The incessant downscaling of building blocks for memory and logic in computer chips requires energy-efficient devices. Thermoelectric-based temperature sensing, cooling as well as energy harvesting could be useful methods to reach reliable device performance with stable operating temperatures. For these applications, complementary metal–oxide–semiconductor (CMOS)-compatible and application ready thin films are needed and have to be optimized. In this work, we investigate the power factor of different phosphorous-doped silicon germanium (SiGe) films fabricated in a 300 mm CMOS-compatible cleanroom. For the thermoelectric characterization, we used a custom-built setup to determine the Seebeck coefficient and sheet resistance. For sample preparation, we used low pressure chemical vapor deposition with in situ doping and subsequent rapid thermal annealing on 300 mm wafers. Thin film properties, such as film thickness (12–250 nm), elemental composition, crystallinity, and microstructure, are studied via spectroscopic ellipsometry, x-ray photoelectron spectroscopy, x-ray diffraction, atomic force microscopy, and TEM. The SiGe-based thin films vary in the ratio of Si to Ge to P and doping concentrations. A power factor of 0.52 mW/m K2 could be reached by doping variation. Our results show that SiGe is a very attractive CMOS-compatible material on the 300 mm wafer level and is immediately ready for production of thermoelectric embedded applications.

Thermoelectric materials are widely used as sensors1 or energy harvesters2 [Seebeck effect: thermopiles or thermoelectric generators (TEGs)] or for heating and cooling, respectively [Peltier effect: thermoelectric coolers (TECs)].3,4 The dimensionless thermoelectric figure of merit (ZT-value) is defined by

(1)

Here, σ describes the electrical conductivity (ρ=1/σ is the specific resistivity), α is the Seebeck coefficient, κ is the thermal conductivity, and T is the absolute temperature. The product PF=α2σ is called power factor (PF) and characterizes the electronic transport properties. As can be seen from Eq. (1), high ZT-values can be achieved for large Seebeck coefficients and low specific resistivity values.5 

The progress in developing thermoelectric materials over the past 30 years is strongly related to the concept of Hicks and Dresselhaus using low dimensionality for increasing α and σ, as well as decreasing κ.6–8 Since the 1990s, this concept has been expanded in terms of defects, critical phenomena, anharmonicity, and spin degree of freedom to maximize the ZT value for several materials.9 Materials such as PbTe,10 Heusler alloys,11 or SnSe12 recently showed the highest values at high temperatures but have the disadvantage that none of them are complementary metal–oxide–semiconductor (CMOS)-compatible.

Silicon germanium (SiGe) is continuously subjected to this thermoelectric optimization route and a well-known material in the semiconductor industry, where it is used as the source/drain or channel material in CMOS (complementary metal–oxide–semiconductor) devices.13,14 Low pressure chemical vapor deposition (LPCVD) SiGe films are also used in solar cells and as the absorption material for bolometer or thermopiles.15–17 The well-established LPCVD (low pressure chemical vapor deposition) technique is commercially used for SiGe thin film depositions in CMOS facilities18,19 and is characterized by the advantages of the high thickness and resistivity uniformity, constant doping profile due to in situ doping, and cost reduction made possible by the use of vertical batch furnaces.20,21 As a CMOS-compatible material, SiGe can be integrated into microelectronics manufacturing lines for thermoelectric approaches in industrial mass production. However, the ZT value improvement has been reported mostly for nanostructured SiGe, such as nanocomposites22,23 and nanowires,24 because the small structures lead to the reduction in lattice thermal conductivity justified by boundary scattering.

Figure 1 shows the comparison of different nanostructures with phosphorous-doped or undoped SiGe and illustrates the resulting increase in ZT values at 300 K. Each approach shows a clear increase in the ZT value due to nanostructuring compared to the respective bulk reference value (blue bar). Another possibility for optimization is modulation doping, where the separation of the charge carriers from ionized dopants reduces their scattering and, thus, increases the charge carrier mobility.25 The transfer of nanoscale approaches into CMOS environments is a challenge in uniformity control and integration. However, optimizing the so-called power factor (PF=α2σ) by varying the doping concentration is another route to improve the thermoelectric performance.26 

FIG. 1.

