Skip to main content
iScience logoLink to iScience
. 2022 Oct 17;25(11):105386. doi: 10.1016/j.isci.2022.105386

Engineering thermal transport within Si thin films: The impact of nanoslot alignment and ion implantation

Sien Wang 1, Yue Xiao 1, Qiyu Chen 1, Qing Hao 1,2,
PMCID: PMC9636053  PMID: 36345333

Summary

In recent years, nanoporous Si films have been intensively studied for their potential applications in thermoelectrics and the thermal management of devices. To minimize the thermal conductivity, ultrafine nanoporous patterns are required but the smallest structure size is largely limited by the spatial resolution of the employed nanofabrication techniques. Along this line, an effectively smaller characteristic length of a nanoporous film can be achieved with offset nanoslot patterns. Compared with periodic circular pores, the nanoslot pattern can achieve an even lower thermal conductivity, where a much smaller porosity is required using ultra-narrow nanoslots. The obtained low thermal conductivity can be understood from the thermally dead volume revealed by phonon Monte Carlo simulations. To further minimize the contribution from short-wavelength phonons, an additional 25% thermal conductivity reduction can be achieved with Ga ions implanted using a focused ion beam.

Subject areas: Nanotechnology, Thermal engineering, Thermal property

Graphical abstract

graphic file with name fx1.jpg

Highlights

  • Offset nanoslot patterns to minimize the in-plane thermal conductivity of thin films

  • Ion implantation to further reduce the thermal conductivity

  • Redistributed heat fluxes revealed by phonon Monte Carlo simulations


Nanotechnology; Thermal engineering; Thermal property

Introduction

Better understanding and then “taming” the thermal transport inside various materials and devices are critical to many applications, including thermoelectric (TE) energy conversion (Shi et al., 2020), cooling of power electronic devices (Hao et al., 2018c; Won et al., 2013), and heat guide (Anufriev et al., 2017; Xiao et al., 2021). Among these, TE materials have a great potential in various energy harvesting and refrigeration applications (Shi et al., 2020). Currently, TE power generators have been developed to produce electricity from the waste heat in the car exhaust gas (Yang and Stabler, 2009) or from solar-radiation heat as a cheap alternative to solar cells (Kraemer et al., 2011). The effectiveness of a TE material is indicated by its dimensionless TE figure of merit, defined as ZT=S2σT/k, where S is the Seebeck coefficient, σ is the electrical conductivity, k is the thermal conductivity, and T is the absolute temperature. Here the thermal conductivity k can be split into the lattice (phonon) contribution kL and the electronic contribution kE. In practice, it has been a challenge to achieve both a low k and a high S2σ within the same material (Slack, 1995). After decades of research, there are still very limited choices for high-performance and abundant TE materials. This material restriction can now be opened up by decoupling the electron and phonon transport using multi-length scale structures (Biswas et al., 2012; Bux et al., 2009; Lan et al., 2009; Poudel et al., 2008). The structures introduced into the materials include point defects such as alloy atoms (Wood, 1988), embedded nanostructures, and nano- to micro-grain boundaries (Hao and Garg, 2021; Vineis et al., 2010; Wan et al., 2010). These structures can minimize the kL across the whole phonon spectrum, while maintaining bulk-like electrical properties. Following this, a high ZT can be potentially achieved in a variety of unconventional materials with a large power factor S2σ, particularly those made of low-cost, abundant, and nontoxic elements. One notable example can be nanostructured bulk silicon (Si) as cost-effective TE materials (Bux et al., 2009; Hao et al., 2010; Kashiwagi et al., 2019).

