We have developed and optimized a method to grow ruthenium films of unprecedented quality. Our three-step process is reminiscent of solid-phase epitaxy. First, c-cut sapphire substrates are terminated at their Al-rich √31 × √31R ± 9° reconstruction by in situ annealing. Second, 3D structured epitaxial Ru films are deposited at low temperatures by laser evaporation of Ru rods. Third, the films, thus, obtained are epitaxially transformed by high-temperature annealing. X-ray diffraction studies reveal good crystallinity of the obtained 15–60 nm-thick films: peak widths of the rocking curve are one order of magnitude smaller than those of the best published films. Scanning transmission electron microscopy and electron energy loss studies show that the interface between the sapphire substrates and the flat Ru films is atomically sharp with very limited intermixing. These results demonstrate the usefulness of postanneal processes for producing high-quality epitaxial films of elemental metals on insulating substrates.

The growth of crystalline films of elemental metals on insulating oxide substrates is challenging due to low film–substrate binding energies.1,2 This difficulty unfortunately also applies to the epitaxial growth of Ru films on sapphire. Such Ru films are of keen interest as ultrathin conductors3–6 and as templates for the epitaxial growth of other metals7–11 and graphene.3,12,13 Sapphire is a preferred substrate for many applications because it is available in high quality and large sizes at comparably low cost.14 

Indeed, the c-plane (0001) surface of sapphire also offers favorable properties for the epitaxial growth of Ru films. Both Ru and the oxygen sublattice of sapphire have hexagonal close-packed (hcp) crystal lattices, which lattice match by rotational alignment with a mismatch of only ≈1.3% [a(Ru) = 2.706 Å, dO–O = 2.747 Å].15 Moreover, the chemical inertness of sapphire is a favorable property, as it avoids chemical reactions between Ru and the substrate during film growth or annealing steps.

Several efforts have been undertaken to grow Ru films on sapphire. Owing to the low vapor pressure of Ru, sputtering has been the preferred growth method. In these studies, Ru films of varying crystalline quality were obtained, including pronounced mosaic structures, growth islands, and significant intermixing with the substrate.10,16,17

Given the capability to prepare sapphire surfaces of improved quality18,19 and to access higher substrate temperatures Ts by means of substrate heating with CO2 laser irradiation,20 we systematically explored the growth of Ru films in this extended parameter range by depositing Ru films via thermal laser evaporation. Thermal laser evaporation has the advantages of low-pressure thermal deposition, potentially boosting film purity and growth rates.20 

We show that high-quality single-crystalline Ru films can indeed be obtained on sapphire substrates, however not by growth at high temperatures, but rather by growth at low Ts followed by high-temperature annealing. The Ru films fabricated in this manner feature superior crystallinity and an atomically sharp interface with the substrate. The surfaces of films as thick as 60 nm follow the step-and-terrace structure determined by the vicinal cut of the sapphire. Microstructural analysis by scanning transmission electron microscopy (STEM) reveals the microscopic origin of this growth behavior. Deposition at low Ts minimizes the formation of pronounced growth islands during the early stages of film growth. For films grown at high Ts, these islands evolve dynamically during further film growth and deteriorate the microstructural quality of the resulting Ru film.

Ru films were grown by thermal laser evaporation using a thermal laser epitaxy (TLE) system equipped with CO2 laser heating20,21 and a base pressure of about 6 × 10−11 mbar on the (0001) surfaces of 5 × 5 mm2, single-side polished Al2O3 substrates from CrysTec. The substrates were terminated by annealing in situ for 200 s at 1700 °C in a background pressure <2 × 10−10 mbar to show the Al-rich √31 × √31R ± 9° reconstruction as described in Ref. 19. This low pressure was maintained during deposition and possible subsequent anneals. Substrate temperatures were measured with an optical pyrometer on the backside of the substrate. Owing to its low vapor pressure, the Ru was evaporated by heating it with a continuous-wave beam of a fiber laser.22 The 1.06-μm laser was operated at 170 W output power. It irradiated ≈1 mm2 areas of cylindrical Ru rods (4.0 N) of 3-mm diameter, yielding growth rates of ≈0.25 Å/s. Film thicknesses were varied by adjusting deposition times.

