Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Letter

Cite This: Nano Lett. 2019, 19, 747−755 pubs.acs.org/NanoLett

Vertical Silicon Nanowire Thermoelectric Modules with Enhanced


Thermoelectric Properties
Seungho Lee,† Kihyun Kim,*,‡ Deok-Hong Kang,§ M. Meyyappan,∥ and Chang-Ki Baek†,‡

Department of Electrical Engineering, Pohang University of Science and Technology (POSTECH), Pohang 37673, Republic of
Korea

Department of Creative IT Engineering and Future IT Innovation Lab, Pohang University of Science and Technology
(POSTECH), Pohang 37673, Republic of Korea
§
Energy Research Group, Research Institute of Industrial Science and Technology (RIST), Pohang 37673, Republic of Korea

See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

NASA Ames Research Center, Moffett Field, Mountain View, California 94035, United States
*
S Supporting Information
Downloaded via CHIANG MAI UNIV on November 12, 2022 at 09:14:42 (UTC).

ABSTRACT: Thermoelectric modules based on silicon


nanowires (Si-NWs) have recently attracted significant
attention as they show an improved thermoelectric efficiency
due to a decrease in thermal conductivity. Here, we adopt a
top-down fabrication method to dramatically reduce the
thermal conductivity of vertical Si-NWs. The thermal
conductivity of a vertical Si-NW is significantly suppressed
with an increasing surface roughness, decreasing diameter, and
increasing doping concentration. This large suppression is
caused by enhanced phonon scattering, which depends on the
phonon wavelength. The boron- and phosphorus-doped
rough Si-NWs with a diameter of 200 nm and surface
roughness of 6.88 nm show the lowest thermal conductivity of 10.1 and 14.8 W·m−1·K−1, respectively, which are 5.1- and 3.6-
fold lower than that of a smooth intrinsic nanowire and 14.8- and 10.1-fold lower than that of bulk silicon. A thermoelectric
module was fabricated using this doped rough Si-NW array, and its thermoelectric performance is compared with previously
reported Si-NW modules. The fabricated module exhibits an excellent performance with an open circuit voltage of 216.8 mV·
cm−2 and a maximum power of 3.74 μW·cm−2 under a temperature difference of 180 K, the highest reported for Si-NW
thermoelectric modules.
KEYWORDS: Vertical silicon nanowire, top-down technique, phonon scattering, thermal conductivity, surface roughness,
thermoelectric module

conductivity, κ is the thermal conductivity, and T is the


A lmost 90% of the world’s energy is generated by fossil fuel
combustion with an efficiency of 30−40% and the rest
is wasted into heat. As the world energy consumption
absolute temperature, respectively. The conversion efficiency
increases with a decreasing thermal conductivity.
continues to rise, thermoelectric energy harvesting, i.e., Silicon nanowires (Si-NWs) have recently attracted much
electricity generation from waste heat, has attracted much attention for chemical and biological sensors,9 photovoltaic
attention as a green technology.1,2 Bulk Bi2Te3 and its alloys cells,10,11 and many other applications due to their small cross-
are the most widely used commercial thermoelectric materials sectional area, large surface-to-volume ratio, and attractive
due to their high conversion efficiency. However, they are not transport properties. They have also been considered as a
suitable for large-scale energy recovery applications due to the strong candidate to enhance the conversion efficiency of
high material cost, insufficient supply, and a manufacturing thermoelectric modules by reducing the thermal conductivity,
method that is not suitable for mass production.3−6 In contrast, while maintaining the electrical properties such as the Seebeck
silicon has the advantages of a low material cost, abundant coefficient and the electrical conductivity of bulk Si.12,13 Most
earth reserves, and matured manufacturing technology for thermal energy within silicon is transported by the lattice
mass production; nevertheless, it has not been used as a vibration rather than by electrons and holes; thus, the thermal
thermoelectric material due to its high thermal conductivity.7,8 conductivity of the Si-NWs is effectively suppressed by phonon
Thermal conductivity of a thermoelectric material is an
important factor determining the conversion efficiency, as Received: September 21, 2018
seen from the dimensionless figure-of-merit (ZT = S2σT/κ), Revised: December 26, 2018
where S is the Seebeck coefficient, σ is the electrical Published: January 13, 2019