Comparison of the ZT-values at 300 K for different SiGe-based nanostructured approaches and comparison to the bulk reference value. Bulk (P-doped),22 nanocomposite (P-doped),22 nanowires (without doping),24 and thin film (P-doped).23 

FIG. 1.

Comparison of the ZT-values at 300 K for different SiGe-based nanostructured approaches and comparison to the bulk reference value. Bulk (P-doped),22 nanocomposite (P-doped),22 nanowires (without doping),24 and thin film (P-doped).23 

Close modal

In this work, we investigate the influence of gas flow parameters on the material and thermoelectric properties of phosphorous-doped SiGe thin films with different compositions. We show that by varying the process gas flows, in addition to the stoichiometry, α and σ change over a large range as a result of doping variation.26 Finally, we discuss different CMOS-compatible thermoelectric application scenarios like infrared sensors, Peltier cooling, or energy harvesting.

All samples were deposited on 300 mm Si wafers with 100 nm thermally grown silicon oxide, which acts as an insulation layer. In situ doped SiGe films were grown using the LPCVD technique in a vertical batch reactor A412 (ASM). Silane (SiH4), germane (GeH4) as well as phosphine (PH3) in helium (He) for doping were used as gaseous deposition sources. All 300 mm wafer samples received a rapid thermal annealing (RTA) for charge carriers activation and damage repair after deposition (Helios XP Mattson system).27 

Spectroscopic ellipsometry FX100 (KLA Tencor) was used for determining the thickness with an relative uncertainty of 1% and uniformity of the whole 300 mm wafers. To analyze the crystallographic phase composition, we employed grazing incidence x-ray diffraction (GI-XRD) and used the Bruker Discover D8 tool with Cu Kα source at a grazing angle of 0.5. The measured data were identified by standard reference patterns supplied by the International Center for Diffraction Data (ICDD). The stoichiometry of all samples is analyzed by x-ray photoelectron spectroscopy (XPS) using PHI Quantes Scanning Microprobe (Physical Electronics GmbH, MN, USA) apparatus with Al Kα radiation. For quantification, the relative sensitivity factors (RSFs) of each element were used. These RSFs (based on the standard sensitivity factor developed by Wagner) are obtained from the MultiPak software developed by Physical Electronics (PHI) and already corrected for the x-ray source angle for geometric asymmetry effects and the transmission function of the spectrometer.28 The Si, Ge, and P concentrations were calculated using the areas under the elemental peaks after Shirley background subtraction and using the extracted RSFs of each element. The root mean square (RMS) roughness of all samples was determined by means of atomic force microscopy (AFM) with a Cypher, AsylumResearch, Oxford Instruments tool and was analyzed with the Gwyddion software.

Both the Seebeck coefficient α and the sheet resistance R were measured with an home-built measurement setup. A LabVIEW programmed software automatically controlled measurement and recording of the thermovoltages at individually defined temperature set points. The average temperatures were set automatically to 40 °C at a constant temperature difference ΔT of 5 K. The Seebeck coefficients were calculated by means of the inversion method to eliminate the offset voltage influence.29 To increase statistical accuracy, each sample was measured three times, and the mean value was formed. The estimated uncertainty for α is 10% and for the resistivity 7%.

Unless otherwise described, specimen pieces were broken from the half-radius area of the wafer for the following investigations. Table I gives an overview of all varied relative gas flow process parameters and the corresponding influences on the stoichiometry, thickness uniformity and RMS roughness of the samples.

TABLE I.

Properties of SiGe films used for experimental studies: overview of nominal gas flow parameters with experimentally obtained stoichiometry via XPS, thickness uniformity via ellipsometry, and RMS roughness via AFM.

SamplePH3 (rel. %)GeH4 (rel. %)SiH4 (rel. %)Si (%)Ge (%)P (%)Thickness uniformity (%)RMS/sq (nm)
100 100 100 60 12 28 7.5 2.69 
57 100 100 68 10 22 2.8 4.33 
34 100 100 72 10 18 2.7 3.11 
23 100 100 70 11 19 0.7 4.22 
17 100 100 70 11 19 2.5 3.88 
11 100 100 69 12 19 3.1 1.9 
23 44 100 80 12 1.5 1.5 
23 18 100 90 0.9 1.22 
23 100 21 57 13 30 9.6 20.6 
SamplePH3 (rel. %)GeH4 (rel. %)SiH4 (rel. %)Si (%)Ge (%)P (%)Thickness uniformity (%)RMS/sq (nm)
100 100 100 60 12 28 7.5 2.69 
57 100 100 68 10 22 2.8 4.33 
34 100 100 72 10 18 2.7 3.11 
23 100 100 70 11 19 0.7 4.22 
17 100 100 70 11 19 2.5 3.88 
11 100 100 69 12 19 3.1 1.9 
23 44 100 80 12 1.5 1.5 
23 18 100 90 0.9 1.22 
23 100 21 57 13 30 9.6 20.6 