The substrates were heated by thermal radiation of the Ru sources while these were hot and additionally, if desired, by CO2 laser radiation. The thermal radiation of the hot sources alone heated the substrates up to ≈280 °C within ≈7 min after the start of the deposition. Possible postgrowth annealing steps performed at various temperatures lasted at least 180 s. Before and after film growth, the substrate and film surfaces were monitored by in situ reflection high-energy electron diffraction (RHEED) with an electron energy of 15 kV. To avoid potentially modifying the film microstructure by electron irradiation,23,24 the electron beam of the RHEED was turned off when substrates were hot.

Aiming to grow high-quality Ru films, we first followed the obvious approach and attempted to deposit films epitaxially by growing them at high Ts on the prepared substrate surfaces. To identify the best Ts for epitaxial growth, we fabricated a first series of samples in which we varied Ts from 280 to 1300 °C. This procedure consists of terminating the substrates in situ and growing the film at elevated temperatures. We will refer to this procedure as “two-step process.”

To reveal the influence of Ts on the microstructure of the Ru films, we start by investigating the structural properties of Ru grown at the lowest Ts ≈ 280 °C.

Figure 1(a) shows the RHEED pattern and atomic force microscopy (AFM) images of such a Ru film. The prominent spots of the RHEED image are noticeably blurred and large, indicating a three-dimensional surface structure with epitaxial alignment, which is consistent with the AFM images, which show many small surface features.

FIG. 1.

RHEED images taken along the sapphire [ 1 1 ¯ 00 ] direction (left) and AFM images (right) of samples with [(a)–(d)] and without [(e)] ≈15-nm-thick Ru layers deposited on thermally annealed c-plane sapphire surfaces at the substrate temperatures listed. These samples were fabricated without a postanneal step using the two-step growth process. In all cases, the growth process comprised a 10-min evaporation of a 3-mm-diameter Ru rod with 170 W of laser output power.

FIG. 1.

RHEED images taken along the sapphire [ 1 1 ¯ 00 ] direction (left) and AFM images (right) of samples with [(a)–(d)] and without [(e)] ≈15-nm-thick Ru layers deposited on thermally annealed c-plane sapphire surfaces at the substrate temperatures listed. These samples were fabricated without a postanneal step using the two-step growth process. In all cases, the growth process comprised a 10-min evaporation of a 3-mm-diameter Ru rod with 170 W of laser output power.

Close modal

Now proceeding to growth at higher temperatures, Fig. 1(b) shows the corresponding images for growth at Ts = 400 °C. The RHEED images now show rather broad lines, indicating a smooth, yet also disordered surface. The corresponding granular crystal structure with grains that are larger than the ones obtained at Ts = 280 °C is well visible in the AFM image.

For growth at 700 °C [Fig. 1(c)], the RHEED image shows for the first time Kikuchi lines in addition to the diffracted spots and lines known from the growth at lower Ts, suggesting an ordered microstructure. However, the surface of this sample is not closed, as shown in the corresponding AFM image. The Ru grains have assembled into a meandering, percolative arrangement. Gaps exist where the grains approach each other.

The surfaces of Ru films grown at 1000 °C [Fig. 1(d)] display a rather different morphology. Also for these films, the RHEED image shows very narrow lines with strongly pronounced Kikuchi lines, again indicating a smooth, slightly disordered surface of an epitaxially grown film. The corresponding AFM image shows many large islands. Whereas different islands have different heights, the surfaces of the islands are atomically flat. The edges of the islands tend to be aligned to the hexagonal crystal lattice of the sapphire. The shape of the islands is reminiscent of a Wulff shape.25,26 These are the films with the best microstructure we could obtain by growing at elevated temperatures. Nevertheless, we have to note that, although the films are epitaxial, they are not continuous. Figure 2(a) shows an STEM cross-sectional image of such a Ru film. As expected from Fig. 1(d), the cross-sectional view also shows pronounced growth islands. Moreover, it reveals that the grooves between the islands extend all the way down to the substrate. Furthermore, EELS reveals undesired intermixing of cations at the interface between these islands and the underlying sapphire [Fig. 2(b)]. Considering the high solubility of Al in Ru, this intermixing is likely a result of Al-diffusion into the Ru film.27 This free Al must be generated by chemical reactions between Ru and the underlying sapphire substrate, the results of which are directly observed at even higher TS.

FIG. 2.

