Sie sind auf Seite 1von 18

Scripta Materialia

Ultrahigh Thermal Conductivity of Interface Materials by Good Wettability of Indium and


Diamond
--Manuscript Draft--

Manuscript Number: SMM-20-1571

Article Type: Regular article

Section/Category: Opt in to First Look

Keywords: Low melting point metal; Diamond; Wetting; thermal conductivity

Abstract: The thermal conductivity of Bi-In-Sn/diamond composites was improved by pre-adding


indium particles fabricated using slice technique. Using in situ imaging and particle
dipping experiment, the wetting behavior of diamond microparticle with pure indium,
indium-based and gallium-based liquid metal (LM) was investigated. The diamond
microparticle was well wetted by molten indium and the wettability of diamond particle
with LM can be improved by alloying with indium. Oxide film would hinder the wetting of
LM and diamond. The highest thermal conductivity of the indium-based composites
obtained was up to ~211.12 W m -1 K -1 , which was close to that of aluminum.

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Electronic copy available at: https://ssrn.com/abstract=3674207
Graphical Abstract

Electronic copy available at: https://ssrn.com/abstract=3674207


Manuscript (word or pdf w/ figures) Click here to view linked References

Ultrahigh Thermal Conductivity of Interface Materials by Good


1
2 Wettability of Indium and Diamond
3
4
5 Chengzong Zeng, Jun Shen*, Jianbo Zhang
6
7
8 State Key Laboratory of Mechanical Transmission, College of Material Science and Engineering,
9
10
11 Chongqing University, Chongqing, 400044, China.
12
13
14
The thermal conductivity of Bi-In-Sn/diamond composites was improved by pre-adding indium particles
15
16 fabricated using slice technique. Using in situ imaging and particle dipping experiment, the wetting behavior
17
18
19 of diamond microparticle with pure indium, indium-based and gallium-based liquid metal (LM) was
20
21
22 investigated. The diamond microparticle was well wetted by molten indium and the wettability of diamond
23
24
25 particle with LM can be improved by alloying with indium. Oxide film would hinder the wetting of LM and
26
27
28 diamond. The highest thermal conductivity of the indium-based composites obtained was up to ~211.12 W m-
29
30 1
31 K-1, which was close to that of aluminum.
32
33
34 Keywords: Low melting point metal; Diamond; Wetting; Thermal conductivity
35
36
37
In the microelectronic industry, more and more attention has been paid to the development of high-
38
39 performance thermal solutions due to the increasing integration and continuous miniaturization [1, 2].
40
41
42 According to Arrhenius’s experience rule [3], the failure rate of electronic components will be doubled for
43
44
45 every 10°C increase. Generally, the failure rate of electronic devices caused by excessive temperature was
46
47
48 more than 80% [4]. In practical applications, interface materials are used to fill the microscopic air gaps
49
50
51 between adjacent components to promote the heat transfer of microelectronic devices. Compared with the low
52
53
54 thermal conductivity (1-5 W m-1 K-1 [5]) and high contact thermal resistance (14-100 mm2 K W-1 [6, 7]) of
55
56
57
58
59 *
60 Corresponding author: Tel: +86-13883111150; E-mail: shenjun@cqu.edu.cn (J. Shen)
61
1
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
polymer-based thermal interface materials (TIMs), low melting point metal (LMPM) present higher thermal
1
2 conductivity (10-30 W m-1 K-1 [8]) and lower interfacial thermal resistance (0.25-5 mm2 K W-1 [9]). However,
3
4
5 it is imminently urgent to find LMPMs with higher thermal conductivity to meet the increasing demand of
6
7
8 thermal design [8].
9
10
11 Most researches have indicated that the thermal conductivity of the matrix material can be improved by
12
13