The deposition of phosphorous-doped SiGe layers is influenced by various adsorption and desorption processes on the substrate surface during pyrolysis of the gaseous precursors SiH4, GeH4, and PH3. For samples #1–6, only the PH3 flow was changed, while for samples #7–9 only the SiH4 and GeH4 flow was changed. The decrease in the PH3 flow of samples #1–6 minimizes the absolute P concentration only for samples #1–3, while the Si concentration increases. In contrast, the Ge content remains relatively constant. By reducing the GeH4 flow, not only the Ge but also the P concentration is reduced as could be observed for samples #7 and 8. On the other hand, in sample #9, the reduction in the SiH4 flow not only increases the Ge concentration but also significantly increases the P incorporation. The larger the GeH4 flow and resulting Ge concentration, the better the P is incorporated. It is assumed that a high phosphine concentration leads to the adsorption of PH3 on silicon surfaces and, after the desorption of hydrogen, a formation of P2 dimers takes place on the substrate surface, which prevents the adsorption of further hydrides and favors the incorporation of P. Significantly lower PH3 flows (samples #3–6) increase the relative GeH4 content and prevent the formation of P2 dimers. There is a uniform incorporation of P in SiGe, whereby the phosphine concentration no longer has a significant effect on the absolute P concentration.30 However, the electrical conductivity of samples 3–6 drops significantly [cf. Fig. 7(a)], which indicates a reduction in the active charge carrier concentration.

The thickness uniformity represents the coefficient of variation in %. A high uniformity value implies a high inhomogeneity in the layer thickness across the 300 mm wafer. The highest inhomogeneity can be seen for sample #1 with the highest PH3 flow and sample #9 with the lowest SiH4 flow. Figure 2 illustrates the very good thickness uniformity of sample #1. Only sample #9 shows a significant increase in RMS while the small thickness uniformity of sample #1 does not affect its RMS value.

FIG. 2.

Thickness distribution of sample #1 over the whole 300 mm wafer. Black points represent measurement positions.

FIG. 2.

Thickness distribution of sample #1 over the whole 300 mm wafer. Black points represent measurement positions.

Close modal

Figure 3 shows the deposition rates depending on the film thickness of all samples from Table I. The deposition rate increases with decreasing PH3 flow and shows the typical flattening of the curve for CVD processes when thicker layers are reached.21,30 With higher PH3 flow, the PH3 partial pressure rises in the reactor and leads to a gas phase decomposition of PH3 and what limits the SiH4 adsorption and prevents the monatomic encapsulation of P in the crystal lattice.30 A decrease in the GeH4 flow also causes a decrease in the deposition rate (samples #7 and 8).

FIG. 3.

Deposition rates of all samples.

FIG. 3.

Deposition rates of all samples.

Close modal

Figure 4 shows the x-ray diffraction patterns and due to clear and distinct reflexes confirming the good crystallinity of all samples. The intensity peaks refer to the lattice planes (111), (220), (311), (400), (331), and (422) as expected for SiGe.20,31 The hkl-values indicate the corresponding sample pattern and represent a cubic crystal system with the space group Fd-3m. The vertical black line at a 2Θ angle of 28.2 shows a drift of the reflex pattern from samples #7–8 (Si90Ge5P5-black line, Si57Ge13P30-pink line) in the direction of larger 2Θ angles, while samples #1 and 2 (Si60Ge12P28-dark blue line, Si68Ge10P22-orange line) shift slightly toward smaller 2Θ values. In the case of either a very large or very small P concentration, a drift can be seen in the reflex pattern. The shift in sample 9 and sample 8 can presumably be attributed to the significant change in the gas flow parameters, which in addition to the change in the stoichiometric ratios also leads to a change in the film strees.32,33

FIG. 4.