© 2019 American Chemical Society 747 DOI: 10.1021/acs.nanolett.8b03822


Nano Lett. 2019, 19, 747−755
Nano Letters Letter

scattering at the NW surface. Several studies have achieved


thermal conductivity suppression by tuning phonon scattering
through the control of diameter, doping concentration, and
surface roughness of the Si-NWs fabricated by the vapor−
liquid−solid (VLS) method13−16 or by electroless etching
(EE).12,17−19 The VLS-grown Si-NWs with a diameter of
several tens of nanometers show about 100 times better ZT
than bulk Si due to reduced thermal conductivity. When the
NW cross section shrinks, the thermal conductivity decreases
due to phonon-boundary scattering. The VLS method can
easily produce Si-NWs with a smaller diameter, but it is not
suitable for mass production since an additional transfer
process is required to fabricate the thermoelectric module. Si-
NWs with a rough surface, produced by the EE method, can
have a further reduced thermal conductivity compared to VLS-
grown Si-NWs.12,20 The surface roughness plays a critical role
to suppress thermal conductivity by generating additional
phonon scattering.12,20−22 As the concentration of impurities
increases, the thermal conductivity decreases and the electrical
conductivity is enhanced. However, the Seebeck coefficient Figure 1. Structural characterization of the vertical Si-NWs with a
degrades at an extremely high doping concentration. There- rough surface. (a) Cross-sectional view scanning electron microscope
fore, the optimal doping concentration, not too high doping (SEM) image of the vertical Si-NW array with a diameter of 200 nm
and height of 2 μm. Inset shows the top-view SEM image. (b−d)
concentration, can improve the performance of Si-NW High-resolution transmission electron microscope (HR-TEM) images
thermoelectric modules. Increasing the doping concentration of the vertical Si-NWs (surface roughness (η) = 2.28, 4.29, and 6.88
of Si-NW or the reaction time significantly in the EE process in nm).
order to reduce the thermal conductivity can degrade the
electrical conductivity and prevent the formation of reprodu-
cible electrical contacts.12,23−25 In contrast to the shortcomings vertical Si-NWs with three different surface morphologies. The
of the VLS and EE techniques, the top-down method, which is SEM images for the 350 nm diameter samples are included in
fully compatible with the current state-of-the-art CMOS the Supporting Information (Figure S1). Two parameters are
technology, can solve the above problems. A well-developed defined to quantify the surface roughness. One is the root-
top-down technique provides an excellent design flexibility and mean-square of the roughness (η), and the other is the average
mass production capability, while allowing precise control of distance between two rough features (L, i.e., correlation
the doping concentration and diameter. Therefore, high- length). The vertical Si-NWs fabricated by three different
performance Si-NW thermoelectric modules can be manufac- etching conditions here have η of 2.28, 4.29, and 6.88 nm,
tured using the top-down method by alternately arranging p- respectively, but they show the similar L value of about 14.3
and n-doped Si-NW arrays with a high density in a chip. nm, as seen in the TEM images (Figure 1b−d). This confirms
In this Letter, we have used a top-down technique to the successful control of the NW surface roughness using the
suppress the thermal conductivity of vertical Si-NWs, addi- top-down technique.
tionally aided by a theoretical investigation. Specifically, the We measured the thermal conductivity of six different NW
thermal conductivity of a vertical Si-NW was controlled for the samples (two diameters and three surface roughness values)
first time by adjusting the surface roughness using dry etching, using the differential 3ω method. The experimental setup for
which has advantages of high reproducibility and mass measuring the thermal conductivity is illustrated in Figure S2.
production. In addition, the diameter and doping concen- First, we measured the thermal conductivity of various film
tration were also controlled to further reduce the thermal samples (SiO2, spin-on-glass (SOG), and polyimide (PI)) and
conductivity. The NW thermal conductivity was measured compared these values with reported data in the literatures to
using a differential 3ω method, and the effect of diameter, confirm the accuracy of our measurement method. The
surface roughness, doping concentration, and temperature was measured values are 1.34, 0.38, and 0.21 W·m−1·K−1 for
theoretically analyzed through the phonon-Boltzmann trans- SiO2, SOG, and PI, respectively, which are consistent with
port equation and various phonon scattering processes. Finally, previous reports26−29 (Figure S3) and confirm the high
a thermoelectric module was fabricated using p-type and n- accuracy and reliability of our measurement method. The
type vertical Si-NW legs connected in series, and its power thermal conductivity of a vertical Si-NW (κSiNW) can be
generation performance (open circuit voltage and power) was calculated using28
assessed. κSiNW = (κcomp − (1 − x) ·κSOG) ·x−1 (1)
Effect of Diameter and Surface Roughness on
Thermal Conductivity. Vertical Si-NWs with different where x is the areal density of the Si-NW, κcomp is the thermal
diameters and surface roughness were prepared using the conductivity of the composite sample, and κSOG is the thermal
top-down method to investigate the thermal conductivity. The conductivity of the SOG. The composite sample consists of the
surface roughness was controlled by tuning the pressure and vertical Si-NWs and the SOG film surrounding the nanowires.
gas flow in the inductively coupled plasma reactive ion etching The areal density of the vertical Si-NWs is obtained from the
(ICP-RIE) reactor. Figure 1 shows the scanning electron top-view SEM image (inset of Figure 1a and Figure S1b).
microscope (SEM) and high-resolution transmission electron Figure 2a,b shows the measured thermal conductivity of
microscope (HR-TEM) images of the 200 nm diameter various vertical Si-NWs with different diameters and surface
748 DOI: 10.1021/acs.nanolett.8b03822
Nano Lett. 2019, 19, 747−755
Nano Letters Letter

Figure 2. Thermal conductivity of vertical Si-NWs with different diameters (DNW) and surface roughness (η). (a) Thermal conductivity of various
vertical Si-NWs (filled symbol: measured data, empty symbol: previously reported data in the literatures, solid line: calculated value using p*). (b)
Thermal conductivity of vertical Si-NWs with DNW of 200 nm (red) and 350 nm (black) as a function of normalized specular parameter p* related
with η (filled symbol: measured data, solid line: calculated value using p*). (c) Phonon mean free path and cumulative thermal conductivity versus
reciprocal of phonon wavelength for vertical Si-NWs with DNW of 200 nm and η = 2.28, 4.29, and 6.88 nm. (d) Phonon mean free path and
cumulative thermal conductivity versus reciprocal of phonon wavelength for the vertical Si-NWs with DNW of 200 nm (red) and 350 nm (blue).

roughness values. Previously reported data for the VLS-grown phonons are specularly reflected at the surface without loss of
Si-NWs are also added in Figure 2a. The thermal conductivity momentum. When p = 1, the boundary scattering does not
is suppressed with decreasing diameter or increasing rough- affect the thermal conductivity due to τb−1 = 0. The surface is
ness; for example, it decreases by about 25.5% for a rough very rough for p = 0, and phonon scattering is purely
surface (η = 6.88 nm) compared with a NW with a smooth diffusive.32 Therefore, p = 0 significantly suppresses the
surface (η = 2.28 nm). In addition, when the diameter shrinks thermal conductivity. The value of p can be calculated using
from 350 to 200 nm, the thermal conductivity decreases by eq 2 and the nanowire surface roughness (η) measured by
27.5%. Especially, the vertical Si-NW with a diameter of 200 TEM, assuming that θ = 0. Figure S4 shows p as a function of
nm and η of 6.88 nm has a lower thermal conductivity than the reciprocal of phonon wavelength for the vertical Si-NWs
that of VLS-grown Si-NWs with a diameter of 115 nm. That is, with η = 2.28, 4.29, and 6.88. It can be seen that p decreases
a vertical rough Si-NW fabricated by our etching recipe can with increasing surface roughness over the entire wave-number
have a smaller thermal conductivity than the VLS-grown case range. Since p is dependent on the phonon wavelength as well
even though the diameter is larger. as surface roughness (Figure S4a), we define a phonon
The effect of diameter and surface roughness on phonon wavelength-independent p (called p*) and use this parameter
scattering can be analyzed using the phonon-Boltzmann for intuitively evaluating the effect of surface roughness on the
equation, boundary scattering rate (τb−1), and Umklapp thermal conductivity of the nanowire. The p* is defined as the
scattering rate (τu−1). The scattering rate is given by τ−1= area of the p normalized by the area assuming p = 1 within
τb−1+ τu−1. The boundary scattering rate is defined as21,30,31 cutoff frequency calculated by Mingo,31,43 as shown in Figure
v 2 2 2 2 S4b. The parameter p* depends on the surface roughness of
τb−1 = b (1 − p), p = e−(16π η / λ )cos θ the nanowire but is independent of the phonon wavelength.
DNW (2)
The vertical Si-NWs with η of 2.28, 4.29, and 6.88 nm have p*
where νb is the sound velocity, DNW is the nanowire diameter, p of 0.41, 0.30, and 0.03, respectively. In Figure 2a,b, the vertical
is the specular parameter, λ is the wavelength, and θ is the Si-NW with η of 6.88 nm has the lowest thermal conductivity
incident angle of the phonon at the surface. The parameter p at the same diameter because p* is close to 0. The solid line
represents the fraction of specularly scattered phonons from indicates the thermal conductivity calculated by substituting p*
the surface with a value generally between 0 and 1. The purely instead of p in eq 2. The solid lines are in good agreement with
specular scattering (p = 1) is an ideal case, in which the the measured thermal conductivities, thus confirming the
749 DOI: 10.1021/acs.nanolett.8b03822
Nano Lett. 2019, 19, 747−755
Nano Letters Letter

