You are on page 1of 6

Effect of phonon-boundary scattering on

phonon-drag factor in Seebeck coefficient


of Si wire
Cite as: AIP Advances 10, 075015 (2020); https://doi.org/10.1063/5.0016043
Submitted: 01 June 2020 • Accepted: 30 June 2020 • Published Online: 16 July 2020

K. Fauziah, Y. Suzuki, T. Nogita, et al.

ARTICLES YOU MAY BE INTERESTED IN

The Seebeck coefficient and phonon drag in silicon


Journal of Applied Physics 116, 245102 (2014); https://doi.org/10.1063/1.4904925

Phonon properties and thermal conductivity from first principles, lattice dynamics, and the
Boltzmann transport equation
Journal of Applied Physics 125, 011101 (2019); https://doi.org/10.1063/1.5064602

Nanoscale thermal transport


Journal of Applied Physics 93, 793 (2003); https://doi.org/10.1063/1.1524305

AIP Advances 10, 075015 (2020); https://doi.org/10.1063/5.0016043 10, 075015

© 2020 Author(s).
AIP Advances ARTICLE scitation.org/journal/adv

Effect of phonon-boundary scattering


on phonon-drag factor in Seebeck coefficient
of Si wire
Cite as: AIP Advances 10, 075015 (2020); doi: 10.1063/5.0016043
Submitted: 1 June 2020 • Accepted: 30 June 2020 •
Published Online: 16 July 2020

K. Fauziah,1,2,3,a) Y. Suzuki,2 T. Nogita,2,4 Y. Kamakura,5 T. Watanabe,6 F. Salleh,7 and H. Ikeda1,2,4,b)

AFFILIATIONS
1
Graduate School of Science and Technology, Shizuoka University, 3-5-1 Johoku, Naka-ku, Hamamatsu 432-8011, Japan
2
Research Institute of Electronics, Shizuoka University, 3-5-1 Johoku, Naka-ku, Hamamatsu 432-8011, Japan
3
National Laboratory for Energy Conversion Technology, Agency for the Assessment and Application of Technology,
15314 Serpong, Indonesia
4
Graduate School of Engineering, Shizuoka University, 3-5-1 Johoku, Naka-ku, Hamamatsu 432-8561, Japan
5
Faculty of Information Science and Technology, Osaka Institute of Technology, 1-79-1 Kitayama, Hirakata, Osaka 573-0196, Japan
6
Graduate School of Engineering, Waseda University, 3-4-1 Ohkubo, Shinjuku-ku, Tokyo 169-8555, Japan
7
Faculty of Engineering, University of Malaya, 50603 Kuala Lumpur, Malaysia

a)
Author to whom correspondence should be addressed: khotimatul.fauziah@gmail.com
b)
Electronic mail: ikeda.hiroya@shizuoka.ac.jp

ABSTRACT
For highly efficient thermoelectric devices with Si nanostructures, we have fabricated and characterized micro/nanometer-scaled Si wires
preserving the phonon-drag effect in order to observe the impact of phonon-boundary scattering on the phonon-drag factor in its Seebeck
coefficient. The observed phonon-drag factor in the Seebeck coefficient decreases with a decrease in the wire width, which is considered due to
an increase in the boundary scattering of phonons. Since the boundary scattering is characterized by the specularity parameter, we measured
the surface roughness of the wire and evaluated the specularity. It was found that the top surface of the Si wire has higher specularity compared
with the sidewall of the wire in the range of phonon wavelength contributing to the phonon drag. This result qualitatively explains the fact
that the phonon drag in the Seebeck coefficient is hardly affected by the wire thickness with a nanometer order, whereas the wire width
influences it significantly even on a micrometer scale. Moreover, it is demonstrated that the phonon-drag effect in the Seebeck coefficient of
Si nanostructures can be preserved while their thermal conductivity is lowered.
© 2020 Author(s). All article content, except where otherwise noted, is licensed under a Creative Commons Attribution (CC BY) license
(http://creativecommons.org/licenses/by/4.0/). https://doi.org/10.1063/5.0016043., s