14
adding high thermal conductivity particles [10-12]. Recent studies have reported successful preparation of
15
16 LMPM-based composites containing microparticles of Ag [13, 14], Cu [15], Mg [16], W [17], carbon
17
18
19 nanotubes [18], diamond [19, 20] and so on. According to Wei’s report [19], the reaction rate of LMPM with
20
21
22 most metal materials is very rapid and transform into intermetallic compounds in a very short time. Thus, the
23
24
25 inorganic nonmetal fillers with high thermal conductivity can be a good choice. Dispersions of original
26
27
28 diamond particles in EBiInSn [20] (eutectic field’s alloy Bi32.5In51Sn16.5) and diamond particles coated in a
29
30
31 chromium wetting layer in pure gallium [19] has reportedly achieved three to fourfold increase in thermal
32
33
34 conductivities from that of the base metal (up to 71 W m-1 K-1 and 112 W m-1 K-1, respectively). However,
35
36
37
the thermal conductivity of LMPM-based composites is still much lower than that of the hot (~150 W m-1 K-
38
39 1
for silicon) and the cold terminal (~237 W m-1 K-1 for aluminum, ~401 W m-1 K-1 for copper). Simultaneously,
40
41
42 the wettability of LMPM with diamond particle lacks further research.
43
44
45 In this study, the pure indium particles fabricated by slice technique as a pre-additive compounded with
46
47
48 diamond and EBiInSn. The wetting process of diamond microparticle with molten indium, EBiInSn, and
49
50
51 gallium-based metal was conducted using in situ imaging and particle dipping experimentation. The highest
52
53
54 thermal conductivity of indium-based composites obtained in this work was up to ~211 W m-1 K-1, which was
55
56
57 higher than that of silicon and close to that of aluminum.
58
59
60 Eutectic field’s metal Bi32.5In51Sn16.5 (EBiInSn) particles and indium particles were prepared using slice
61
2
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
technique [21] (the purity of the metal is at least 99.995%, Shenyang Jiabei Trading Co., Ltd., China). 30 g
1
2 (approx.) metal was added in citric acid solution (0.2 wt% for EBiInSn, 0.1 wt% for indium particles) in
3
4
5 glycerol. The solution prepared in a glass beaker (100 mL) was kept in an oil bath as determined temperature
6
7
8 (120°C for EBiInSn and 220°C for indium) for at least 10 min before slice. The slice rate was 15000 rpm
9
10
11 (using a homogenizer, FSH-2A) and the lasting time was 5 min. Excess citric acid and glycerol were washed
12
13
14
out with ethanol. Particles were dried in a vacuum oven (BZF-30) and kept in an airtight container. Diamond
15
16 particles with MBD8 quality were used in this study (Zhecheng Hongxiang Superhard Material Co., Ltd.,
17
18
19 China). The diamond particles, EBiInSn powder, and indium powder were mechanically mixed and pressed
20
21
22 into tablets with the diameter of 12.5 mm. Then the specimens were sintered (at 180°C for 30 min) in an oven
23
24
25 (LTKC-4-10A) with vacuum condition.
26
27
28 The microstructures and crystallographic structures of the samples were characterized by field emission
29
30
31 scanning electron microscope (FESEM, Zeiss Auriga-39-60) equipped with energy dispersive X-ray
32
33
34 spectrometer (EDS) and X-ray diffraction (XRD, Empryean) with Cu Kα radiation. The melting process and
35
36
37
the specific heat capacity Cp of the samples were measured using a differential scanning calorimeter (DSC,
38
39 TGA/DSC1/1100LF). The density of the specimens  c was measured using the Archimedes principle. The
40
41
42 thermal diffusivity α was measured by a LFA467 MicroFlash (Netzsch) thermal analyzer at room temperature.
43
44
45 Each sample was tested 3~5 times to reduce testing errors of experimental data. The thermal conductivity Kc
46
47
48 was calculated from the equation K c =c C p .
49
50
