (41) for only two points (N = 2).

3.2.2 Diffusion activation energy calculation in pure iron

A method of dislocation pipe diffusion parameter determination during the type B diffusion kinetics was suggested by the model of dislocation pipe diffusion involving outflow [6, 20]. The method involves diffusion dislocation pipe kinetics for two different annealing times at the same temperature during the type B kinetics and dislocation pipe kinetics for one annealing time at other lower temperature during the type C kinetics. Transition time for type B kinetics to type A kinetics (volume diffusion) and kinetics law t1/6 [7] for cone top rate are used in this method.

Bulk diffusion coefficients, DV, for the diffusion of 59Fe in the high-purity iron were calculated in [21] using type B → A kinetics: DV = 1.5*10−18 m2s−1 at T1 = 973 K for tB → A = 67.5ks (Tm/T1 = 1.86, Tm is melting point of iron). Only one experiment was carried out at the same temperature for two annealing times t1 and t2 (t1 < < t2, t2 = 40 t1). Dislocation diffusion coefficients for the diffusion of 59Fe in the iron were calculated in [21] using type C kinetics: Dd = 3*10−16 m2s−1 at T2 = 753 K for tC = 2.4ks (Tm/T2 = 2.4). One can find ratio Dd/DV: ytCB=Dd6Dδ, where δ = 1 nm, DdD=4.3x106. Ratio Dd/DV increases remarkably for lower temperature. Dislocation pipe and volume diffusion activation energies and pre-exponential factors were not calculated in [21]. It is possible to calculate Ed and D0: Ed=lnDdT1DdT2kBT1T2T1T2, D0=DdT1expEdkBT1,Ed=1.1eV;Qd=106kJ/mol, D0=6.85109m2s1. One can calculate dislocation pipe diffusion coefficient for temperature 973 K directly (T1 = 753 K and T2 = 693 K (type C kinetics)): Dd ≈ 10−14 m2s−1. Such value corresponds to value calculated using the proposed method. The Fisher law (t1/4) gives Dd ≈ 10−16 ÷10−15 m2s−1. Such value is in two orders lower than experimentally obtained in [21]. The volume diffusion activation energy EV can be calculated: EV=lnD0DVT1kBT1, EV=1.85eV;QV179kJ/mol. Ratio EdEV=0.6 as described in [8].

Advertisement

4. Conclusions

The Al atoms diffuse faster than the Cu atoms at a temperature higher than 475°C, but the Cu atoms diffuse faster than the Al atoms at a temperature lower than 100°C. The diffusion activation energy of Al is less than the diffusion activation energy of Cu at a temperature higher than 475°C, but diffusion activation energy of Cu is less than the diffusion activation energy of Al at a temperature lower than 100°C. Our investigations show that it is possible because the Cu2+ ions are less mobile than Cu+ ions.

Volume diffusion activation energy of Fe is higher than volume diffusion activation energy of Cu or Al, but dislocation pipe diffusion activation energy of Fe is smaller than volume diffusion activation energy of Cu or Al, so the Fe atoms diffuse faster along the dislocation line, but the Cu or Al atoms diffuse faster in volume.