Considering the difficulty in exactly controlling the nanostructures within a synthesized bulk material, fabricated nanoporous thin films (Hao and Xiao, 2020; Hao et al., 2018a; Lee et al., 2017; Lim et al., 2016; Marconnet et al., 2012; Tang et al., 2010; Yu et al., 2010) and graphene (Gunst et al., 2011; Oh et al., 2017; Xu et al., 2019) have been intensively studied to demonstrate the ZT improvement with the nanostructuring approach. Tremendous attention has been paid to nanoporous Si thin films that can be directly used for the thermal management of semiconductor devices. In such films, diffuse nanopore-edge scattering of phonons is usually dominant and kL can be reduced by classical phonon size effects, whereas wave effects may become important at ∼10 K or below (Graczykowski et al., 2017; Lee et al., 2017; Maire et al., 2017; Nomura et al., 2022; Wagner et al., 2016; Xiao et al., 2019). Following the Matthiessen’s rule, nanopores can modify the phonon mean free paths (MFPs) with a characteristic length (LC) (Hao et al., 2016a; Xiao et al., 2019). To maximize the ZT, this LC should be longer than majority electron MFPs but shorter than majority phonon MFPs (Hao et al., 2016b). In a similar case for nanocrystalline materials, LC becomes the grain size and the recommended LC value is 20 nm for Si at a carrier concentration of 1×1019 cm−3 (Qiu et al., 2015). In practice, challenges exist in achieving such a small LC in bulk materials due to significant grain growth during the hot press (Lan et al., 2009). For thin films, the effective LC can be estimated with the radiative mean beam length (MBL) for optically thin samples, defined as MBL=4Vsolid/Apore (Hao et al., 2016a, 2017a). Here, Vsolid is the solid region volume and Apore is the pore surface area. For aligned circular through-film pores, the recommended LC = 20 nm requires a pore diameter of 6.77 nm with a typical 25% porosity. In nanofabrication, such sub-10 nm patterns can be defined with electron beam lithography (EBL) but the actual pore diameter d is often restricted by the film thickness t and aspect ratio of dry etching, i.e., d>t/3 (Marconnet et al., 2013). Further reducing the nanopore size and spacing may also lead to the degradation of the mechanical strength (Winter et al., 2017). Keeping this in mind, the nanoporous patterns should be optimized to minimize the effective LC (Romano and Grossman, 2014). Among various patterns, long and narrow rectangular nanopores (i.e., nanoslots) exhibit great potentials in thermal designs, i.e., a smaller neck width w between adjacent nanoslots to yield a smaller LC (Hao and Xiao, 2020; Hao et al., 2019). This LC can be further reduced by offsetting adjacent rows of nanoslots (Figure 1A), which also adjusts the in-plane thermal anisotropy of a thin film (Xiao and Hao, 2021). Perpendicular to the overall heat flow direction, the pitch p is fixed at 500 nm for all devices. The dimensions for a device can be uniquely described with (l, b, w) in Figure 1A. In general, nanoslot offsets can vary from zero (i.e., aligned nanoslots) to half of the pitch p (i.e., staggered nanoslots). In this work, the focus is on staggered nanoslots to minimize the in-plane thermal conductivity perpendicular to nanoslot rows. For circular, square, and rectangular nanopores, comparisons between aligned and staggered patterns can be found (Huang et al., 2017; Romano and Kolpak, 2017; Song and Chen, 2004; Tang et al., 2013; Verdier et al., 2017). The use of nanoslots also introduces a “thermally dead volume” with little contribution to heat transfer, as found in similar structures such as nanoladders (Park et al., 2018) and fishbone-like structures (Maire et al., 2018; Yang et al., 2019b). Such a thermally dead volume can also be revealed by the simulated heat flux distribution within the structure, which can be obtained by solving the phonon Boltzmann transport equation (BTE). With a similar thermally dead volume, zigzag Si nanowires exhibited a kL reduction compared with the straight Si nanowire (Anufriev et al., 2019; Heron et al., 2010; Yang et al., 2019a; Zhao et al., 2019). The reduction ranges from ∼5% to 25% at the room temperature and is up to 50% at 25 K (Yang et al., 2019a; Zhao et al., 2019). In a separate study, no kL suppression was observed above 150 K due to the relatively large structure size compared with the phonon MPFs (Anufriev et al., 2019). Following this, heat transfer is mainly along a zigzag path across offset nanoslots (Figure 1B) to achieve a similar kL reduction. Compared with individual zigzag nanowires, a nanoporous film can transport significantly more heat and electrical currents, along with advantages in the device fabrication.

Figure 1.

Figure 1

Demonstration of a nanoporous thin film with offset nanoslots

(A) Offset nanoslots patterned across a thin film. The geometry parameters are defined in the inset.

(B) Heat flux intensity map of a typical offset pattern, with overall heat flowing from the left end to the right end. The magnitude of the heat flux is taken as common logarithm in the plot.