X-ray diffraction patterns of deposited phosphorous-doped SiGe samples. A reference pattern from the ICDD database has been added for comparison.

FIG. 4.

X-ray diffraction patterns of deposited phosphorous-doped SiGe samples. A reference pattern from the ICDD database has been added for comparison.

Close modal

From the TEM images in Fig. 5, we can see a polycrystalline structure with small grain sizes that are in the AFM grain size range of 40–210 nm. Figure 5(b) shows smaller grain sizes compared to the AFM data, which is known from other work.22 Surface topography measurements of all samples are performed by AFM, where selection is depicted in Fig. 6. The samples exhibit small grain structures ranging in the nanometer range with a roughened surface. In particular, sample #9 [Fig. 6(c)] shows a clearly different structure in comparison to the others, which is confirmed by the high RMS value caused by the variation of gas flow. Due to the insufficient thickness uniformity and high roughness, the processes of samples #1 and 9 are not sufficient for industrial applicability. A grain size distribution of all samples between 40 and 210 nm was determined on the basis of 30 manual measured grains. In contrast, the remaining samples show a homogeneously structured surface such as sample #3 (Si72Ge10P18). Also the determined RMS roughness (cf. Table I) in combination with the good uniformity proves the beneficial applicability for CMOS devices.

FIG. 5.

TEM images with (a) low and (b) high magnifications of as-deposited sample #3.

FIG. 5.

TEM images with (a) low and (b) high magnifications of as-deposited sample #3.

Close modal
FIG. 6.

Surface topography measured by AFM: (a) sample #3 (Si72Ge10P18), (b) sample #8 (Si90Ge5P5), and (c) sample #9 (Si57Ge13P30).

FIG. 6.

Surface topography measured by AFM: (a) sample #3 (Si72Ge10P18), (b) sample #8 (Si90Ge5P5), and (c) sample #9 (Si57Ge13P30).

Close modal

Figure 7 shows the measured Seebeck coefficient α, electrical conductivity σ, and calculated power factor PF of all nine samples dependent on the applied mean temperature (Tav) of 40 °C. The cross marked data points represent the Seebeck coefficient data, and the circle marked graphs the electrical conductivity. It can be seen from samples #1–6 that α increases and σ decreases with a decreasing PH3 flow. Interestingly, in samples #3–6, the P concentration is constant despite a strong reduction in PH3 flow. From this, it has to be concluded that the decreasing conductivity σ for these samples is caused by a decrease in the active charge carrier concentration. Sample #3 (Si72Ge10P18) with a medium PH3 flow as well as #8 (Si90Ge5P5) with the lowest Ge and P concentrations show the greatest PF. In the case of sample #3, this can be explained by the opposite trend of α and σ or reciprocal ρ, which inevitably reach a maximum in the product of PF. A decrease in the Ge content in sample #8 leads to an increase in the conductivity and, finally, to the enhancement of PF. Otherwise, it is well known34 that the higher Ge content inhibits the P activation and leads to the decrease in σ and decrease in PF sample #9.

FIG. 7.

Thermoelectric characterization of all samples at 40 °C: (a) Seebeck coefficient α (cross-marked) and electrical conductivity σ (circle-marked); (b) power factor PF.

FIG. 7.

Thermoelectric characterization of all samples at 40 °C: (a) Seebeck coefficient α (cross-marked) and electrical conductivity σ (circle-marked); (b) power factor PF.

Close modal

In our investigations, we also looked to see if there is a correlation between thermoelectric properties (Fig. 7) and roughness RMS or thickness uniformity (Table I). However, the measurement results did not reveal any significant relationship. To evaluate the potential of our phosphorous-doped SiGe thin films, we have compared the power factor PF with results from previously published results in Table II. These devices were also intended for the usage in thermoelectric generators (TEGs) or infrared sensors. As can be seen from Table II, the highest PF achieved in this work is about twice the size as previous results. It is also worth mentioning that, compared to silicon, SiGe provides a higher efficiency, e.g., in TEGs, through material optimization in the particular CMOS processes.36 

TABLE II.

Overview of CMOS or MEMS compatible thermoelectric thin film materials at room temperature.