Figure 3. Temperature dependence of heat transport for vertical Si-NWs. (a) Thermal conductivity of vertical Si-NWs with different diameters and
surface roughness in the temperature range 300−500 K. (b) Phonon mean free path versus reciprocal of phonon wavelength for bulk Si and vertical
Si-NWs (DNW = 200 and 350 nm) with η = 6.88 nm at a temperature of 300 (blue), 400 (black), and 500 K (red).

validity of our calculation. By using p* instead of p in phosphorus-doped (P-doped) vertical Si-NWs for η = 2.28
calculating thermal conductivity, the thermal conductivity of and 6.88 nm. The doping concentrations of boron and
rough Si-NW according to the surface roughness can be easily phosphorus are 2.3 × 1019 and 1.4 × 1019 cm−3, respectively,
predicted. Using the scattering model with p, we can calculate and the intrinsic Si-NW has a doping concentration of 1015
the phonon mean free path and cumulative thermal cm−3. Regardless of the diameter and roughness, the B-doped
conductivity at room temperature shown in Figure 2c,d as a and P-doped vertical Si-NWs tend to have a lower thermal
function of reciprocal of phonon wavelength. As the surface conductivity compared to the intrinsic Si-NW. More
becomes rougher, the wavelength of the core phonons carrying specifically, the boron ion further contributes to the decrease
heat (corresponding to 20−80% of the cumulative thermal in thermal conductivity of the Si-NW than the phosphorus ion.
conductivity) shifts to shorter phonon wavelength, and the When the rough nanowires with DNW = 200 nm are doped with
mean free path of phonons with long wavelength are boron and phosphorus ions, respectively, the thermal
significantly reduced. The rough surface dominantly reduces conductivity of doped nanowires is 71.5% and 58.4% less
the mean free path of phonons with a larger wavelength, than the intrinsic rough nanowire, respectively. The B-doped
resulting in suppression of thermal conductivity. The surface and P-doped vertical rough Si-NWs (DNW = 200 nm and η =
roughness causes additional phonon scattering, which can 6.88 nm) exhibit the lowest thermal conductivity of 10.1 and
reduce the mean free path, for phonons with a larger 14.8 W·m−1·K−1, respectively, which are about 14.8- and 10.1-
wavelength.33−35 The phonon mean free path is reduced fold lower than that of bulk Si (150 W·m−1·K−1).6 The B-
with a shrinking NW diameter in the entire wavelength range. doped rough Si-NW (DNW = 200 nm and η = 6.88 nm)
When the diameter decreases, phonon scattering becomes exhibits the lowest thermal conductivity. It is well-known that
stronger in all wavelength ranges and their mean free path is when the doping concentration is less than about 1020 cm−3,
suppressed; these contribute to the reduction of thermal the thermal conductivity is suppressed with an increasing
conductivity, as shown in Figure 2d. The temperature doping concentration due to the phonon scattering between
dependence of thermal conductivity for vertical Si-NWs (η = the host atom and the impurity atom.36,37 Figure 4b,c shows
2.28 and 6.88 nm) was investigated in the temperature range the phonon mean free path and cumulative thermal
300−500 K (Figure 3a). The calculated thermal conductivity conductivity vs. the reciprocal of phonon wavelength for the
of bulk Si is also added as a solid line for comparison. Unlike B-doped and P-doped rough nanowire with DNW = 200 and
the bulk Si, the vertical Si-NWs show a negligible change in 350 nm, as calculated using the combined scattering rate as
thermal conductivity in the temperature range 300−500 K τ−1= τb−1+ τu−1+ τi−1 where τi−1 is the impurity scattering rate
because the phonon mean free path is hardly affected by the given by

nV 2 ij δM yz 4 2nV 2Q 0 2γ 2 ij δR yz2 4
temperature (Figure 3b). In contrast, the thermal conductivity
3j
j zz ω + jj zz ω
2