(a) STEM image showing a cross-sectional view of a Ru film grown by the two-step process at Ts = 1000 °C on c-plane sapphire, revealing a discontinuous film characterized by growth islands extending down to the film–substrate interface. The image was taken along the sapphire [ 1 1 ¯ 00 ] zone axis. (b) EELS maps of one of the islands shown in (a). Intermixing between the film and the substrate is observed.

FIG. 2.

(a) STEM image showing a cross-sectional view of a Ru film grown by the two-step process at Ts = 1000 °C on c-plane sapphire, revealing a discontinuous film characterized by growth islands extending down to the film–substrate interface. The image was taken along the sapphire [ 1 1 ¯ 00 ] zone axis. (b) EELS maps of one of the islands shown in (a). Intermixing between the film and the substrate is observed.

Close modal

Increasing TS further does not improve the film quality. Indeed, as highlighted by Fig. 1(e), which shows the data obtained after deposition at 1300 °C, the processes occurring during film growth change. The RHEED patterns of such samples taken after the growth procedure [Fig. 1(e)] represent the √31 × √31R ± 9 reconstruction of the Al2O3 (0001) substrate, suggesting that the Ru atoms deposited at this temperature do not remain on the hot substrate surface. The AFM images show numerous triangular holes, suggesting that some of the Ru atoms form a volatile oxide by reducing the Al2O3 and, thus, etching the triangular holes into the sapphire. These holes likely decorate dislocation cores.28,29

The systematic variation of Ts during film deposition, therefore, showed that, by increasing Ts to the highest point at which growth could still be observed (Ts = 1000 °C), the growth islands become larger and develop flat surfaces. However, even for Ts = 1000 °C, at which the film in the islands have the best microstructural quality, grooves remain that reach through the entire film thickness between the islands. These films are, therefore, not continuous.

These observations suggest that the formation of extended islands driven by the increased adatom diffusion, ripening and dewetting at high Ts must be suppressed to obtain continuous films. As these effects are weaker at lower Ts, and because good film crystallinity requires high-temperature diffusion, we explored a three-step process for the growth of Ru films that is reminiscent of solid-phase epitaxy: (solid-phase epitaxy) the films are deposited on the prepared substrates at low Ts and are subsequently annealed at high temperatures. In the process we applied, the deposited Ru films are 3D epitaxially aligned prior to the annealing step. This is subtly different from solid-state epitaxy, which is defined such that amorphous films are deposited and subsequently crystallized.30 

We deposited films at Ts = 280 °C and annealed them for at least 180 s at 1000 °C. This is the temperature at which the two-step process described above yielded the best film quality.

Figure 3 shows the RHEED diffraction pattern and the AFM surface image of a 14-nm-thick Ru film layer grown by this three-step process. The surface morphology of this film shown in the AFM image differs from those of films fabricated with the conventional two-step process described above. First, the surface reveals a continuous film. Second, the surface shows a complex surface structure on a length scale <50 nm and a height <1 nm, which resembles small flames or the sun’s surface imaged in the ultraviolet range.31 The variations of the sample height at these structures are significantly smaller than the lattice parameter. Thus, it is unlikely that they result from dislocations that are close to the surface. However, it seems possible that this surface structure is shaped by strain fields of dislocations located well below the film surface, for example, at the substrate–film interface, as suggested by the protruding lines of constant width, which consist of straight or arced segments of constant curvature. Third, a series of lines crosses the surface from the lower left to the upper right. These reflect the terrace steps of the sapphire substrate, indicating a uniform thickness of the Ru layer. The remarkably low RMS surface roughness of this film is 0.18 nm. As for films deposited at 1000 °C in the two-step process [Fig. 1(d)], the RHEED image exhibits a two-dimensional surface structure and a highly ordered microstructure. The elongation of the spots indicates a variance in diffraction conditions, which can be attributed to the complex surface structure observed in the AFM image and/or to atomic-level surface disorder.

FIG. 3.

RHEED image taken along the sapphire [ 1 1 ¯ 00 ] direction and AFM image of a ≈15-nm-thick Ru film deposited on thermally annealed c-plane sapphire surfaces at Ts = 280 °C. This sample was fabricated using the three-step growth process with a 1000 °C postanneal.

FIG. 3.

RHEED image taken along the sapphire [ 1 1 ¯ 00 ] direction and AFM image of a ≈15-nm-thick Ru film deposited on thermally annealed c-plane sapphire surfaces at Ts = 280 °C. This sample was fabricated using the three-step growth process with a 1000 °C postanneal.