GroupN-type materialPF (mW/m K2)Application
Strasser2  LPCVD SiGe 0.25 TEG 
Wang35  LPCVD SiGe 0.288 TEG 
Roncaglia19  SiGe 0.18 IR sensors 
This work LPCVD SiGe 0.52 TEG/TEC/IR sensors 
GroupN-type materialPF (mW/m K2)Application
Strasser2  LPCVD SiGe 0.25 TEG 
Wang35  LPCVD SiGe 0.288 TEG 
Roncaglia19  SiGe 0.18 IR sensors 
This work LPCVD SiGe 0.52 TEG/TEC/IR sensors 

In this work, we investigate the influence of gas flow process parameters on phosphorous-doped SiGe thin films fabricated under CMOS-compatible conditions in a LPCVD batch reactor. The results show that by optimizing the doping concentration, a maximization of the thermoelectric performance could be achieved. Furthermore, it could be observed that by reducing the GeH4 flow, the Ge concentration is minimized and the resistance decreases. The fully optimized process with an optimal composition of Si72Ge10P18 reached a power factor PF of 0.52 mW/m K2. This value is about twice the size of previously reported results. This proves the promising potential of the used CMOS-compatible fabrication technology for applications on the industrial 300 mm wafer level.

This research was funded by the German Bundesministerium für Wirtschaft (BMWI) and by the State of Saxony in the frame of the Important Project of Common European Interest (IPCEI).

The authors have no conflicts to disclose.