51 The micrograph of EBiInSn particles was presented in Fig. 1a, highly spherical EBiInSn microparticles
52
53
54 were obtained. Surface of the EBiInSn microparticles was smooth, non-pits and defects were observed.
55
56
57 Diameters of the EBiInSn particles were measured and calculated (Fig. 1b). Whose diameters were
58
59
60 concentrated in 20~30 μm. Depending on the XRD results (Fig. 1c). The EBiInSn particles were mainly
61
3
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
composed of Bi, In, Sn and BiIn2 phases, which were consistent with that of bulk EBiInSn [22]. Indium
1
2 microparticles were successfully fabricated using slice technique. Quasi-spherical indium powders were
3
4
5 obtained (Fig. 1d). The surface of indium particle was not as smooth as that of EBiInSn particles, which is
6
7
8 chiefly because the lower surface tension of molten indium (~0.556 N m-1 [23]) than that of EBiInSn (~0.58N
9
10
11 m-1 [24]) and the lower Moh’s hardness (1 for indium, 1.8 for tin, and 2.25 for bismuth) of indium [25]. Thus,
12
13
14
indium particles were too soft to maintain higher and better spherical shape during the slice process of
15
16 collisions between particles caused by high-speed shear force. Equivalent diameters of the indium
17
18
19 microparticles were mostly concentrated in 15~20 μm (Fig. 1e). The XRD results indicate that the indium
20
21
22 particles obtained were only the indium without other phases through slice technique (Fig. 1f). Meantime, the
23
24
25 melting point of EBiInSn and indium did not change after slice process (Fig. 1g). Fig. 1h was the schematic
26
27
28 representation of mixing diamond particles, indium particles, and EBiInSn particles in different proportions.
29
30
31 Macrograph of the final manufactured specimens was displayed in Fig. 1i.
32
33
34 To study the wettability of diamond microparticles in molten indium and molten EBiInSn, simple in situ
35
36
37 experiments were conducted. Fig. 2a presented the schematic of in situ wetting experiments. The temperature
38
39 of the low melting point metal (LMPM) in wetting process was ~30°C higher than that of its melting point,
40
41
42 which was measured by the computer through a data acquisition (UNI-T UT325). The diamond microparticles
43
44
45 of 35/40 mesh were bonded on quartz capillary tubes using silicone glue with heat resistance of 250°C. The
46
47
48 wetting process was recorded using a CCD microscope camera (DM3). Fig. 2b displays the micrograph of
49
50
51 bulk indium with diamond microparticles before and after heating. The diamond particles were pre-attached
52
53
54 to bulk indium. After the bulk indium melted, parts of the diamond particles were “eaten” by molten indium.
55
56
57 This phenomenon could be owing to the diamond wetted by molten indium. Fig. 2c shows the single diamond
58
59
60 microparticle wetting experiment in molten indium. To avoid wetting results of the particle surface affected
61
4
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
by the oxide layer of indium, the surface of molten indium was cleaned using a polytetrafluoroethylene scraper
1
2 (viewed by a CCD microscope) and the time of single wetting experiment was controlled about 15s. The
3
4
5 diamond microparticle was immersed into the molten indium and then pulled out until the diamond
6
7
8 microparticle and molten indium were completely separated. The diamond particle and molten indium showed
9
10
11 excellent adhesion during the pull-out stage. The wetting process of the single diamond particle with molten
12
13
14
indium showed that the diamond particle was completely wetted by molten indium and the quartz capillary
15
16 tube was also well wetted.
17
18
19 Using the same wetting method, the diamond microparticle was immersed into the molten EBiInSn with
20
21