Close modal

The cross section of this sample is shown in the STEM image, see Figs. 4(a) and 4(b). The Ru film is continuous and features a sharp, apparently defect-free interface to the substrate. Film–substrate interdiffusion is very limited, as shown by EELS investigations [Fig. 4(c)]. According to the STEM images, the films have a highly ordered microstructure. Taking a closer look, we see that the STEM cross section [Fig. 4(b)] also reveals a subtle, wavelike shift of the atomic layers with a period of ≈2 nm. These shifts may contribute to the surface morphology as shown in Fig. 3.

FIG. 4.

(a) and (b) STEM images showing cross-sectional views of a Ru film grown by the three-step process at Ts = 280 °C on c-plane sapphire with a 1000 °C postannealing step. The film is continuous and of high crystalline quality. The film–substrate interface is sharp and well-defined. (c) EELS maps of the film shown in (a) and (b). No intermixing between the substrate and the film is observed within the resolution limits.

FIG. 4.

(a) and (b) STEM images showing cross-sectional views of a Ru film grown by the three-step process at Ts = 280 °C on c-plane sapphire with a 1000 °C postannealing step. The film is continuous and of high crystalline quality. The film–substrate interface is sharp and well-defined. (c) EELS maps of the film shown in (a) and (b). No intermixing between the substrate and the film is observed within the resolution limits.

Close modal

Further insight into the effects of the postgrowth anneal on the microstructure of the Ru films is provided by x-ray diffraction studies. Figure 5 shows θ/2θ diffraction measurements of two characteristic samples. These are a first sample fabricated with the two-step process by deposition at 1000 °C without being subjected to a subsequent anneal, and a second sample fabricated with the three-step process by deposition at 280 °C with a subsequent anneal at 1000 °C for at least 180 s. The two scans show exclusively the sapphire 0006 and 00012 and the Ru 0002 and 0004 peaks. Compared with the diffraction pattern of the two-step sample, that of the three-step sample shows a stronger and narrower Ru 0002 peak as well as more pronounced Laue oscillations [Fig. 5(b)], indicating a higher crystallinity of the film obtained by the three-step process and an excellent film–substrate interface, consistent with the STEM cross-sectional results. The high degree of crystallinity of films produced by the three-step process is evidenced directly by the rocking curves of the Ru 0002 peaks [Fig. 5(c)]. The full-width at half-maximum (FWHM) of the curve for the sample shown is only 0.063°, which is about one order of magnitude smaller than the best values published in the literature.6,15

FIG. 5.

(a) and (b) θ/2θ x-ray diffraction patterns of two Ru films grown by the two-step and three-step processes, respectively. The film grown in two steps was deposited at Ts = 1000 °C without postanneal, whereas the film grown in three steps was deposited at Ts = 280 °C and postannealed at 1000 °C. The diffraction pattern indicates a higher quality of the latter film. (c) Rocking curves recorded at the Ru 0002 peak of the samples shown in (a) and (b).

FIG. 5.

(a) and (b) θ/2θ x-ray diffraction patterns of two Ru films grown by the two-step and three-step processes, respectively. The film grown in two steps was deposited at Ts = 1000 °C without postanneal, whereas the film grown in three steps was deposited at Ts = 280 °C and postannealed at 1000 °C. The diffraction pattern indicates a higher quality of the latter film. (c) Rocking curves recorded at the Ru 0002 peak of the samples shown in (a) and (b).

Close modal

As the film–substrate interdiffusion is greater for samples grown with the two-step than three-step process, we speculate that the interdiffusion in the two-step samples is fostered by stress fields associated with the pronounced Ru growth islands in these samples. During growth at high temperatures, the substrate surface is covered only under the growth island by Ru. At the islands’ edges, stress field gradients may enhance Al–Ru interdiffusion such that the Ru islands may appear to be slightly sunk into the sapphire [Fig. 2(a)]; a process referred to as endotaxy.32 During growth at low temperatures, the island density is much higher, disallowing relatively long-range stress fields to develop. Obviously, interdiffusion is also suppressed. Heating a surface fully covered by Ru yields no lateral variation in diffusion rates, and, thus, substrate–film interdiffusion is suppressed. Another explanation for the lack of interdiffusion is that the closed Ru layer kinetically limits the formation of the volatile ruthenium oxide.