In addition to the nanoporous pattern variation, ion implantation by a focused ion beam (FIB) may also be used as an effective method to locally decrease the kL. Such ion implantation can introduce point defects such as silicon vacancies and interstitials (Lim et al., 2016). For ion irradiation with sufficient energy and/or beyond a threshold dose, amorphization can also be observed to lower the kL (Scott et al., 2020; Tamura et al., 1986; Xiao et al., 2020; Zhao et al., 2017). For the doping purpose, an one order of magnitude reduction in the k was reported for 100-nm diameter ZnO nanowires with implanted gallium ions (Ga+), with a simultaneous one order of magnitude increase in the σ (Xia et al., 2014). The local k value along a single Si nanowire was tuned with selective helium ion irradiation (Zhao et al., 2017). For a single-crystal Si substrate, doses of 1012 to 1018 Ga+/cm2 led to amorphization within the top surface layer with a typically <70 nm thickness. The lowest out-of-plane k was about 1 W/m⋅K at a dose of roughly 1016 Ga+/cm2 for the irradiated Si volume (Alaie et al., 2018). In practice, ion implantation was used to dope nanoporous Si films, and further k reduction was found (Lim et al., 2016).

In this work, the in-plane k of Si thin films with varied nanoslot patterns was systematically studied. Figure 2 shows a typical device used for 3ω measurements. The typical data analysis to extract the k value is presented in Figure 3. The detailed dimensions for all fabricated samples are summarized in Table 1.

Figure 2.

Figure 2

SEM image of a Si thin film with offset nanoslot (sample 1o)

Terminals for the 4-probe measurement are labeled.

Figure 3.

Figure 3

3 ω voltage of Sample 1a at 300 K

Circles are experimental data, and the solid curve is the fitting of the experimental data with Equation 3.

Table 1.

Critical dimensions for four groups of nanoslot-patterned Si thin films

Sample index l (nm) p (nm) b (nm) w (nm)
1a 1000 500 200 80
1o
2a 500 500 200 80
2o
3a 250 500 100 160
3o
4o 1000 500 120 160
4o+
4o++

The definition of geometry parameters (pitches l and p, depth b, and neck width w) is given in Figure 1. The subscript “a” stands for aligned nanoslots and “o” stands for offset nanoslots. Samples 4o+ and 4o++ are further implanted with Ga+ ions with a nominal dose of 2.1×1015 cm−2 and 2.1×1016 cm−2, respectively.

The aligned nanoslot patterns can be cut into fishbone nanowires or nanoladders as a strip along the heat conduction direction (Hao and Xiao, 2020). More controls of the in-plane k are achieved by changing the nanoslot alignment. An up to 33.7% k reduction from aligned to offset nanoslot patterns was found at 125 K, which was comparable to the 34% k reduction from straight to zigzag nanowires at 75 K (Yang et al., 2019a). The experimental result was consistent with phonon Monte Carlo (MC) simulations. When Ga+ ions were implanted with an FIB, the k was further decreased by 25% for one heavily doped sample. By minimizing the energy for ion irradiation, no amorphization was found in transmission electron microscopy (TEM) studies on ion-implanted thin films. These implanted ions simply function as point defects to dramatically scatter short-wavelength phonons. The finding here shows the potential of combining nanofabrication and ion implantation to locally tune the in-plane thermal conductivity of a thin film or atomic-thick material.

Results and discussion

Phonon MFP distributions and heat flux maps

Phonon MC simulations were used to compute the kL of nanoporous thin films, along with the temperature and heat flux profiles under an applied temperature difference. Figure 4 presents the accumulated in-plane kL as a function of the in-plane phonon MFP (Λin) (Hao et al., 2019, 2020) for a 70-nm-thick solid film at 300 K, where 67% of the kL is contributed by phonons with Λin less than 100 nm. This Λin is modified from the bulk phonon MFP (ΛBulk) using the Fuchs-Sondheimer model (Chen, 2005) to incorporate the influence of possibly diffusive phonon scattering at the top/bottom film surfaces, given as