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

1.
J.
Schieferdecker
, “
Infrared thermopile sensors with high sensitivity and very low temperature coefficient
,”
Sens. Actuators, A
47
,
422
427
(
1995
).
2.
M.
Strasser
, “
Miniaturized thermoelectric generators based on poly-Si and poly-SiGe surface micromachining
,”
Sens. Actuators, A
97–98
,
535
542
(
2002
).
3.
Y.
Su
,
J.
Lu
,
D.
Villaroman
,
D.
Li
, and
B.
Huang
, “
Free-standing planar thermoelectric microrefrigerators based on nano-grained SiGe thin films for on-chip refrigeration
,”
Nano Energy
48
,
202
210
(
2018
).
4.
M.
Haras
and
T.
Skotnicki
, “
Thermoelectricity for IoT: A review
,”
Nano Energy
54
,
461
476
(
2018
).
5.
Thermoelectrics Handbook: Macro to Nano
, edited by
D. M.
Rowe
(
CRC/Taylor & Francis
,
Boca Raton
,
2006
).
6.
L. D.
Hicks
and
M. S.
Dresselhaus
, “
Effect of quantum-well structures on the thermoelectric figure of merit
,”
Phys. Rev. B
47
,
12727
(
1993
).
7.
L. D.
Hicks
and
M. S.
Dresselhaus
, “
Thermoelectric figure of merit of a one-dimensional conductor
,”
Phys. Rev. B
47
,
16631
16634
(
1993
).
8.
J. P.
Heremans
,
M. S.
Dresselhaus
,
L. E.
Bell
, and
D. T.
Morelli
, “
When thermoelectrics reached the nanoscale
,”
Nat. Nanotechnol.
8
,
471
473
(
2013
).
9.
J.
He
and
T. M.
Tritt
, “
Advances in thermoelectric materials research: Looking back and moving forward
,”
Science
357
,
eaak9997
(
2017
).
10.
H. J.
Wu
,
L.-D.
Zhao
,
F. S.
Zheng
,
D.
Wu
,
Y. L.
Pei
,
X.
Tong
,
M. G.
Kanatzidis
, and
J. Q.
He
, “
Broad temperature plateau for thermoelectric figure of merit ZT > 2 in phase-separated PbTe0.7S0.3
,”
Nat. Commun.
5
,
4515
(
2014
).
11.
B.
Hinterleitner
,
I.
Knapp
,
M.
Poneder
,
Y.
Shi
,
H.
Müller
,
G.
Eguchi
,
C.
Eisenmenger-Sittner
,
M.
Stöger-Pollach
,
Y.
Kakefuda
,
N.
Kawamoto
,
Q.
Guo
,
T.
Baba
,
T.
Mori
,
S.
Ullah
,
X.-Q.
Chen
, and
E.
Bauer
, “
Thermoelectric performance of a metastable thin-film Heusler alloy
,”
Nature
576
,
85
90
(
2019
).
12.
L.-D.
Zhao
, “
Ultralow thermal conductivity and high thermoelectric figure of merit in SnSe crystals
,”
Nature
508
,
373
377
(
2014
).
13.
K. J.
Kuhn
,
A.
Murthy
,
R.
Kotlyar
, and
M.
Kuhn
, “
Past, present and future: SiGe and CMOS transistor scaling
,”
ECS Trans.
33
,
3
17
(
2010
).
14.
S.
Barraud
,
V.
Lapras
,
M. P.
Samson
,
L.
Gaben
,
L.
Grenouillet
,
V.
Maffini-Alvaro
,
Y.
Morand
,
J.
Daranlot
,
N.
Rambal
,
B.
Previtalli
,
S.
Reboh
,
C.
Tabone
,
R.
Coquand
,
E.
Augendre
,
O.
Rozeau
,
J. M.
Hartmann
,
C.
Vizioz
,
C.
Arvet
,
P.
Pimenta-Barros
,
N.
Posseme
,
V.
Loup
,
C.
Comboroure
,
C.
Euvrard
,
V.
Balan
,
I.
Tinti
,
G.
Audoit
,
N.
Bernier
,
D.
Cooper
,
Z.
Saghi
,
F.
Allain
,
A.
Toffoli
,
O.
Faynot
, and
M.
Vinet
, “
Vertically stacked-nanowires MOSFETs in a replacement metal gate process with inner spacer and SiGe source/drain
,” in
2016 IEEE International Electron Devices Meeting (IEDM)
(
IEEE
,
2016
), pp.
17.6.1
17.6.4
.
15.
P.
van Gerwen
,
T.
Slater
,
J.
Chévrier
,
K.
Baert
, and
R.
Mertens
, “
Thin-film boron-doped polycrystalline silicon70%-germanium30% for thermopiles
,”
Sens. Actuators, A
53
,
325
329
(
1996
).
16.
S.
Abdul Hadi
,
P.
Hashemi
,
N.
DiLello
,
E.
Polyzoeva
,
A.
Nayfeh
, and
J. L.
Hoyt
, “
Thin-film Si1–xGex HIT solar cells
,”
Sol. Energy
103
,
154
159
(
2014
).
17.
S.
Sedky
,
P.
Fiorini
,
K.
Baert
,
L.
Hermans
, and
R.
Mertens
, “
Characterization and optimization of infrared poly SiGe bolometers
,”
IEEE Trans. Electron Devices
46
,
675
682
(
1999
).
18.
A. E.
Franke
,
J. M.
Heck
,
T.-J.
King
, and
R. T.
Howe
, “
Polycrystalline silicon-germanium films for integrated microsystems
,”
J. Microelectromech. Syst.
12
,
160
171
(
2003
).
19.
A.
Roncaglia
and
M.
Ferri
, “