22 or without oxide present. Fig. 2d was wetting behavior between diamond microparticle and molten EBiInSn
23
24
25 with an oxide layer (prepared by melting EBiInSn placed in air for one hour). The results showed that only a
26
27
28 small part of the diamond microparticle was wetted by molten EBiInSn. However, the majority of the diamond
29
30
31 microparticle were well wetted by clean molten EBiInSn (Fig. 2e). It noted that the oxide film on the molten
32
33
34 EBiInSn would hinder the diamond particle from being wetted. The wetting behavior of gallium and diamond
35
36
37
was also conducted (Figure S1a, Supporting Information), which showed that the wettability of diamond and
38
39 gallium was poor. However, the wetting behavior of gallium and diamond particle can be improved by alloying
40
41
42 with indium (Figure S1b, Supporting Information). This phenomenon is significant for the preparation of
43
44
45 gallium-based low melting point metal composites containing diamond particles.
46
47
48 Fig. 3a illustrates the morphology of EBiInSn-diamond-indium composite (Bi-In-Sn/diamond composite)
49
50
51 before sintering. The EBiInSn particles still kept spherical morphology, while the initial indium particles were
52
53
54 compressed into a flat shape due to their excellent softness (Fig. 3b and c). After sintering, the EBiInSn
55
56
57 particles and indium particles were melted together. The diamond particles were well surrounded by the
58
59
60 indium-based metal (Fig. 3d) owing to good wettability between them. Fig. 3e was the enlarged view of the
61
5
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
matrix metal formed by the re-melting of EBiInSn and indium particles. Combined with the microstructure
1
2 and its element distribution analysis (Fig. 3f), the Sn element from original EBiInSn had obviously diffused
3
4
5 to the initial indium particles. Fig. 3g presented the fracture surface of the Bi-In-Sn/diamond composite, most
6
7
8 surfaces of diamond particles were well coated with indium-based metal and the diamond microparticles were
9
10
11 effectively connected by the indium-based metal. From the enlarged view (Fig. 3h), the diamond particle was
12
13
14
obviously and completely wetted by indium-based metal and no obvious gap existed between the diamond
15
16 and the matrix metal. It indicated that the diamond particles can be well wetted by pre-adding indium and the
17
18
19 effective heat conduction path for diamond particles can be formed.
20
21
22 Priorly, the thermal conductivity of Bi-In-Sn/diamond composites influenced by the pre-adding indium
23
24
25 particles volume fraction was analyzed (Fig. 4a). The volume fraction of diamond particles was set at 30%
26
27
28 and the rest was EBiInSn except for pre-adding indium particles. The thermal conductivity of the composites
29
30
31 increased from ~43.19 W m-1 K-1 to 55.15 W m-1 K-1 as the volume fraction of pre-adding indium particles
32
33
34 increased from 0% to 30%. Next, the thermal conductivity of the composites affected by the volume fraction
35
36
37
of diamond particles was measured. On the one hand, the volume fraction of EBiInSn was at the rate of 50%
38
39 and the rest was pre-adding indium particles. The thermal conductivity of the composites increased with the
40
41
42 diamond particles volume fraction as the diamond content was no more than 30% (Fig. 4b). When the diamond
43
44
45 volume fraction was 40%, the thermal conductivity of the composites was no significant increase. This may
46
47
48 be attributable to the fact that the ratio of diamond to pre-adding indium was in relative equilibrium. On the
49
50
51 other hand, the pre-adding indium volume fraction was fixed at 30% and the rest was EBiInSn. The thermal
52
53
54 conductivity of the composites increased from ~31.84 W m-1 K-1 to ~105.84 W m-1 K-1 as the volume fraction
55