4πv b k M { πv b 3 k R {
is changed in all temperature ranges when changing the τi−1 = τδ m−1 + τδ r −1 =
diameter or the surface roughness. This means that τb−1 is (3)
significantly dominant compared to τu−1 in the scattering rate
of vertical Si-NW (τ−1= τb−1+ τu−1). Therefore, even if the where τδm−1 is the mass-difference impurity scattering rate,
surface roughness increases or the diameter decreases, the τδr−1 is radius-difference impurity scattering rate, n is the
influence of temperature change on the phonon mean free path volumetric concentration, V is the crystal volume of the host
is negligible (Figure S5). atom, M is the mass of the host atom, δM is the difference of
Effect of Doping Concentration on Thermal Con- mass, Q0 is how the nearest and further-out linkages combine
ductivity. The addition of dopants within the Si crystalline in the scattering matrix, γ is the Grüneisen constant, R is the
lattices is an effective method for suppression of thermal radius of the host atom, and δR is the difference of radius. The
conductivity.12,36,37 Figure 4a shows the measured thermal phonon-impurity scattering is dominantly caused by the
conductivity for intrinsic, boron-doped (B-doped), and difference in mass ( δM ) and radius ( δR ) between the host
M R
750 DOI: 10.1021/acs.nanolett.8b03822
Nano Lett. 2019, 19, 747−755
Nano Letters Letter

mainly cause the additional scattering of short wavelength


phonons (Figure 4b,c). Therefore, the phonons with a short
wavelength significantly contribute to the reduction in thermal
conductivity. Impurity scattering is further enhanced when the
vertical rough Si-NW is implanted with boron ions (dose of 2.3
× 1019 cm−3) rather than phosphorus ions with a dose of 1.4 ×
1019 cm−3. Boron and phosphorus ions implanted in Si show a
relationship of ( δMM )B ≈ 6( δMM )P and ( δRR )B ≈ 4( δRR )P. 36
At
the same doping concentration, the B-doped Si-NW has a
lower thermal conductivity since the mass and radius scattering
coefficients of boron are larger than those of phosphorus due
to the large difference in mass and radius between the Si and
boron ion. It is known that the change in the thermal
conductivity due to the change in doping concentration is
negligible, when the doping concentrations is higher than 1018
cm−3 at room temperature.36 Therefore, we can see that the B-
doped rough Si-NW has a lower thermal conductivity than the
P-doped Si-NW due to large mass and radius scattering
coefficients, not a high doping concentration. Ion implantation
is one of the effective methods to reduce the thermal
conductivity of vertical Si-NWs, and the type of dopant
contributes strongly to the reduction of thermal conductivity.
Demonstration of a Thermoelectric Generator. Si-
NWs fabricated by the EE method have a very low thermal
conductivity comparable to amorphous SiO2 and high
ZT.12,16,20 However, it is difficult to make the thermoelectric
module with a large number of electrical series connections
between p-type and n-type Si-NWs because they cannot be
precisely formed at the desired position using the EE method.
Here, we have fabricated a thermoelectric module with many
connections between the p-type and n-type rough Si-NW legs
using the top-down approach and demonstrated its thermo-
electric power generation performance. The thermoelectric
modules were fabricated on a 8 in. SOI wafer (p-type, 1−10 Ω·
cm) using B-doped and P-doped vertical Si-NW structures (η
= 6.88 nm, DNW = 200 nm, pitch of 750 nm, and height of 2
μm). The main fabrication process flow is shown in Figure S9
and the detailed fabrication process is described in the Methods
section. The thermoelectric device with a size of 1 cm × 1 cm
was successfully fabricated, and then the vertical rough Si-NW
thermoelectric module was made by attaching the ceramic
substrate to both sides of the thermoelectric device chip
(Figure 5a).
The open circuit voltage, short circuit current (ISC), and
generated power of the thermoelectric modules were measured
Figure 4. Dopant dependence of the thermal conductivity of vertical using an in-house measurement system built with a commercial
Si-NWs. (a) Comparison of thermal conductivity according to heater plate and heat sink consisting of a commercial Peltier
diameter and dopant in vertical Si-NWs with η = 2.28 and 6.88 cooler and a fan (Figure S6). The temperature on the hot side
nm. (b) Phonon mean free path and cumulative thermal conductivity was adjusted via a PID controller unit driving the heater plate.
versus reciprocal of phonon wavelength for intrinsic (black),
In contrast, the temperature of the Peltier cooler was not
phosphorus-doped (red), and boron-doped (blue) vertical rough Si-
NWs with a diameter of 350 nm. (c) Phonon mean free path and controlled by feedback of the actual temperature data
cumulative thermal conductivity versus the reciprocal of phonon measured by the temperature sensor. The temperature on
wavelength for intrinsic (black), phosphorus-doped (red), and boron- the cold side was set by applying a specific current to the
doped (blue) vertical rough Si-NWs with a diameter of 200 nm. cooler. The heater temperature was increased from 273 to 473
K at 10 K intervals, while the cooler temperature was always
set at 273 K. The temperature difference is calculated from the
atom and impurity atoms. Previously reported ( δM ) and ( δR ) experimental setting values of the hot and cold sides. In order
M R
values for B-doped and P-doped Si were used here for to eliminate the effect of the module size, the open circuit
calculation of the phonon mean free path and the cumulative voltage per unit area (VOC) is defined by dividing the open
thermal conductivity.36,37 The phonon mean free path in the circuit voltage by the module area. Figure 5b shows the
short wavelength is further reduced than that in the long measured VOC of the thermoelectric modules according to the
wavelength since the injected ions (boron and phosphorus) temperature difference, showing a linear increase with the
751 DOI: 10.1021/acs.nanolett.8b03822
Nano Lett. 2019, 19, 747−755
Nano Letters Letter