ΛinΛBulk=13[1P(λ)]ΛBulk2h01(xx3)1exp(hΛBulkx)1P(λ)exp(hΛBulkx)dx (Equation 1)

where λ is the phonon wavelength, and P(λ) is the specularity of film-surface phonon reflection for a film with its thickness h. For our measured temperature range, completely diffusive film-surface phonon scattering can be assumed, i.e., P(λ) 0. The computed Λin distribution can be used to compute the in-plane thermal conductivity of thin-film-based nanostructures, such as etched nanowires and nanoporous thin films (Hao et al., 2020). For the 70 nm film thickness, our calculated Λin distribution is similar to the Λin distribution computed based on first-principles (first P) ΛBulk given by Esfarjani et al. (2011) but the temperature-dependent phonon MFPs are not available for first P calculations. In our previous study, a good agreement was found between temperature-dependent measurements on nanoporous Si thin films and phonon MC simulations using fitted phonon MFPs by Wang et al. (2011). The phonon MFPs for a 100-nm-thick Si film were also predicted by Malhotra and Maldovan using a Boltzmann-transport-based reduced MFP model (Malhotra and Maldovan, 2016). In comparison, the ΛBulk distribution given in a separate study (Jain et al., 2013) was converted into the Λin distribution for a 70-nm-thick thin film. For a 145 nm film thickness, Anufriev et al. reconstructed the phonon MFPs from the measured in-plane k of Si films with nanoslits of various aperture sizes (Anufriev et al., 2020).

Figure 4.

Figure 4

Room temperature phonon MFP distributions for the in-plane heat conduction along a 70-nm-thick Si thin film

The red dashed line is computed with the fitted bulk phonon MFPs by Wang et al. (Wang et al., 2011), and the black solid line is based on first P calculations by Esfarjani et al. (Esfarjani et al., 2011). The compared room temperature MFP distributions are predicted for 100-nm-thick Si thin films (Malhotra and Maldovan, 2016), predicted for a 70-nm-thick Si thin film using bulk-Si phonon MFPs given by Jain et al. (Jain et al., 2013), and measured for 145-nm-thick Si films (Anufriev et al., 2019).

In practice, the phonon MFP distribution can be used to guide the thermal designs of nanoslot patterns. When the (lb) value in Figure 1A is much longer than majority phonon Λin values, little kL difference can be found between aligned [Figure 5A for Sample 1a (1000,200,80)] and offset [Figure 5B for Sample 1o (1000,200,80)] nanoslot patterns. In this situation, phonons coming out of the neck region can spread out in the middle of the solid region and then converge at the entrance of the following neck. The thermal resistance of the entire structure can be largely attributed to the ballistic thermal resistance for phonons passing through the mostly “isolated” narrow necks between adjacent nanoslots (Hao et al., 2019; Prasher, 2005). The nanoslot alignment has limited influence here. In contrast, clearly defined heat transfer “channels” can be found when the (lb) value is comparable to or smaller than majority phonon Λin values (Figures 5C and 5D). For Sample 3a (250,100,160) (Figure 5C), heat flow is mostly along straight channels similar to parallel nanowires, as suggested in one previous study on nanoporous Si films (Yu et al., 2010). The interconnecting parts between channels are thermally inactive for most of their volume. When the nanoslots are offset [Figure 5D for Sample 3o (250,100,160)], phonons are enforced to flow along a zigzag path with larger thermal resistance than that for straight channels in Figure 5C.

Figure 5.

Figure 5

Simulated temperature profiles (in Kelvin)

The arrows are the local heat flux. Scale bars are 250 nm.

(A) Temperature profile for Sample 1a (1000,200,80).

(B) Temperature profile for Sample 1o (1000,200,80).

(C) Temperature profile for Sample 3a (250,100,160).

(D) Temperature profile for Sample 3o (250,100,160).

Temperature-dependent kL measured for varied nanoslot patterns

The device layer of an SOI wafer has a low electrical conductivity σ = 5–100 S/m so that kkL can be approximated. The measured in-plane kL with different nanoslot arrangements at different temperatures are plotted as symbols in Figure 6. The maximum temperature rise in the self-heating 3ω measurements is generally controlled within 5 K to avoid the uncertainty in the sample temperature.

Figure 6.

Figure 6

Temperature-dependent kL for different nanoslot-patterned Si thin films

The geometry for each sample is listed in Table 1. Symbols are experiment (exp.) data from the 3ω measurements and curves are predictions from the phonon MC simulation. Error bars are estimated for the 95% confidence interval.