Thermoelectric materials in MEMS and NEMS: A review
,”
Sci. Adv. Mater.
3
,
401
419
(
2011
).
20.
A.
Baudrant
and
M.
Sacilotti
, “
The LPCVD polysilicon phosphorus doped in situ as an industrial process
,”
J. Electrochem. Soc.
129
,
1109
1116
(
1982
).
21.
J. M.
Hartmann
, “
SiGe growth kinetics and doping in reduced pressure-chemical vapor deposition
,”
J. Cryst. Growth
263
,
10
20
(
2001
).
22.
X. W.
Wang
,
H.
Lee
,
Y. C.
Lan
,
G. H.
Zhu
,
G.
Joshi
,
D. Z.
Wang
,
J.
Yang
,
A. J.
Muto
,
M. Y.
Tang
,
J.
Klatsky
,
S.
Song
,
M. S.
Dresselhaus
,
G.
Chen
, and
Z. F.
Ren
, “
Enhanced thermoelectric figure of merit in nanostructured n-type silicon germanium bulk alloy
,”
Appl. Phys. Lett.
93
,
193121
(
2008
).
23.
J.
Lu
,
R.
Guo
, and
B.
Huang
, “
Nanograin-enhanced in-plane thermoelectric figure of merit in n-type SiGe thin films
,”
Appl. Phys. Lett.
108
,
141903
(
2016
).
24.
E. K.
Lee
,
L.
Yin
,
Y.
Lee
,
J. W.
Lee
,
S. J.
Lee
,
J.
Lee
,
S. N.
Cha
,
D.
Whang
,
G. S.
Hwang
,
K.
Hippalgaonkar
,
A.
Majumdar
,
C.
Yu
,
B. L.
Choi
,
J. M.
Kim
, and
K.
Kim
, “
Large thermoelectric figure-of-merits from SiGe nanowires by simultaneously measuring electrical and thermal transport properties
,”
Nano Lett.
12
,
2918
2923
(
2012
).
25.
B.
Yu
,
M.
Zebarjadi
,
H.
Wang
,
K.
Lukas
,
H.
Wang
,
D.
Wang
,
C.
Opeil
,
M.
Dresselhaus
,
G.
Chen
, and
Z.
Ren
, “
Enhancement of thermoelectric properties by modulation-doping in silicon germanium alloy nanocomposites
,”
Nano Lett.
12
,
2077
2082
(
2012
).
26.
F.
Völklein
, “
Review of the thermoelectric efficiency of bulk and thin-film materials
,”
Sens. Mater.
8
(
6
),
389
408
(
1996
), see https://myukk-org.ssl-xserver.jp/SM2017/article.php?ss=10253.
27.
J.
Calvo
, “
LPCVD in-situ doped silicon for thermoelectric applications
,”
Mater. Today
5
,
10249
10256
(
2018
).
28.
C. D.
Wagner
,
L. E.
Davis
,
M. V.
Zeller
,
J. A.
Taylor
,
R. H.
Gale
, and
L. H.
Gale
, “
Empirical atomic sensitivity factors for quantitative analysis by electron spectroscopy for chemical analysis
,”
Surf. Interface Anal.
3
,
211
225
(
1981
).
29.
J.
Liu
,
Y.
Zhang
,
Z.
Wang
,
M.
Li
,
W.
Su
,
M.
Zhao
,
S.
Huang
,
S.
Xia
, and
C.
Wang
, “
Accurate measurement of Seebeck coefficient
,”
Rev. Sci. Instrum.
87
,
064701
(
2016
).
30.
S.-M.
Jang
,
K.
Liao
, and
R.
Reif
, “
Chemical vapor deposition of epitaxial silicon–germanium from silane and germane: II. In situ boron, arsenic, and phosphorus doping
,”
J. Electrochem. Soc.
142
,
3520
3527
(
1995
).
31.
T.-J.
King
and
K. C.
Saraswat
, “
Deposition and properties of low-pressure chemical-vapor deposited polycrystalline silicon-germanium films
,”
J. Electrochem. Soc.
141
,
2235
2241
(
1994
).
32.
Y.-C.
Jeon
,
T.-J.
King
, and
R. T.
Howe
, “
Properties of phosphorus-doped poly-SiGe films for microelectromechanical system applications
,”
J. Electrochem. Soc.
150
,
H1
H6
(
2003
).
33.
C.
Rusu
,
S.
Sedky
,
B.
Parmentier
,
A.
Verbist
,
O.
Richard
,
B.
Brijs
,
L.
Geenen
,
A.
Witvrouw
,
F.
Lormer
,
F.
Fischer
,
S.
Kronm-ller
,
V.
Leca
, and
B.
Otter
, “
New low-stress PECVD poly-SiGe layers for MEMS
,”
J. Microelectromech. Syst.
12
,
816
825
(
2003
).
34.
T.-J.
King
,
J. P.
McVittie
,
K. C.
Saraswat
, and
J. R.
Pfiester
, “
Electrical properties of heavily doped polycrystalline silicon-germanium films
,”
IEEE Trans. Electron Devices
41
,
228
232
(
1994
).
35.
Z.
Wang
,
V.
Leonov
,
P.
Fiorini
, and
C.
Van Hoof
, “
Micromachined thermopiles for energy scavenging on human body
,” in
TRANSDUCERS 2007, 2007 International Solid-State Sensors, Actuators and Microsystems Conference
(
2007
).
36.
S. M.
Yang
and
S. H.
Wang
, “
Development of a thermoelectric energy generator with double cavity by standard CMOS process
,”
IEEE Sens. J.
21
,
250
256
(
2021
).