temperature difference of 180 K. This PMAX is higher than that


of a previously reported Si-NW thermoelectric generator.41
When the surface roughness increases from 2.28 to 6.88 nm,
the rough Si-NW shows about a 40% reduction in the thermal
conductivity compared to the smooth Si-NW. The lower
thermal conductivity can generate a larger temperature
difference within the nanowire structure at a given heat flux,
leading to increased VOC and PMAX.
To clarify the reason for the improved thermoelectric
performance (VOC and PMAX), the effect of surface roughness
on the Seebeck coefficient and the electrical conductivity was
also investigated using the P-doped Si-NW legs with DNW =
200 nm and η = 2.28 and 6.88 nm (Figure S7). The vertical Si-
NWs with η of 2.28 and 6.88 nm show Seebeck coefficients of
266 and 298 μV·K-1, respectively. The rough Si-NW shows a
15% improvement in the Seebeck coefficient and a negligible
reduction in the electrical conductivity compared to the
smooth Si-NW. The B-doped Si-NWs are expected to show a
similar trend to P-doped Si-NWs. Based on these results, the
improved power generation performance (VOC and PMAX) of
the rough Si-NW module is attributed to the low thermal
conductivity and high Seebeck coefficient of the rough Si-NW.
In particular, a significant reduction in thermal conductivity
(∼40%) has the greatest effect on improving the thermo-
electric performance.
In conclusion, we have dramatically reduced the thermal
conductivity of vertical Si-NWs by controlling the surface
roughness, diameter, and doping concentration using a top-
down fabrication method. The vertical Si-NW with a smaller
diameter, rougher surface, and higher doping concentration
can lead to more phonon scattering, resulting in a lower
thermal conductivity. This thermal conductivity reduction was
theoretically analyzed by applying various scattering models to
the phonon-Boltzmann equation. The boundary scattering
becomes stronger as the surface roughness increases or the
diameter decreases, resulting in suppression of thermal
conductivity. When the dopant is implanted into an intrinsic
Si-NW, the phonons with a short wavelength are strongly
scattered due to impurity scattering and cause a drop in
thermal conductivity. The impurity scattering is maximized
when the difference in mass and radius between the Si atom
and the dopant becomes larger. Whereas the thermal
conductivity of bulk Si decreases with an increasing temper-
ature due to phonon scattering at short wavelengths, the
vertical Si-NWs show almost no change in thermal
Figure 5. Power generation performance of a vertical rough Si-NW
conductivity due to temperature change. The B-doped and
thermoelectric module. (a) Photograph of a fabricated thermoelectric P-doped rough Si-NWs (DNW = 200 nm and η = 6.88 nm)
device chip and module with a size of 1 cm × 1 cm. (b) Open circuit exhibit a low thermal conductivity of 10.1 and 14.8 W·m−1·
voltage versus temperature difference for various Si-NW thermo- K−1, respectively, which are 5.1- and 3.6-fold lower than an
electric modules. (c) Output voltage and power generation of the undoped smooth Si-NW (DNW = 200 nm and η = 2.28 nm).
rough Si-NW thermoelectric module for various temperature These vertical rough Si-NWs with a low thermal conductivity
differences. were used to fabricate a thermoelectric module, which
exhibits an excellent open circuit voltage of 216.8 mV·cm−2
temperature difference. The results for various thermoelectric and a power of 3.74 μW·cm−2 under a temperature difference
modules are also included in Figure 5b for comparison.38−41 A of 180 K. The improvement in thermoelectric performance
VOC of 216.8 mV·cm−2 is observed when the temperature comes from the rough Si-NW structure with a significant
difference across the whole experimental setup is 180 K. reduction in thermal conductivity through a control of the
Although the rough Si-NW thermoelectric module has a larger surface roughness, diameter, and doping concentration. Our
diameter and smaller number of legs per unit area than other top-down fabrication method can provide a route toward the
Si-NW modules, it exhibits at least a 1.5 times higher VOC than efficient and cost-effective fabrication of thermoelectric
the rest (Table S1). The generated maximum power (PMAX ) modules for generating power from waste heat.
increases with the temperature difference (Figure 5c). The Methods. Sample Preparation for Measuring Thermal
module shows a PMAX of 3.74 μW and ISC of 63 μA under a Conductivity. The starting material was an 8 in. (100) Si wafer
752 DOI: 10.1021/acs.nanolett.8b03822
Nano Lett. 2019, 19, 747−755
Nano Letters Letter