References

  1. 1. Kizaki T, Minho O, Kajihara M. Rate-controlling process of compound growth in Cu-Clad Al wire during isothermal annealing at 483–543 K. Materials Transactions. 2020;61(1):188-194. DOI: 10.2320/matertrans.MT-M2019207
  2. 2. Braunovic M, Alexandrov N. Intermetallic compounds at aluminum-to-copper electrical interfaces: Effect of temperature and electric current. IEEE Transactions on Components, Packaging, and Manufacturing Technology: Part A. 1994;17(1):78-85. DOI: 10.1109/95.296372
  3. 3. Goh CS, Chong WLE, Lee TK, Breach C. Corrosion study and intermetallics formation in gold and copper wire bonding in microelectronics packaging. Crystals. 2013;3(3):391-404. DOI: 10.3390/cryst3030391
  4. 4. Yarmolenko MV. Copper and aluminum electric corrosion investigation and intermetallics disappearance in Cu-Al system analysis. Phys. Chem. Solid St. 2020;21(2):294-299. https://journals.pnu.edu.ua/index.php/pcss/article/view/3055
  5. 5. Moisy F, Sauvage X, Hug E. Investigation of the early stage of reactive interdiffusion in the Cu-Al system by in-situ transmission electron microscopy. Materialia. 2020;9:100633. DOI: 10.1016/j.mtla.2020.100633
  6. 6. Yarmolenko MV. Method of dislocation and bulk diffusion parameters determination. Metallofizika i Noveishie Tekhnologii. 2020;42(11):1537-1546. https://mfint.imp.kiev.ua/article/v42/i11/MFiNT.42.1537.pdf
  7. 7. Yarmolenko MV. Intermediate phase cone growth kinetics along dislocation pipes inside polycrystal grains. AIP Advances. 2018;8:095202. DOI: 10.1063/1.5041728
  8. 8. Mehrer H. Diffusion in Solids. New York: Springer; 2007. 651p. http://users.encs.concordia.ca/∼tmg/images/7/79/Diffusion_in_solids_Helmut_Mehrer.pdf
  9. 9. Hentzell HTG, Tu KN. Interdiffusion in copper–aluminum thin film bilayers. II. Analysis of marker motion during sequential compound formation. Journal of Applied Physics. 1983;54:6929-6937. DOI: 10.1063/1.332000
  10. 10. Darken LS. Diffusion, mobility and their interrelation through free energy in binary metallic systems. Transactions AIME. 1948;175:184-201. http://garfield.library.upenn.edu/classics1979/A1979HJ27500001.pdf
  11. 11. Yarmolenko MV. Intermetallics disappearance rate analysis in double multiphase systems. DDF. 2021;407:68-86. DOI: 10.4028/www.scientific.net/ddf.407.68
  12. 12. Włodarczyk PP, Włodarczyk B. Effect of hydrogen and absence of passive layer on corrosive properties of aluminium alloys. Materials. 2020;13(7):1580-1593. DOI: 10.3390/ma13071580
  13. 13. Kumar S, Handwerker CA, Dayananda MA. Intrinsic and interdiffusion in Cu-Sn system. JPEDAV. 2011;32:309-319. DOI: 10.1007/s11669-011-9907-91547-7037
  14. 14. Tu KN. Electronic Thin-Film Reliability. 1st ed. New York: Cambridge University Press; 2010. 392p. DOI:https://www.amazon.com/Electronic-Thin-Film-Reliability-King-Ning-Tu-ebook-dp-B00QIT3LXA/dp/B00QIT3LXA/ref=mt_other?_encoding=UTF8&me=&qid=
  15. 15. Epishin A, Chyrkin A, Camin B, Saillard R, Gouy S, Viguier B. Interdiffusion in CMSX-4 related Ni-base alloy system at a supersolvus temperature. DDF. 2021;407:1-10. DOI: 10.4028/www.scientific.net/ddf.407.1
  16. 16. Prawoto Y. Synergy of erosion and galvanic effects of dissimilar steel welding: Field failure analysis case study and laboratory test results. Journal of King Saud University – Engineering Sciences. 2013;25:59-64. DOI: 10.1016/j.jksues.2011.12.001
  17. 17. Yarmolenko MV. Intrinsic diffusivities ratio analysis in the Al-Cu system. Phys. Chem. Solid St. 2020;21(4):720-726. https://journals.pnu.edu.ua/index.php/pcss/article/view/4440
  18. 18. Yarmolenko MV. Intermetallics disappearance rates and intrinsic diffusivities ratios analysis in the Cu-Zn and the Cu-Sn systems. Phys. Chem. Solid St. 2021;22(1):80-87. https://journals.pnu.edu.ua/index.php/pcss/article/view/4744
  19. 19. Funamizu Y, Watanabe K. Interdiffusion in the Al–Cu system. Transactions of the Japan Institute of Metals. 1971;12(3):147-152. DOI: 10.2320/matertrans1960.12.147
  20. 20. Yarmolenko MV. Analytically solvable differential diffusion equations describing the intermediate phase growth. Metallofizika i Noveishie Tekhnologii. 2018;40(9):1201-1207. https://mfint.imp.kiev.ua/article/v40/i09/MFiNT.40.1201.pdf
  21. 21. Shima Y, Ishikawa Y, Nitta H, Yamazaki Y, Mimura K, Isshiki M, et al. Self-diffusion along dislocations in ultra high purity iron. Materials Transactions. 2002;43(2):173-177. https://www.jim.or.jp/journal/e/43/02/173.html

Written By

Mykhaylo Viktorovych Yarmolenko

Submitted: 28 March 2021 Reviewed: 03 September 2021 Published: 10 December 2021