Consistent with simulation results in Figure 5, the measured kL values show small difference between Sample 1a (1000,200,80) and Sample 1o (1000,200,80) (red symbols in Figure 6) with l-b≈ 800 nm. For samples in Group 2 (500,200,80) (green symbols), a reduced lb 300 nm yields a ∼7.3% reduction in kL at room temperature and a 11.7% reduction at 83 K by offsetting the nanoslots. In Group 3 (250,100,160) with lb≈ 150 nm (blue symbols), a 30.2% kL reduction at room temperature and 33.7% at 125 K is observed for the offset nanoslot pattern. In this situation, phonons are enforced to travel along interconnected nanowires and experience more phonon boundary scattering at nanoslot edges.

The suppression factor, defined as =1(kL,O/kL,A), is used to quantify the impact of the nanoslot offset (Figure 7). Due to the increased Λin at decreased temperatures, the suppression effect becomes stronger for Groups 2 (500,200,80) and 3 (250,100,160) at cryogenic temperatures.

Figure 7.

Figure 7

Temperature-dependent suppression factors for two groups of nanoslot patterns. Error bars are estimated for the 95% confidence interval.

Figure 8 shows the correlation between l and , with a fixed p = 500 nm. Based on phonon MC simulations (solid curves), increases rapidly when l is decreased below 300 nm. An over 90% suppression is found for lb = 20 nm, which can largely benefit thermoelectric applications. The for square nanopores (Tang et al., 2013), circular nanopores (Verdier et al., 2017), and segmented silicon nanowires (NWs) (Anufriev et al., 2019) are also added in comparison. In addition, Gluchko et al. also studied the packman-shaped nanopores (Gluchko et al., 2019). The staggered pattern led to a of 17% at 4 K.

Figure 8.

Figure 8

Suppression factors for nanoporous Si thin films with various pore designs

The (l, b, w) values are indicated in the legend.

Impact of crystal defects induced by ion implantation

Temperature-dependent kL for ion-implanted samples was also measured as Group 4 (1000,120,160) (Figure 6). Compared with Sample 4o without ion implantation, the room temperature kL is reduced by 10.8% and 20.0% for Sample 4o+ and Sample 4o++, respectively. The reduction factors further increase to 13.9% and 30.0% at 100 K. In phonon MC simulations, the scattering coefficient A1 for point-defect scattering is increased to match the experimental data. Here A1=1.69×1045s3 for the undoped Si was increased to 8.00×1045s3 and 1.80×1044s3 for Samples 4o+ and 4o++, respectively.

For ion implantation studies, solid Si thin films were released from the SOI and transferred onto TEM grids. The detailed wet transfer process was similar to those used in the film-wafer bonding (Wang et al., 2020). The ion implantation condition followed that for measured nanoporous samples. As shown by the electron backscatter diffraction (EBSD) patterns and TEM images (Figures 10A and 10B), the 70-nm-thick Si film maintained a high crystal quality under the nominal dose of 2.1×1015 cm−2 and 2.1×1016 cm−2. The minimized structure damage is mostly attributed to the employed low ion implantation voltage and is critical to doping applications (Xia et al., 2014). The Ga concentration is further measured with energy-dispersive X-ray spectroscopy (EDS). Figure 9C shows a typical EDS spectrum taken from a region with a 2.1×1016 cm−2 nominal dose. In addition to Si and Ga atoms, O with an atomic percentage (at%) of 1.9–4.7 at% was found but no C contamination was detected. Detailed Ga concentration for each implantation dose is listed in Table 2. Due to the low acceleration voltage, 45%–50% of the irradiated Ga+ were captured by the thin film instead of penetrating the film. A similarly high Ga at% was observed by Roediger et al. (2011). For ion implantation, it should be noted that more Ga atoms should be found on the top part of a thin film. As a simple way to match the experimental data, the phonon MC simulations are thus focused on an effective A1 value across the whole film thickness.

Figure 10.

Figure 10

Electron microscopy results for the sample with a Ga+ dose of 2.1×1017 cm−2

(A) STEM image of Ga precipitation on the 70-nm-thick Si film (2.1×1017 cm−2 dose) supported by a holey TEM grid. The scale bar is 5 μm. The dark circular regions are the suspended part of the 70-nm-thick Si film on the TEM grid.

(B) EDS mapping of Ga clusters across the Si thin film. The scale bar is 1 μm.

(C) Corresponding secondary electron image for the EDS mapping in (b). The scale bar is 1 μm.

(D) TEM image of one polycrystalline Ga cluster. The scale bar is 50 nm.