(p-type, 1−10 Ω·cm), which corresponds to a doping of the fabricated composite film (i.e., vertical Si-NW array filled
concentration of about 1015 cm−3. A 400 nm thick SiO2 with an SOG film) and SOG film samples were obtained by
layer was thermally grown on the Si wafer using a wet measuring the V3ω in the frequency range from 100 Hz to 1
oxidation process. For the formation of the vertical Si-NW, kHz.
circular patterns with diameters of 200 and 350 nm at a pitch Computational Analysis of Thermal Conductivity. The
of 750 nm were defined on the SiO2 layer using a KrF scanner. computational analysis was carried out using the phonon-
The SiO2 layer used as a hard mask was first etched using the Boltzmann transport equation to theoretically study the
ICP-RIE. For the formation of Si-NW, the Si wafer was then thermal conductivity of the vertical Si-NW. The phonon-
etched to a depth of 2 μm using the ICP-RIE with a mixture of Boltzmann equation can generally be used for analyzing heat
SF6, O2, and He gases. The roughness of the vertical Si-NW transport and is given by45,46
was controlled by the chamber pressure and the flow rates of
SF6 and O2. SF6 serves to etch Si, and the O2 provides sidewall kB 4T 3 ℏωc / kBT ey
protection during the etching process. Thus, as the flow rate of κ= 2
2π vbℏ 3 ∫0 τ (T , y )y 4 y
(e − 1)2
dy
(5)
SF6 and pressure increase, the surface of the vertical Si-NW
becomes rougher. In contrast, a smooth surface is formed by where κ is the thermal conductivity, kB is the Boltzmann’s
increasing the flow rate of O2. The surface roughness of three constant, ℏ is the Planck’s constant, ωc is the cutoff frequency
different cases was realized by using three kinds of etching for Si, νb is the phonon group velocity for Si, T is the
conditions. To remove surface damage by the plasma, a 4 nm temperature, τ is the phonon lifetime, and y ≡ ℏω/kBT. The
thick SiO2 was thermally grown as a sacrificial oxide, followed thermal conductivity can be theoretically analyzed by various
by immersing in diluted HF solution. In addition, doped Si- models proposed by Callaway, Holland, and Mingo.43,45,46
NW samples were also prepared to investigate the effect of Among them, Mingo’s model, which is based on linearized
doping concentration on the thermal conductivity. Boron or dispersion with fitting parameters, has been widely utilized for
phosphorus ions with a dose of about 1019 cm−3 were theoretical analysis of Si-NW because this model has shown a
implanted into different Si-NW samples using a tilted ion good agreement with the measured thermal conductivity of the
implantation technique, followed by annealing at 1000 °C for Si-NWs wider than 35 nm.21,33,43 The parameters used in the
90 min. The tilted implantation was performed to uniformly calculation are from Mingo et al.31,43,44 The temperature,
distribute the implanted ions within the entire nanowire, and surface roughness, and impurity can lead to additional phonon
the tilted angle was defined by considering the pitch, diameter, scattering mechanisms in nanostructures to affect the heat
and height of the nanowire. Through three-dimensional atom transport. Therefore, it is necessary to consider the anharmonic
probe tomography (3D APT), we confirmed that the doping (Umklapp), boundary, and impurity scattering in theoretical
concentrations for B-doped and P-doped Si-NWs were 2.3 × calculations to accurately analyze their effects on the thermal
1019 and 1.4 × 1019 cm−3, respectively. The gaps between the conductivity of Si-NW. All scattering processes can be taken
vertical Si-NWs were filled with SOG, which has a low thermal into account using Mathiessen’s rule. Therefore, the combined
conductivity. This sample is called the composite sample. A 30 scattering rate (τ−1) is written as follows: τ−1= τb−1+ τu−1+ τi−1.
nm thick SiO2 film was deposited for electrical isolation, and Fabrication of the Thermoelectric Module. An 8 in. (100)
then the π-shaped metal line serving as a heater and a SOI wafer (p-type, 1−10 Ω·cm) with a top silicon thickness of
thermometer were finally formed on the SiO2 film using a lift- 5 μm and buried oxide thickness of 500 nm was used as the
off technique, as shown in Figure S8a. SiO2, SOG, and PI starting substrate. The thermoelectric device is designed to
samples were also prepared to confirm the accuracy of our have 160 × 160 legs in an area of 1 cm × 1 cm, and each leg is
measurement system. located apart with a gap of 10 μm. Each leg consists of 50 × 50
Thermal Conductivity Measurement. The differential 3ω nanowires with η of 6.88 nm, diameter of 200 nm, height of 2
technique is a well-known method for measuring the thermal μm, and pitch of 750 nm. The main fabrication process flow is
conductivity of a thin film using a π-shaped metal line.26,42 shown in Figure S9. The vertical Si-NW legs comprising a Si-
Figure S8b shows a schematic of the measurement principle. NW array were formed using conventional KrF lithography
When an AC current with the ω frequency is applied to the and ICP-RIE. The vertical rough nanowires were uniformly
metal line, a temperature oscillation of ΔT2ω is generated due fabricated over the large area (Figure S10). The thermoelectric
to the Joule heat and the electrical resistance of the metal line module is generally designed such that the p-type and n-type
is modulated, resulting in the generation of a voltage drop at a legs are arranged to be thermally parallel and electrically in
frequency of 3ω (V3ω). The V3ω was measured using an in- series. The p-type and n-type legs were therefore defined by
house built measurement system consisting of a vacuum probe selectively tilted ion implantation of boron and phosphorus
station with a chuck temperature controller, a signal processing ions so that they were alternately placed. Ion implantation was
circuit, and a lock-in amplifier (SR830), as shown in Figure performed under the same conditions (concentration and
S2a. The ΔT2ω is determined by the V3ω, and the thermal tilted angle) as for the samples for measuring thermal
conductivity of the thin film is given by26 conductivity described earlier. During the formation of p-
Pheater ·t film type legs, the regions of n-type legs were protected from ion
κfilm = implantation of boron using photoresist, and vice versa for the
w·l·ΔTfilm (4) formation of n-type legs. After ion implantation, the bottom
electrode was formed by a selective cobalt silicidation process
where Pheater is the power generated from the metal line, tf ilm is to connect the p-type and n-type legs in series (Figure S10a).
the thickness of the thin film, w and l are the width and length For the selective cobalt silicidation, a 30 nm thick SiO2 film
of the metal line, respectively, and ΔTf ilm is the difference of was grown on the entire wafer surface using the wet oxidation
T2ω between the thin film samples. To extract the thermal process to prevent the silicide formation by blocking the
conductivity of the vertical Si-NW, the thermal conductivities reaction between Si and cobalt. Next, only the SiO2 film at the
753 DOI: 10.1021/acs.nanolett.8b03822
Nano Lett. 2019, 19, 747−755
Nano Letters Letter

position where the cobalt silicide is to be formed was technique of its characterization of a Nd:YAG laser" (2018-
selectively removed using optical lithography and ICP-RIE. 0-01283) supervised by the IITP, the "Smart Industrial Energy
Anisotropic etching was done so that the SiO2 film on the ICT Convergence Consortium" (NIPA-C1601-17-1007) su-
sidewalls of Si-NW was not etched. Next, 50 nm thick cobalt pervised by the NIPA (National IT Industry Promotion
was deposited on the wafer surface, followed by annealing at Agency), and the "Nanomaterial Technology Development
800 °C for the formation of cobalt silicide. Unreacted cobalt Program" (2009-0082580) supervised by the NRF (National
over the SiO2 film was subsequently removed by immersing Research Foundation of Korea).


the wafer in a heated piranha solution (H2SO4:H2O2). 3D APT
measurements confirmed the doping concentrations for B-
REFERENCES
doped and P-doped Si-NWs to be 2.3 × 1019 and 1.4 × 1019
cm−3, respectively, which are the same as samples for (1) Kim, F.; Kwon, B.; Eom, Y.; Lee, J. E.; Park, S.; Jo, S.; Park, S.
measuring the thermal conductivity described earlier. The H.; Kim, B.-S.; Im, H. J.; Lee, M. H.; Min, T. S.; Kim, K. T.; Chae, H.
polyimide was coated and cured at 400 °C to fill the gaps G.; King, W. P.; Son, J. S. 3D printing of shape-conformable
between vertical Si-NWs, and the top of the vertical rough Si- thermoelectric materials using all-inorganic Bi2Te3-based inks. Nature
NWs was then exposed using O2 plasma etching. A 20 nm Energy 2018, 3 (4), 301−309.
(2) Avery, A. D.; Zhou, B. H.; Lee, J.; Lee, E.-S.; Miller, E. M.; Ihly,
thick titanium (Ti) layer and a 200 nm thick gold (Au) layer
R.; Wesenberg, D.; Mistry, K. S.; Guillot, S. L.; Zink, B. L.; Kim, Y.-
were deposited and patterned using a lift-off method to H.; Blackburn, J. L.; Ferguson, A. J. Tailored semiconducting carbon
electrically connect the p-type and n-type legs. After deposition nanotube networks with enhanced thermoelectric properties. Nature
of SiO2 as a passivation layer, the thermoelectric devices were Energy 2016, 1, 16033.
diced into 1 cm × 1 cm. Finally, the ceramic plates were (3) Bell, L. E. Cooling, heating, generating power, and recovering
attached to both the front and back sides of the thermoelectric waste heat with thermoelectric systems. Science 2008, 321 (5895),
device chip using a thermal interface paste to make the 1457−1461.
thermoelectric module. (4) Poudel, B.; Hao, Q.; Ma, Y.; Lan, Y.; Minnich, A.; Yu, B.; Yan,