As a final point, we note that, to yield high-quality Ru films, the optimal three-step growth process requires the deposited Ru film to be thicker than ≈10 nm. Thinner films do not appear to be sufficiently continuous to stabilize the Ru layer during the postdeposition anneal or dewet as a result of thermodynamic instability.33 The anneal then yields Ru layers that again are marred by islands and trenches. This dewetting threshold temperature of 1000 °C for ≈10-nm-thick epitaxial Ru films on Al2O3 (0001) is much higher than reported for polycrystalline Ru on SiO2.34 

In summary, we have provided a systematic study of the growth behavior of Ru films on c-plane sapphire substrates, a case study for the epitaxial growth of films of elemental metals on oxide insulators. The films were deposited by laser evaporation on thermally prepared (0001) sapphire surfaces with an Al-termination. Without further processing, films deposited at any substrate temperature were 3D, although epitaxial. Growth at 1000 °C yielded films of the highest crystallinity, yet these films were still discontinuous. A three-step process was found to ameliorate the underlying detrimental effects of island formation on film quality. Following in situ substrate termination, Ru films were deposited at low temperatures and then annealed in a high-temperature step. With this process, epitaxial Ru films of unprecedented crystalline quality have been obtained on sapphire (0001).

The authors acknowledge Ute Salzberger for assistance with TEM sample preparation.

The authors have no conflicts to disclose.