(E) TEM image of the interface between crystalline Si and crystalline Ga. The scale bar is 5 nm.

Figure 9.

Figure 9

Electron microscopy results for ion-implanted samples

(A) TEM images of the Si thin film with a nominal Ga+ dose of 2.1×1015 cm−2. Inset is the SAED taken from the region. The scale bar is 5 nm for the TEM image.

(B) TEM images of the Si thin film with a nominal Ga+ dose of 2.1×1016 cm−2. Inset is the SAED taken from the region. The scale bar is 5 nm.

(C) EDS spectrum from the Si thin film with a nominal implanted Ga dose of 2.1×1016 cm−2.

Table 2.

Parameters and results for ion implantation studies. The Ga at% is measured with EDS

Sample index Nominal dose
(cm−2)
Ga at% kL/kL,undoped at 300 K
4o (1000,120,160) 0 0.0 100%
4o+ 2.1×1015 0.3 89.2%
4o++ 2.1×1016 2.7 80.0%

With a high Ga+ dose of 2.1×1017 cm−2, Ga precipitation was observed (Figure 10). Polycrystalline Ga clusters were found across the Si thin film. When the same dose was applied to a suspended nanoporous Si thin film, the suspended part warped due to the significant heating effect from bombarded ions. The kL of the nanoporous thin film cannot be measured for this dose.

Temperature-dependent Cp

Within periodic nanoporous films, the phonon wave effect (Maldovan, 2015) may play an important role in the kL reduction. There are several experimental methods to justify the importance of the wave effect, such as comparing the kL of periodic and aperiodic nanoporous thin films (Lee et al., 2017; Maire et al., 2017; Wagner et al., 2016), comparing the Cp of nanoporous thin films with the bulk value (Hao et al., 2018b), and measuring the thermal resistance of nanoporous thin films with added periodic nanopores (Xiao et al., 2020). To compare Cp values, all measured Cp values are divided by (1Φ) to yield the solid Cp. The porosity Φ is estimated based on the sample geometry but errors may occur due to tapered sidewalls for dry-etched nanopores (Hao et al., 2018a). Considering the thickness variation for each material layer and 10% error in estimated Φ, all solid Cp values are within the range for the predictions using bulk specific heat values (Figure 11). This result suggests negligible influence of the wave effect for the measured temperature range. It should also be noted that the ion implantation did not significantly affect the Cp value. In practice, doping heavy atoms to the material (i.e., increased mean atomic mass) has been widely used to reduce the k by decreasing the phonon group velocity (Kargar et al., 2018; Pei et al., 2012; Winter and Clarke, 2007). Based on measured Cp values, the phonon dispersion modification cannot be remarkable with implanted Ga+ ions here.

Figure 11.

Figure 11

Measured solid Cp for nanoporous thin films

The predicted range is based on the bulk material properties of the metal-coated film. Prediction band 1 considers the uncertainty of the film thickness (± 5 nm for Si, ± 2 nm for Cr, and ± 4 nm for Pt). Prediction band 2 further considers the uncertainty of porosity measurements (± 10%) in addition to the thickness uncertainty.

Conclusions

In this work, nanoslot alignment and ion implantation are explored for their potential applications in reducing the kL of Si thin films. Without changing the feature size and porosity, a 33.7% kL reduction at 125 K can be achieved simply by offsetting nanoslots. This nanoporous structure can be easily implemented in thin films and atomic-thick materials to tune the thermal transport. For aligned periodic nanoslots, the characteristic length can be approximated as Lc3wl/4Hw(pw), where Hw is the correction factor to account for the reduced thermal conductance due to decreased cross-section for heat conduction inside a porous structure (Hao and Xiao, 2020). For thermoelectrics, the recommended Lc20 nm for heavily doped Si (Qiu et al., 2015) can be achieved by patterning aligned nanoslots with l = 60 nm, b = 10 nm, w = 13 nm, and p = 120 nm. The dimensions here are much larger than the aforementioned pore diameter of 6.77 nm for periodic circular pores with a 25% porosity. Following this, offset nanoslots can provide a comparable Lc value with even larger feature sizes and thus minimize the challenges in nanofabrication. The alignment modification can also be applied to three-dimensional nanoporous structures (Ma et al., 2016; Yang et al., 2014) when classical phonon size effects are dominant. Along another line, the low-energy ion implantation introduced with an FIB did not damage the overall lattice structure but created local defects to suppress the phonon transport, which can be combined with nanoporous patterns to minimize the kL but still maintain bulk-like electrical properties. Attention should also be paid to charge carriers trapped by pore-edge defects. The charge carrier depletion near pore edges may largely reduce the carrier concentration (Tang et al., 2010). A potential field can build up to further scatter electrons and holes within the pore-edge depletion region (Hao and Xiao, 2019; Hao et al., 2017b), which also affects the electrical properties.