*
ASSOCIATED CONTENT
S Supporting Information
X.; Wang, D.; Muto, A.; Vashaee, D.; et al. High-thermoelectric
performance of nanostructured bismuth antimony telluride bulk
alloys. Science 2008, 320 (5876), 634−638.
(5) Yazawa, K.; Shakouri, A. Cost-efficiency trade-off and the design
The Supporting Information is available free of charge on the
of thermoelectric power generators. Environ. Sci. Technol. 2011, 45
ACS Publications website at DOI: 10.1021/acs.nano- (17), 7548−7553.
lett.8b03822. (6) LeBlanc, S.; Yee, S. K.; Scullin, M. L.; Dames, C.; Goodson, K. E.
Structural characterization of the vertical Si-NW array, Material and manufacturing cost considerations for thermoelectrics.
thermal conductivity measurement, thermal conductivity Renewable Sustainable Energy Rev. 2014, 32, 313−327.
analysis, device fabrication step, SEM images of the (7) Tang, J.; Wang, H.-T.; Lee, D. H.; Fardy, M.; Huo, Z.; Russell,
fabricated device, thermoelectric module measurement, T. P.; Yang, P. Holey silicon as an efficient thermoelectric material.
Seebeck coefficient, electrical conductivity, and open Nano Lett. 2010, 10 (10), 4279−4283.
circuit voltage comparison of Si-NW modules (PDF) (8) Curtin, B. M.; Codecido, E. A.; Krämer, S.; Bowers, J. E. Field-


effect modulation of thermoelectric properties in multigated silicon
nanowires. Nano Lett. 2013, 13 (11), 5503−5508.
AUTHOR INFORMATION (9) Kim, K.; Park, C.; Kwon, D.; Kim, D.; Meyyappan, M.; Jeon, S.;
Corresponding Author Lee, J.-S. Silicon nanowire biosensors for detection of cardiac
*E-mail: khkim85@postech.ac.kr. troponin I (cTnI) with high sensitivity. Biosens. Bioelectron. 2016,
77, 695−701.
ORCID (10) Ko, M.-D.; Rim, T.; Kim, K.; Meyyappan, M.; Baek, C.-K. High
Kihyun Kim: 0000-0001-7578-7977 efficiency silicon solar cell based on asymmetric nanowire. Sci. Rep.
Author Contributions 2015, 5, 11646.
S.L. carried out fabrication and characterization of the (11) Garnett, E.; Yang, P. Light trapping in silicon nanowire solar
thermoelectric device, analyzed the data, and wrote the cells. Nano Lett. 2010, 10 (3), 1082−1087.
paper. K.K. conceived the idea, designed the experiments, (12) Hochbaum, A. I.; Chen, R.; Delgado, R. D.; Liang, W.; Garnett,
E. C.; Najarian, M.; Majumdar, A.; Yang, P. Enhanced thermoelectric
fabricated and characterized the thermoelectric device,
performance of rough silicon nanowires. Nature 2008, 451 (7175),
analyzed the data, and wrote the paper. D.K. supported the
163−167.
design and measurement of the thermoelectric module. M.M. (13) Boukai, A. I.; Bunimovich, Y.; Tahir-Kheli, J.; Yu, J.-K.;
and C.B. supported the research and provided experimental Goddard, W. A., III; Heath, J. R. Silicon nanowires as efficient
advice. All authors discussed the results and contributed to the thermoelectric materials. Nature 2008, 451 (7175), 168.
writing of the manuscript. (14) Peng, K.-Q.; Wang, X.; Li, L.; Hu, Y.; Lee, S.-T. Silicon
Notes nanowires for advanced energy conversion and storage. Nano Today
The authors declare no competing financial interest. 2013, 8 (1), 75−97.


(15) Li, D.; Wu, Y.; Fan, R.; Yang, P.; Majumdar, A. Thermal
ACKNOWLEDGMENTS conductivity of Si/SiGe superlattice nanowires. Appl. Phys. Lett. 2003,
83 (15), 3186−3188.
We acknowledge the financial support from the MSIT (16) Lee, J.; Lee, W.; Lim, J.; Yu, Y.; Kong, Q.; Urban, J. J.; Yang, P.
(Ministry of Science and ICT), Korea, under the "ICT Thermal transport in silicon nanowires at high temperature up to 700
Consilience Creative program" (IITP-2018-2011-1-00783) K. Nano Lett. 2016, 16 (7), 4133−4140.
supervised by the IITP (Institute for Information & (17) Bin, X.; Khouri, W.; Fobelets, K. Two-sided silicon nanowire
Communications Technology Promotion), the "Development array/bulk thermoelectric power generator. IEEE Electron Device Lett.
of a highly sensitive Si photodetector and optimization 2014, 35 (5), 596−598.