56
57 of the diamond increased from 10% to 50% (Fig. 4c). Further, the thermal conductivity of indium-based metal
58
59
60 composites containing diamond particles influenced by the dimension of the diamond was analyzed. The
61
6
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
thermal conductivity of the composites increased with the size of diamond particles (Fig. 4d). The highest
1
2 thermal conductivity achieved in this work was indium/diamond composite, which was ~211.12 W m-1 K-1.
3
4
5 Fig. 4e was the DSC curves of the composites varied with the ratio of original EBiInSn and pre-adding indium.
6
7
8 Two obvious melting stages were observed. The initial melting temperature of the composites was close to the
9
10
11 melting point of EBiInSn (~60°C) and the complete melting temperature was delayed with the increase of pre-
12
13
14
adding indium volume fraction. The effect of aging on the DSC curve of the composites was small (Fig. 4f).
15
16 Compared with the melting point of uniform alloying of Bi, In, and Sn (Fig. 4g), the melting point of pre-
17
18
19 adding indium into EBiInSn was much lower. This is mainly because the pre-adding indium and original
20
21
22 EBiInSn were difficult to form a homogeneous alloy at a lower temperature with the obstacles of diamond
23
24
25 particles. Namely this material can be melted at a relatively low temperature.
26
27
28 A simplified 2D model with the size of 500 μm×500 μm was built to predict the thermal conductivity of
29
30
31 the composites. Before simulation, some assumptions were made: 1. The pre-adding indium particles were
32
33
34 circular and the diamond particles were orthohexagnal, 2. The gap between pre-adding indium and diamond
35
36
37
was filled with EBiInSn, 3. The mutual diffusion of EBiInSn and pre-adding indium was ignored. The
38
39 temperature of the bottom wall and the top wall was set at 57oC and 20oC. The left and the right were set as
40
41
42 insulation walls. The thermal conductivity Kc of this model can be calculated using the equation
43
44
K c  qavg D Thot  Tcold  , where qavg is the average surface heat flux, D is the distance between the cold end
1
45
46
47
48 and the hot end, Thot and Tcold were the temperature of the hot end and the cold end, respectively. Fig. 5a
49
50
51 presents the heat flux density with vectors in the composites varied with the volume fraction of diamond
52
53
54 particles. It states that the heat flux mainly passed through the diamond particles and the pre-adding indium
55
56
57 particles nearing the diamond. The heat flux density was increased with the increase of diamond volume
58
59
60 fraction, which was directly proportional to the thermal conductivity calculated by the model. Compared with
61
7
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
the results without pre-adding indium particles (Fig. 5a iii), the heat flux density of the model with pre-adding
1
2 indium was significantly improved (Fig. 5a iV). The changing trend of experimental results and simulation
3
4
5 results was basically consistent (Fig. 5b). However, the difference between experimental results and
6
7
8 simulation results became larger as the diamond volume fraction higher than 40%, which could be due to the
9
10
11 gap occurred between diamond and matrix metal. Fig. 5c shows the comparison of thermal conductivity of
12
13
14
interface materials obtained in this work with that of other relative researches. It indicates that the thermal
15
16 conductivity of the materials in this study was in a better level and the highest thermal conductivity achieved
17
18
19 was ~211.12 W m-1 K-1, which was close to that of aluminum (~224 W m-1 K-1 [27]).
20
21
22 In summary, quasi-spherical indium particles and spherical EBiInSn particles were successfully
23
24