Limitations of the study

This study focuses on suppressing the phonon transport inside Si thin films. In principle, the method (i.e., well-designed nanoporous patterns and ion implantation) could be extended to other thin films and atomic-thick materials to locally tune the thermal properties. For thermoelectric applications, the electrical properties of the material should be experimentally measured for ion implantation. More complicated staggered patterns should also be investigated.

STAR★Methods

Key resources table

REAGENT or RESOURCE SOURCE IDENTIFIER
Software and algorithms

Fortran Intel Corporation https://www.fortran90.org/
MATLAB MathWorks, Inc. https://www.mathworks.com/
SigmaPlot SPSS Inc. https://systatsoftware.com/sigmaplot/

Resource availability

Lead contact

Further information and requests for resources including data and code should be directed to and will be fulfilled by the lead contact, Dr. Qing Hao (qinghao@arizona.edu).

Materials availability

This study did not generate new unique materials.

Data and code availability

Any additional information required to re-analyze the data reported in this paper is available from the lead contact upon request.

The phonon Monte Carlo simulation code is unavailable for now.

Method details

Device fabrication

All devices were fabricated from the 70-nm-thick device layer of a silicon-on-insulator (SOI) wafer. Nanoslots of various sizes and pitches were defined with EBL (Elionix ELS-7000). Reactive ion etching (RIE) with an SF6/Ar recipe was used to remove the Si layer and thus produce the nanoslot-patterned Si beam. The sample was then wet etched in the diluted hydrofluoric acid (DHF) solution to remove the underlying 2-μm-thick buried oxide (BOX) layer and fully suspend the nanoporous Si thin film. All thermal measurements were based on the AC self-heating of a suspended device, i.e., the 3ω method (Hao et al., 2018b, 2020, 2018b; Xiao et al., 2019). As the heater and electrical resistance thermometer, an electrically conductive layer consisting of 10 nm Cr and then 40 nm Pt was deposited onto the Si device via electron beam evaporation. Before thermal measurements, each sample was carefully examined under a scanning electron microscope (SEM) to ensure the quality of the suspended bridge and measure the exact dimension of nanoslot patterns. The suspended region is 20 × 2 μm in the top view (Figure 2). SEM images for different device configurations can be found in the supplemental information (Figure S1). It is acknowledged that pore-edge defects (e.g., roughness, oxidation, amorphization) introduced during the nanofabrication process could affect the phonon transport inside the nanoporous structure (Hao et al., 2018a), which can effectively expand the pore size (Ravichandran and Minnich, 2014). For the feature sizes of our fabricated nanoslot patterns, the ∼2 nm surface layer introduced by dry etching (Tang et al., 2010) can be neglected for its impact and the comparison between aligned and offset patterns is unaffected when both patterns are fabricated under the same etching condition.

Ion implantation with an FIB

To locally introduce defects on a nanoslot-patterned Si thin film, ion implantation was performed within a Ga FIB (FEI Helios NanoLab 660) before the metal deposition. The dose was calculated by multiplying the beam current, exposure area and exposure time. Since the ion beam was bombarded onto a small area, rapid ion implantation with high doses can be achieved. During the exposure, the acceleration voltage for the Ga+ ions was set to 500 eV to minimize the milling effect. The sputter yield, which is the ratio between the number of removed atoms and irradiated ions, is around 0.7 for Si under 500 eV ion bombardment (Rossnagel, 2001). Based on this value, the thickness of the Si removed during the exposure is estimated to be 0.14 nm for the largest dose, which is negligible compared with the 70 nm film thickness.