I. INTRODUCTION originates from the phonon-surface scattering.9 The phonon-


scattering at the boundaries primarily depends on the surface
Introduction of Si nanostructures into Si-based thermoelectric roughness and the incident phonon wavelength. Diffusive scatter-
devices has been widely studied in order to enhance their perfor- ing occurs when the phonons are reflected by a boundary that
mance by reducing their thermal conductivity through the suppres- has surface roughness comparable to or higher than the phonon
sion of phonon transport.1–4 Several studies reported the reduc- wavelength. This dependency is characterized by the specularity
tion of thermal conductivity in a low dimensional structure owing parameter, p, expressed as10,11
to the promotion of the boundary scattering in phonon trans-
port.5–8 Moreover, Martin et al. reported that the surface roughness
strongly affects the thermal conductivity in thin Si nanowires, which p = exp(−4η2 k2 cos2 θ), (1)

AIP Advances 10, 075015 (2020); doi: 10.1063/5.0016043 10, 075015-1


© Author(s) 2020
AIP Advances ARTICLE scitation.org/journal/adv

where η is the rms roughness, k is the phonon wavenumber, and θ is determined to be 3.6 × 1017 cm−3 by the Hall effect measurement.
the incident angle with respect to the surface normal. The specular- Al electrodes were patterned at both ends of the wire.
ity indicates the probability of the phonon being specularly reflected The Seebeck coefficient was measured by a probe measure-
at the boundary surface. The impact of the surface roughness on ment system equipped with an infrared camera, which is explained
the phonon-boundary scattering has been studied in the scope of in the previous report.22 The Seebeck coefficient was evaluated by
thermal transport.10–14 giving a temperature gradient to the sample using a resistive heater
On the other hand, the Seebeck coefficient S is also influenced near room temperature. The temperature difference was observed
by the phonon transport through the phonon-drag effect, which by an infrared camera placed above the sample. A couple of probes
originates from the momentum transfer to electrons from phonons were directly attached to the Al electrode to measure the ther-
flowing under a temperature gradient. In general, the Seebeck coef- moelectromotive force (TEMF). The Seebeck coefficient was eval-
ficient of a semiconductor comprises a phonon-drag factor Sph as uated from the slope of the linear relationship between the mea-
well as an electronic factor Se originating from carrier transport.15,16 sured TEMF, ΔV = V H − V L , and the temperature difference,
Geballe and Sadhu reported that Sph is significantly observed in ΔT = T H − T L by S = −ΔV/ΔT, where the subscripts H and L indi-
bulk Si at room temperature.17,18 In addition, we previously reported cate the high- and low-temperature region. The uncertainty in the
that in quite thin Si-on-insulator (SOI) layers with thickness below measured Seebeck coefficient evaluated from the deviation between
10 nm, Sph is not influenced by the thickness.19,20 However, Sph is the data and the linearly fitted line was ±0.11 mV/K. The mea-
likely affected by reducing the dimensions of the layer into wire.21,22 surement was performed for several single wires with the same
The phonon-boundary scattering seems to have less impact on Sph dimensions in order to confirm the repeatability of the measure-
in ultrathin Si layers, and its relation to Sph in different dimensions ment. These measured values are included in Fig. 1, as an error
of Si is still not clear. In the present study, we investigate the See- bar.
beck coefficient of Si wires with micro/nanometer dimensions and
discuss the influence of the sample dimensions on its phonon-drag
component from the viewpoint of the specularity. III. RESULTS AND DISCUSSION
Figure 1 shows the measured Seebeck coefficient of the p-type
II. EXPERIMENTAL
Si wire as a function of wire width. The blue dotted line shows the
The Si wires discussed in this study were fabricated in SOI wafer linear fitted and the red broken line indicates the calculated Se of
consisting of a top Si layer, a SiO2 layer, and a p-type Si substrate bulk Si with a carrier concentration of 3.6 × 1017 cm−3 .20,23 The
with the same process as the previous investigation.22 Si wires were arrow at the right axis corresponds to the reported value of the bulk
patterned in the top Si layer of SOI wafer by a lithography method Si with a carrier concentration of ∼1015 cm−3 .17 From Fig. 1, the
with a thickness of 30 nm, a length of 1 mm, and a width of 10 μm, Seebeck coefficient is clearly observed to be higher than the calcu-
5 μm, and 1 μm. The p-type wires were prepared by thermal diffu- lated Se , which indicates the contribution of the phonon-drag effect.
sion of B atoms. The carrier concentration of the B-doped wire was Moreover, the Seebeck coefficient decreases with a decrease in the
wire width, and the widest wire has the Seebeck coefficient that is
close to the value of bulk Si despite its small thickness of 30 nm. This
result seems strange since the phonon-drag contribution in the Si
wires is affected by their width even though they are on a micrometer
order.
Since the decrease in the Seebeck coefficient is considered to
originate from the enhancement of phonon-boundary scattering, the
discrepancy between the case in thin layer and wire can be qual-
itatively explained by considering that the boundary scattering of
phonons is strongly related to the roughness on the wire surface.
In order to clarify this, the specularity characterizing the phonon-
boundary scattering is determined by the surface roughness of the
boundary using Eq. (1). The top surface roughness and sidewall
surface roughness of the wire were observed by using atomic force
microscopy (AFM) and are shown in Fig. 2. The average rms rough-
ness of the top surface was evaluated to be 0.4 ± 0.1 nm. On the other
hand, the sidewall surface roughness was observed by constructing
a contour diagram of the wire edge shown in Fig. 2(b) from the top
view. The rms roughness of the sidewall surface was estimated to be
1.4 ± 0.1 nm.
It has been reported that the phonons contributing to the
FIG. 1. The Seebeck coefficient as a function of wire width. The blue dotted line and phonon-drag factor are spectrally distinct from phonons contribut-
the red broken line represent the linear fitted and the calculated Se , respectively. ing to the thermal conductivity.24 The phonon modes, which sig-
The black arrow indicates the reported value of Seebeck coefficient of bulk Si. The nificantly contribute to the phonon drag, have smaller wavenum-
inset shows the schematic diagram of the Si wire.
bers compared with the phonon modes contributing to the thermal