25 manufactured by slice technique. Wetting behavior of diamond microparticle with indium, EBiInSn, and
26
27
28 gallium-based metal was studied using in situ imaging and particle dipping experiment. The results presented
29
30
31 that the diamond microparticle was well wetted by molten indium and the wettability of diamond particle with
32
33
34 LMPM can be improved by alloying with indium. The wettability between diamond microparticle and LMPM
35
36
37
would be affected by oxide film of LMPM. The thermal conductivity of indium-based composites containing
38
39 diamond particles varied with the volume fraction of pre-adding indium, the diamond volume fraction, and
40
41
42 the dimension of diamond particles. It had little effect on the initial melting temperature of EBiInSn by pre-
43
44
45 adding indium particles and the complete melting temperature of the composite was relatively low. Namely
46
47
48 this material can be operated at a relatively low temperature. The highest thermal conductivity of the indium-
49
50
51 based composites obtained in this work was up to ~211.12 W m-1 K-1, which was close to that of aluminum.
52
53
54 Acknowledgement
55
56
57 This research is supported by a Fundamental Research Funds for the Central Universities of China (Grant
58
59
60 No. 2018CDGFCL0003).
61
8
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
References
1 [1] X.C. Tong, Advanced Materials for Thermal Management of Electronic Packaging, 2011.
2
3 [2] M.M. Waldrop, Nature 530(7589) (2016) 144-147.
4 [3] C. Harper, Electronic Packaging and Interconnection Handbook, McGraw-Hill, Inc.2004.
5
6 [4] A.D. Kraus, A. Bar-Cohen, Thermal analysis and control of electronic equipment, 1983.
7 [5] M.A. Raza, A.V.K. Westwood, A.P. Brown, C. Stirling, Journal of Materials Ence Materials in Electronics
8
9
23(10) (2012) 1855-1863.
10 [6] K. Zhang, M.M.F. Yuen, N. Wang, J.Y. Miao, H.B. Fan, Thermal interface material with aligned CNT and
11 its application in HB-LED packaging, Electronic Components & Technology Conference, 2006.
12
13 [7] Y.H. Zhao, Z.K. Wu, S.L. Bai, International Journal of Heat & Mass Transfer 101(oct.) (2016) 470-475.
14 [8] X.H. Yang, J. Liu, Advances in Liquid Metal Science and Technology in Chip Cooling and Thermal
15
16 Management, Elsevier2018.
17 [9] C.K. Roy, S. Bhavnani, M.C. Hamilton, R.W. Johnson, J.L. Nguyen, R.W. Knight, D.K. Harris,
18
19 International Journal of Heat & Mass Transfer 85 (2015) 996-1002.
20 [10] Y.-S. Jhong, H.-T. Tseng, S.-J. Lin, J. Alloy. Compd. 801 (2019) 589-595.
21
[11] Y. Xu, D. Kraemer, B. Song, Z. Jiang, J. Zhou, J. Loomis, J. Wang, M. Li, H. Ghasemi, X. Huang, X. Li,
22
23 G. Chen, Nat. Commun. 10(1) (2019) 1771.
24 [12] S. Bhanushali, P.C. Ghosh, G.P. Simon, W. Cheng, Advanced Materials Interfaces 4(17) (2017) 1700387.
25
26 [13] J. Tang, X. Zhao, J. Li, J. Liu, Advanced Materials Interfaces (2018) 1800406.
27 [14] Z. Lin, H. Liu, Q. Li, L. Han, C. Sheng, Y. Yang, G. Chu, Applied Physics A 124(5) (2018) 368.
28
29 [15] J. Tang, X. Zhao, J. Li, R. Guo, Y. Zhou, J. Liu, ACS Appl. Mater. Interfaces (2017) acsami.7b10256.
30 [16] X. Wang, W. Yao, R. Guo, X. Yang, J. Tang, J. Zhang, W. Gao, V. Timchenko, J. Liu, Advanced
31
32
Healthcare Materials (2018) e1800318.
33 [17] W. Kong, Z. Wang, M. Wang, K.C. Manning, K. Rykaczewski, Adv Mater 31(44) (2019).
34 [18] L.Y. Zhao, S. Chu, X.C. Chen, G. Chu, Bull. Mater. Sci. 42(4) (2019) 5.
35
36 [19] S. Wei, Z.F. Yu, L.J. Zhou, J.D. Guo, J. Mater. Sci.: Mater. Electron. 30(7) (2019) 7194-7202.
37 [20] C.Z. Zeng, J. Shen, C. He, H. Chen, Scr. Mater. 170 (2019) 140-144.
38
39 [21] S. Cinar, I.D. Tevis, J. Chen, M. Thuo, Sci. Rep. 6 (2016).
40 [22] V.T. Witusiewicz, U. Hecht, B. Böttger, S. Rex, J. Alloy. Compd. 428(1) (2007) 115-124.
41
42 [23] H. Tostmann, E. Dimasi, P.S. Pershan, B.M. Ocko, O.G. Shpyrko, M. Deutsch, Physical Review B 59(2)