Lena N. Majer: Conceptualization (equal); Data curation (lead); Formal analysis (equal); Investigation (lead); Methodology (equal); Validation (equal); Visualization (lead); Writing – original draft (equal). Sander Smink: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Supervision (equal); Validation (equal); Writing – review & editing (equal). Wolfgang Braun: Conceptualization (equal); Data curation (supporting); Formal analysis (equal); Investigation (equal); Methodology (equal); Supervision (equal); Validation (equal); Writing – review & editing (equal). Hangguang Wang: Data curation (equal); Formal analysis (equal); Investigation (equal); Validation (equal); Visualization (equal); Writing – review & editing (equal). Peter A. van Aken: Supervision (equal); Validation (equal); Writing – review & editing (equal). Jochen Mannhart: Conceptualization (equal); Investigation (equal); Supervision (equal); Validation (equal); Writing – original draft (equal); Writing – review & editing (equal). Felix V. E. Hensling: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Supervision (equal); Validation (equal); Visualization (supporting); Writing – original draft (equal); Writing – review & editing (lead).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
A.
Picone
,
M.
Riva
,
A.
Brambilla
,
A.
Calloni
,
G.
Bussetti
,
M.
Finazzi
,
F.
Ciccacci
, and
L.
Duò
,
Surf. Sci. Rep.
71
,
32
(
2016
).
3.
A.
Al Taleb
,
L. N.
Majer
,
E. W.
Selfors
,
S.
Smink
,
B.
Holst
,
J.
Mannhart
,
W.
Braun
, and
D.
Farías
,
Thin Solid Films
758
,
139449
(
2022
).
4.
J.
Kaczmarski
,
A.
Taube
,
M. A.
Borysiewicz
,
M.
Mysliwiec
,
K.
Piskorski
,
K.
Stiller
, and
E.
Kaminska
,
IEEE Trans. Electron Devices
65
,
129
(
2018
).
5.
A. S.
Samardak
et al,
Phys. Chem. Chem. Phys.
24
,
8225
(
2022
).
6.
E.
Milosevic
,
S.
Kerdsongpanya
,
A.
Zangiabadi
,
K.
Barmak
,
K. R.
Coffey
, and
D.
Gall
,
J. Appl. Phys.
124
,
041601
(
2018
).
8.
J. C.
Vickerman
,
K.
Christmann
,
G.
Ertl
,
P.
Heimann
,
F. J.
Himpsel
, and
D. E.
Eastman
,
Surf. Sci.
134
,
367
(
1983
).
9.
C.
Egawa
,
T.
Aruga
, and
Y.
Iwasawa
,
Surf. Sci.
188
,
563
(
1987
).
10.
K.
Barmak
and
R. R.
Gusley
,
J. Electrochem. Soc.
169
,
082517
(
2022
).
11.
M.
Minniti
,
D.
Farías
,
P.
Perna
, and
R.
Miranda
,
J. Chem. Phys.
137
,
074706
(
2012
).
12.
P. W.
Sutter
,
J.-I.
Flege
, and
E. A.
Sutter
,
Nat. Mater.
7
,
406
(
2008
).
13.
P. W.
Sutter
,
P. M.
Albrecht
, and
E. A.
Sutter
,
Appl. Phys. Lett.
97
,
213101
(
2010
).
14.
M. S.
Akselrod
and
F. J.
Bruni
,
J. Cryst. Growth
360
,
134
(
2012
).
15.
S.
Yamada
,
Y.
Nishibe
,
M.
Saizaki
,
H.
Kitajima
,
S.
Ohtsubo
,
A.
Morimoto
,
T.
Shimizu
,
K.
Ishida
, and
Y.
Masaki
,
Jpn. J. Appl. Phys.
41
,
L206
(
2002
).
16.
K.
Barmak
,
K.
Sentosun
,
A.
Zangiabadi
,
E.
Milosevic
,
D.
Gall
,
M.
Zecevic
,
R. A.
Lebensohn
, and
J. A.
Floro
,
J. Appl. Phys.
128
,
045304
(
2020
).
17.
J. E.
Prieto
,
E. M.
Trapero
,
P.
Prieto
,
E.
García-Martín
,
G. D.
Soria
,
P.
Galán
, and
J.
de la Figuera
,
Appl. Surf. Sci.
582
,
152304
(
2022
).
18.
W.
Braun
,
M.
Jäger
,
G.
Laskin
,
P.
Ngabonziza
,
W.
Voesch
,
P.
Wittlich
, and
J.
Mannhart
,
APL Mater.
8
,
071112
(
2020
).
19.
S.
Smink
,
L. N.
Majer
,
H.
Boschker
,
J.
Mannhart
, and
W.
Braun
,
Adv. Mater.
36
,
2312899
(
2024
).
20.
F. V. E.
Hensling
,
W.
Braun
,
D. Y.
Kim
,
L. N.
Majer
,
S.
Smink
,
B. D.
Faeth
, and
J.
Mannhart
,
APL Mater.
12
, 040902 (
2024
).
21.
W.
Braun
and
J.
Mannhart
,
AIP Adv.
9
,
085310
(
2019
).
22.
T. J.
Smart
,
J.
Mannhart
, and
W.
Braun
,
J. Laser Appl.
33
,
022008
(
2021
).
23.
I. N.
Sergeev
,
V. K.
Kumykov
, and
V. A.
Sozaev
,
Bull. Russ. Acad. Sci. Phys.
72
,
1120
(
2008
).
24.
J. E.
Boschker
,
E.
Tisbi
,
E.
Placidi
,
J.
Momand
,
A.
Redaelli
,
B. J.
Kooi
,
F.
Arciprete
, and
R.
Calarco
,
AIP Adv.
7
,
015106
(
2017
).
25.
R. F.
Sekerka
,
Cryst. Res. Technol.
40
,
291
(
2005
).
26.
T.L.
Einstein
, “
Equilibrium shape of crystals
,” in
Handbook of Crystal Growth
(
Elsevier
,
New York
,
2015
), pp.
215
264
.
27.
H.
Okamoto
,
J. Phase Equilib. Diffus.
27
,
426
(
2006
).
28.
D.
Lu
,
Q.
Jiang
,
X.
Ma
,
Q.
Zhang
,
X.
Fu
, and
L.
Fan
,
Crystals
12
,
1549
(
2022
).
29.
S. G.
Kim
,
J. S.
Yang
, and
W. T.
Kim
,
J. Am. Ceram. Soc.
89
,
3194
(
2006
).
30.
M. A.
Herman
,
W.
Richter
, and
H.
Sitter
, “
Solid phase epitaxy
,” in
Epitaxy
, (
Springer, Berlin, Heidelberg, 2004
), pp.
45
62
.
31.
F.
Rincon
and
M.
Rieutord
,
Living Rev. Sol. Phys.
15
,
6
(
2018
).
32.
W.
Braun
,
A.
Trampert
,
V. M.
Kaganer
,
B.
Jenichen
,
D. K.
Satapathy
, and
K. H.
Ploog
,
J. Cryst. Growth
301–302
,
50
(
2007
).
33.
C. V.
Thompson
,
Annu. Rev. Mater. Res.
42
,
399
(
2012
).
34.
P. R.
Gadkari
,
A. P.
Warren
,
R. M.
Todi
,
R. V.
Petrova
, and
K. R.
Coffey
,
Surf. Films
23
,
1152
(
2005
).