AIP Advances 10, 075015 (2020); doi: 10.1063/5.0016043 10, 075015-2


© Author(s) 2020
AIP Advances ARTICLE scitation.org/journal/adv

FIG. 2. (a) AFM image of the Si wire sur-


face and (b) contour lines of the sidewall
surface.

conductivity. By calculating the phonon participating in intraval- By considering the specularity, the phonon-boundary scatter-
ley electron–phonon scattering in Si, as shown in Fig. 3, the typical ing rate is defined by12
wavenumber of phonons, which mainly contribute to the phonon
1−p v
drag, kphonon drag , is estimated to be 7 × 106 cm−1 . That is, this τB−1 = , (2)
phonon wavenumber is smaller than the major contributor in the 1+p l
thermal transport kthermal conductivity , which is estimated to be ∼4 × 107 where v is the phonon velocity, and l is the length normal to the
at 300 K.13 This result indicates that the high phonon wavelength is boundary surface, which corresponds to the wire width and thick-
mainly responsible for the phonon-drag effect since the wavenum- ness for sidewall and top surfaces, respectively. The total phonon-
−1 −1
ber of electrons in the Si conduction valley is small and they interact scattering rate τtotal_thickness and τtotal_width were calculated by adding
mostly with small-wavenumber phonons. the phonon-boundary scattering and the phonon–phonon scatter-
−1
Figure 4 shows the calculated specularity of the top and side- ing τph including the phonon–phonon Umklapp scattering τU−1 and
wall surfaces by using Eq. (1) with an assumption of θ = 0○ (normal the normal scattering τN−1 calculated by20,25
incidence) as a function of phonon wavenumber. The specularity of
the top and sidewall surfaces for the phonon modes participating in ̵ 2 2 −H

τU−1 = ω Te 3T , (3)
the phonon-drag is obtained to be 0.73 and 0.02, respectively. These Mv2 H
results indicate that the boundary scattering at the top surface is
much less effective than the sidewall surface in the range of phonon k3B γ2 V 2 3
wavenumber contributing to the phonon-drag effect. On the other τN−1 = ωT , (4)
M h̵2 v5
hand, in the range of phonon wavenumber, which participates in
the thermal conductivity, the specularity is quite low p ≈ 0 (i.e., dif-
fusive scattering). This leads to a significant reduction in the thermal
conductivity due to phonon boundary scattering.