43 (1999) 783-791.
44
45
[24] E.E.M. Noor, N.M. Sharif, C.K. Yew, T. Ariga, A.B. Ismail, Z. Hussain, J. Alloy. Compd. 507(1) (2010)
46 290-296.
47 [25] M.B. Peterson, S.F. Murray, J.J. Florek, A S L E Transactions 2(2) (1959) 225-234.
48
49 [26] S. Rex, B. Böttger, V. Witusiewicz, U. Hecht, Mater. Sci. Eng. A-Struct. Mater. Prop. Microstruct. Process.
50 413-414 (2005) 249-254.
51
52 [27] Z. Tan, D.B. Xiong, G. Fan, Z. Chen, Q. Guo, C. Guo, G. Ji, Z. Li, D. Zhang, J. Mater. Sci. (2018).
53 [28] X.-H. Yang, S.-C. Tan, Y.-J. Ding, L. Wang, J. Liu, Y.-X. Zhou, Int Commun Heat Mass 87 (2017) 118-
54
55
124.
56
57
58
59
60
61
9
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
Figures
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
Fig. 1. SEM images of (a) EBiInSn particles and (d) indium particles; size distribution of (b) EBiInSn particles
32
33 and (e) indium particles; X-ray diffractogram of (c) EBiInSn particles and (f) indium particles; (g) DSC cures
34 of EBiInSn and indium (both bulk metal and corresponding particles) as measured at a ramping rate of 5°C
35
36 min-1; (h) schematic diagram of diamond, indium and EBiInSn mixtures (not to scale); (i) macrograph of the
37 mixtures.
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
10
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19 Fig. 2. (a) The schematic of in-situ wetting experimental platform (not to scale); (b) micrograph of diamond
20 particles with indium before and after heating; (c) wetting behavior of diamond microparticle with molten
21
22 indium; (d) immersion of diamond microparticle into EBiInSn with oxide attachment (digital microscope
23 view); (e) immersion of diamond microparticle into clean molten EBiInSn.
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
11
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22 Fig. 3. SEM images of (a) EBiInSn-diamond-indium composite before sintering and (b) the enlarged view
23 with (c) EDS image; SEM images of (d) EBiInSn-diamond-indium composite after sintering and (e) the
24
25 enlarged view with (f) EDS mapping; (g) the fracture surface of the composite and (h) an enlarged view.
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
12
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29 Fig. 4. The thermal conductivities of Bi-In-Sn/diamond composites varied with (a) initial indium particles
30 volume fraction (the diamond was fixed at 30 vol%) and diamond volume fraction: (b) the initial EBiInSn
31
was fixed at 50 vol%, (c) the initial indium was fixed at 30%; (d) measured thermal conductivities along with
32
33 particle size of diamond at different initial mixing ratios; DSC curves of the Bi-In-Sn/diamond composites:
34 (e) at different initial ratios of EBiInSn and indium particles, (f) influenced after aging; (g) liquidus projection
35
36 of the ternary system Bi-In-Sn [26].
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
13
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29 Fig. 5. Thermal flux in Bi-In-Sn/diamond composites varied with the volume fraction of diamond particles;
30 (b) comparison between experimental and simulation results; (c) comparison of thermal conductivity with
31
literature [14, 15, 17-20, 28].
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
14
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
Manuscript (word or pdf w/ figures) Click here to view linked References

1 Conflict of interest
2
3 The authors declare that they have no conflict of interest.
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
Electronic copy available at: https://ssrn.com/abstract=3674207
Supplementary Material

Click here to access/download


Supplementary Material
Supporting information.docx

Electronic copy available at: https://ssrn.com/abstract=3674207

Das könnte Ihnen auch gefallen