754 DOI: 10.1021/acs.nanolett.8b03822


Nano Lett. 2019, 19, 747−755
Nano Letters Letter

(18) Weisse, J. M.; Marconnet, A. M.; Kim, D. R.; Rao, P. M.; generator with polyimide filling. IEEE Electron Device Lett. 2012, 33
Panzer, M. A.; Goodson, K. E.; Zheng, X. Thermal conductivity in (5), 715−717.
porous silicon nanowire arrays. Nanoscale Res. Lett. 2012, 7 (1), 554. (41) Dávila, D.; Tarancon, A.; Calaza, C.; Salleras, M.; Fernández-
(19) Feser, J. P.; Sadhu, J. S.; Azeredo, B. P.; Hsu, K. H.; Ma, J.; Kim, Regúlez, M.; San Paulo, A.; Fonseca, L. Monolithically integrated
J.; Seong, M.; Fang, N. X.; Li, X.; Ferreira, P. M.; et al. Thermal thermoelectric energy harvester based on silicon nanowire arrays for
conductivity of silicon nanowire arrays with controlled roughness. J. powering micro/nanodevices. Nano Energy 2012, 1 (6), 812−819.
Appl. Phys. 2012, 112 (11), 114306. (42) Cahill, D. G. Thermal conductivity measurement from 30 to
(20) Lim, J.; Hippalgaonkar, K.; Andrews, S. C.; Majumdar, A.; 750 K: the 3ω method. Rev. Sci. Instrum. 1990, 61 (2), 802−808.
Yang, P. Quantifying surface roughness effects on phonon transport in (43) Mingo, N. Calculation of Si nanowire thermal conductivity
silicon nanowires. Nano Lett. 2012, 12 (5), 2475−2482. using complete phonon dispersion relations. Phys. Rev. B: Condens.
(21) Malhotra, A.; Maldovan, M. Impact of phonon surface Matter Mater. Phys. 2003, 68 (11), 113308.
scattering on thermal energy distribution of Si and SiGe nanowires. (44) Mingo, N.; Hauser, D.; Kobayashi, N.; Plissonnier, M.;
Sci. Rep. 2016, 6, 25818. Shakouri, A. Nanoparticle-in-Alloy” approach to efficient thermo-
(22) Moore, A. L.; Saha, S. K.; Prasher, R. S.; Shi, L. Phonon electrics: silicides in SiGe. Nano Lett. 2009, 9 (2), 711−715.
backscattering and thermal conductivity suppression in sawtooth (45) Callaway, J. Model for lattice thermal conductivity at low
nanowires. Appl. Phys. Lett. 2008, 93 (8), 083112. temperatures. Phys. Rev. 1959, 113 (4), 1046.
(23) Bouaïcha, M.; Khardani, M.; Bessais, B. Evaluation of the (46) Holland, M. Analysis of lattice thermal conductivity. Phys. Rev.
electrical conductivity of nano-porous silicon from photolumines- 1963, 132 (6), 2461.
cence and particle size distribution. Mater. Sci. Eng., B 2008, 147 (2−
3), 235−238.
(24) Zhang, T.; Wu, S.; Xu, J.; Zheng, R.; Cheng, G. High
thermoelectric figure-of-merits from large-area porous silicon nano-
wire arrays. Nano Energy 2015, 13, 433−441.
(25) Hochbaum, A. I.; Gargas, D.; Hwang, Y. J.; Yang, P. Single
crystalline mesoporous silicon nanowires. Nano Lett. 2009, 9 (10),
3550−3554.
(26) Lee, S.-M.; Cahill, D. G. Heat transport in thin dielectric films.
J. Appl. Phys. 1997, 81 (6), 2590−2595.
(27) Costescu, R. M.; Bullen, A. J.; Matamis, G.; O’Hara, K. E.;
Cahill, D. G. Thermal conductivity and sound velocities of hydrogen-
silsesquioxane low-k dielectrics. Phys. Rev. B: Condens. Matter Mater.
Phys. 2002, 65 (9), 094205.
(28) Curtin, B. M.; Fang, E. W.; Bowers, J. E. Highly ordered vertical
silicon nanowire array composite thin films for thermoelectric devices.
J. Electron. Mater. 2012, 41 (5), 887−894.
(29) Ghose, S.; Watson, K.; Delozier, D.; Working, D.; Connell, J.
W.; Smith, J.; Sun, Y.; Lin, Y. Thermal Conductivity of Polyimide/
Carbon Nanofiller Blends; Key Engineering Materials, Trans Tech
Publ, 2007; pp 749−752.
(30) Ziman, J. M. Electrons and phonons: the theory of transport
phenomena in solids; Oxford University Press, 1960.
(31) Wang, Z.; Mingo, N. Diameter dependence of SiGe nanowire
thermal conductivity. Appl. Phys. Lett. 2010, 97 (10), 101903.
(32) Casimir, H. Note on the conduction of heat in crystals. Physica
1938, 5 (6), 495−500.
(33) Lee, J.; Lim, J.; Yang, P. Ballistic phonon transport in holey
silicon. Nano Lett. 2015, 15 (5), 3273−3279.
(34) Xie, G.; Guo, Y.; Wei, X.; Zhang, K.; Sun, L.; Zhong, J.; Zhang,
G.; Zhang, Y.-W. Phonon mean free path spectrum and thermal
conductivity for Si1− xGex nanowires. Appl. Phys. Lett. 2014, 104
(23), 233901.
(35) Satyala, N.; Vashaee, D. The effect of crystallite size on
thermoelectric properties of bulk nanostructured magnesium silicide
(Mg2Si) compounds. Appl. Phys. Lett. 2012, 100 (7), 073107.
(36) Asheghi, M.; Kurabayashi, K.; Kasnavi, R.; Goodson, K.
Thermal conduction in doped single-crystal silicon films. J. Appl. Phys.
2002, 91 (8), 5079−5088.
(37) Liu, W.; Asheghi, M. Thermal conduction in ultrathin pure and
doped single-crystal silicon layers at high temperatures. J. Appl. Phys.
2005, 98 (12), 123523.
(38) Li, Y.; Buddharaju, K.; Singh, N.; Lee, S. Top-down silicon
nanowire-based thermoelectric generator: design and characterization.
J. Electron. Mater. 2012, 41 (6), 989−992.
(39) Li, Y.; Buddharaju, K.; Singh, N.; Lo, G.; Lee, S. Chip-level
thermoelectric power generators based on high-density silicon
nanowire array prepared with top-down CMOS technology. IEEE
Electron Device Lett. 2011, 32 (5), 674−676.
(40) Li, Y.; Buddharaju, K.; Tinh, B. C.; Singh, N.; Lee, S. J.
Improved vertical silicon nanowire based thermoelectric power

755 DOI: 10.1021/acs.nanolett.8b03822


Nano Lett. 2019, 19, 747−755

You might also like