FIG. 4. Specularity as a function of wavenumber. The red and blue lines represent
the specularity of the top and sidewall surfaces, respectively. The dashed-dotted
FIG. 3. Distribution of the phonon wavenumber participating in intravalley electron– and dotted lines indicate the wavenumbers, which mainly contribute to phonon
phonon scattering. drag and thermal conductivity, respectively.

AIP Advances 10, 075015 (2020); doi: 10.1063/5.0016043 10, 075015-3


© Author(s) 2020
AIP Advances ARTICLE scitation.org/journal/adv

where h̵ is the reduced Planck constant, γ is the Grüneisen con-


stant, ω is the angular phonon frequency, T is the absolute tem-
perature, H is the Debye temperature, M is the average mass of Si,
kB is the Boltzmann constant, and V is the volume per atom of Si.
−1
Equations (3) and (4) show that τph has a dependency on temper-
ature. Hence, a higher temperature decreases the mean free time of
phonons, thus leading to the phonon-drag effect. However, although
the average temperature increased to 370 K when the tempera-
ture gradient was applied during the measurement, the relationship
between the TEMF and temperature difference had a good lin-
ear relation. Therefore, the Seebeck coefficient is constant, and the
phonon-drag effect is preserved, at least, up to 370 K. The calculated
mean free time is shown in Fig. 5, as a function of length. In this
calculation, it was assumed that the roughness at the Si/SiO2 inter-
face is identical to the roughness of the top surface of the Si wire.
From Fig. 5, it is found that τ total_width is affected by the wire width
in the range below 10 μm. This is consistent with the experimental
results, where the Seebeck coefficient decreases with a decrease in
the wire width below 10 μm. On the other hand, the thickness affects FIG. 6. The calculated β as a function of the specularity.
τ total_thickness at a smaller length, in the range below 1 μm, which is
qualitatively consistent with the experimental result that the influ-
That is, with an increase in the β value, the critical length, where
ence of τ B_thickness on τ total_thickness at a smaller length than that the
the influence of τ B begins to be significant, becomes small. Figure 6
influence of τ B_width on τ total_width .
shows the calculated β as a function of specularity. In this figure,
However, it is quantitatively discrepant from the fact that the Si
with increasing p, β increases monotonously. Above p = 6, the gradi-
wire as thin as 30 nm has the Seebeck coefficient value close to the
ent of the β–p graph also increases significantly, and then, an abrupt
bulk Si. From Eq. (2),
change in β appears at the high specularity value (p > 0.9). This
1+p result indicates that the critical length significantly decreases even
log τB = log l + log − log v. (5)
1−p if the evaluation of rms roughness fluctuates slightly to be smaller at
p = 0.73, which is more sensitive compared with the case of p = 0.02.
Therefore, the straight line of τ B in Fig. 5 shifts vertically in accor- This is a possible reason for the quantitative discrepancy between the
dance with the term, experimental and theoretical critical lengths in the thickness direc-
1+p
β = log . (6) tion. Thus, it is clarified that the roughness on the wire surface is an
1−p important factor for the phonon-drag effect in the Si Seebeck coef-
ficient. Hence, for preserving the phonon-drag contribution to the
Seebeck coefficient in Si nanowires, the wire surface should be quite
smooth.

IV. CONCLUSIONS
We have fabricated micro/nanometer-scaled Si wires contain-
ing the phonon-drag contribution to their Seebeck coefficient. The
size dependency of the phonon-drag factor of the Seebeck coef-
ficient was clearly observed, which is due to the enhancement of
the phonon-boundary scattering. The calculation of the specular-
ity of the wire surface reveals that the boundary scattering at the
top surface is much less effective than the sidewall surface in the
range of phonon wavenumber contributing to the phonon drag.
Thus, the roughness of the wire surface is an important parameter
for preserving the phonon-drag contribution to the nanostructured
Si Seebeck coefficient while the thermal conductivity is reduced by
nanostructuring.

FIG. 5. The calculated phonon mean free time as a function of length. The black
ACKNOWLEDGMENTS
and red solid lines represent the total mean free time considering the specularity
of the thickness and width directions, respectively. The black and red dotted lines The authors thank the Ministry of Research, Technology and
represent the mean free time of the phonon-boundary scattering considering the Higher Education, the Republic of Indonesia, for the RISET-Pro
specularity of the thickness and width directions, respectively. The blue dotted line
scholarship. This work was financially supported in part by JST
represents the mean free time of the phonon–phonon scattering.
CREST Grant No. JPMJCR19Q5, Japan.

AIP Advances 10, 075015 (2020); doi: 10.1063/5.0016043 10, 075015-4


© Author(s) 2020
AIP Advances ARTICLE scitation.org/journal/adv

10
DATA AVAILABILITY J. M. Ziman, Electrons and Phonons (Clarendon Press, Oxford, 1960).
11
A. Malhotra and M. Maldovan, Sci. Rep. 6, 25818 (2016).
The data that support the findings of this study are available 12
J. Anaya, T. Rodríguez, and J. Jiménez, Sci. Adv. Mater. 7, 1097 (2015).
from the corresponding author upon reasonable request. 13
X. Wang and B. Huang, Sci. Rep. 4, 6399 (2014).
14
A. A. Maznev, Phys. Rev. B 91, 134306 (2015).
REFERENCES 15
C. Herring, Phys. Rev. 96, 1163 (1954).
1 16
L. D. Hicks and M. S. Dresselhaus, Phys. Rev. B 47, 8 (1993). G. D. Mahan, L. Lindsay, and D. A. Broido, J. Appl. Phys. 116, 245102 (2014).
2 17
A. Hochbaum, R. Chen, R. Delgado, W. Liang, E. C. Garnett, M. Najarian, T. H. Geballe and G. W. Hull, Phys. Rev. 98, 940 (1955).
18
A. Majumdar, and P. Yang, Nature 451, 163 (2008). J. Sadhu, H. Tian, J. Ma, B. Azeredo, J. Kim, K. Balasundaram, C. Zhang, X. Li,
3 P. M. Ferreira, and S. Sinha, Nano Lett. 15, 3159 (2015).
A. I. Boukai, Y. Bunimovich, J. Tahir-Kheli, J.-K. Yu, W. A. Goddard III, and J. R.
19
Heath, Nature 451, 168 (2008). F. Salleh, K. Asai, A. Ishida, and H. Ikeda, Appl. Phys. Express 2, 071203
4 (2009).
E. B. Ramayya, L. N. Maurer, A. H. Davoody, and I. Knezevic, Phys. Rev. B 86,
20
115328 (2012). F. Salleh, T. Oda, Y. Suzuki, Y. Kamakura, and H. Ikeda, Appl. Phys. Lett. 105,
5 102104 (2014).
D. Li, Y. Wu, P. Kim, L. Shi, P. Yang, and A. Majumdar, Appl. Phys. Lett. 83,
21
2934 (2003). F. Salleh, T. Oda, Y. Suzuki, Y. Kamakura, and H. Ikeda, Makara J. Tech. 19, 1
6 (2015).
R. Chen, A. I. Hochbaum, P. Murphy, J. Moore, P. Yang, and A. Majumdar, Phys.
22
Rev. Lett. 101, 105501 (2008). K. Fauziah, Y. Suzuki, Y. Narita, Y. Kamakura, T. Watanabe, F. Salleh, and
7 H. Ikeda, IEICE Trans. Electron. E102.C, 475 (2019).
D. L. Joulain, D. Terris, D. Lemonnier, D. Lacroix, and K. Joulain, Appl. Phys.
23
Lett. 89, 103104 (2006). E. Behnen, J. Appl. Phys. 67, 287 (1990).
8 24
M. Asheghi, M. N. Touzelbaev, K. E. Goodson, Y. K. Leung, and S. S. Wong, J. Zhou, B. Liao, B. Qiu, S. Huberman, K. Esfarjani, M. S. Dresselhaus, and
J. Heat Transfer 120, 30 (2008). G. Chen, Appl. Phys. Sci. 112, 14777 (2015).
9 25
P. Martin, Z. Aksamija, E. Pop, and U. Ravaioli, Phys. Rev. Lett. 102, 125503 D. T. Morelli, J. P. Heremans, and G. A. Slack, Phys. Rev. B 66, 195304
(2009). (2002).

AIP Advances 10, 075015 (2020); doi: 10.1063/5.0016043 10, 075015-5


© Author(s) 2020

You might also like