Thermal characterization

The in-plane thermal conductivity measurement was carried out in a high vacuum (<1 × 10−4 Torr) with the 3ω method (Cahill, 1990; Lu et al., 2001). The measurement temperature range was from 83 to 300 K for all samples. An AC source (Keithley 6221) was used to generate a heating current at an angular frequency ω. Using differential circuits to eliminate the 1ω signal, the third-harmonic root-mean-square (RMS) voltage (V3ω) of the sample and the corresponding phase lag were recorded by a lock-in amplifier (Stanford Research SR830).

In the data analysis, the effective in-plane thermal conductivity k and specific heat of the metal-coated film can be extracted by fitting V3ω as a function of the angular frequency, given as (Lu et al., 2001)

V3ω4I3LRR/(π4kS1+(2ωγ)2) (Equation 2)

In Equation 2, I is the RMS heating current, R is the electrical resistance, R=dR/dT, T is the absolute temperature, L is the length and S is the cross-sectional area of the suspended film. The volumetric specific heat (Cp) can be further calculated from the thermal time constant γ=CpL2/π2k for the metal-coated film. The linear regression is performed by transforming Equation 2 into the following form:

1V3ω2=[2π4Skγ4I3LR(dR/dT)]2ω2+[π4Sk4I3LR(dR/dT)]2 (Equation 3)

Figure 3 shows the analysis result of Sample 1a at 300 K, with R2 > 0.9995 for the linear fitting.

To obtain the in-plane thermal conductivity of the Si film itself, the k contribution from the metal layer should be subtracted. The thermal conductivity of the metal layer can be estimated from its electrical conductivity σ using the Wiedemann-Franz law, kE=LσT. The employed temperature-dependent Lorenz number L has been calibrated in an earlier study on an identical metal layer (Hao et al., 2018b). For the specific heat of the metal layer, bulk values for each material are used in the calculation.

Phonon MC simulation

In phonon MC simulations, the movement and scattering of individual phonons are tracked to yield a statistical description of the transport process, as an alternative way to solve the phonon BTE. By employing the periodic heat flux boundary condition (Hao et al., 2009; Hao and Xiao, 2020), a single period can be selected as the computational domain to minimize the computational load. The computational efficiency can be further improved using a variance-reduced MC technique developed by Péraud and Hadjiconstantinou, where the tracked “useful” phonons are related to the distribution function distortion from an equilibrium distribution function, i.e., the Bose-Einstein distribution at a reference temperature (Péraud and Hadjiconstantinou, 2011). These tracked phonons are directly associated with the net heat flow across a structure. Phonon boundary scattering and internal scattering are two scattering events in the simulation. The former includes scattering with the nanoslot edge and top/bottom film surfaces, which are both treated as diffusive. Considering impurity scattering and Umklapp (U) scattering, the internal phonon scattering rate is expressed as 1/τ(ω)=A1ω4+B1ω2Texp(B2/T). In the expression, τ(ω) is the phonon relaxation time depending on the phonon angular frequency ω, A1 is the parameter for the impurity scattering, B1 and B2 are the parameters for the U processes. The employed parameters were obtained by fitting the temperature-dependent kL of bulk Si (Hao, 2014; Wang et al., 2011). Following these works, A1=1.69×1045s3, B1=1.53×1019s/K and B2=140K were used in the simulation of undoped samples. For samples with point defects introduced by the ion implantation, the A1 value should be further increased to account for stronger point-defect phonon scattering (Hao et al., 2010).

Acknowledgments

The authors thank the support from National Science Foundation CAREER Award (grant number CBET-1651840). This work was performed in part at the OSC Micro/Nano Fabrication Cleanroom and Nano Fabrication Center at the University of Arizona. SEM, FIB, and TEM analyses were performed at the Kuiper Materials Imaging and Characterization facility at the University of Arizona. Wang thanks Dr. Jerry Chang for his help with the TEM. An allocation of computer time from the UA Research Computing High Performance Computing (HPC) and High Throughput Computing (HTC) at the University of Arizona is gratefully acknowledged.

Author contributions

Qing Hao supervised the work. Sien Wang performed the experiments. Yue Xiao and Sien Wang performed the simulations. Sien Wang, Yue Xiao, and Qiyu Chen performed the data analysis. Sien Wang, Yue Xiao, and Qing Hao wrote the article.

Declaration of interests

The authors declare no competing interests.

Published: November 18, 2022

Footnotes

Supplemental information can be found online at https://doi.org/10.1016/j.isci.2022.105386.

Supplemental information

Document S1. Figure S1
mmc1.pdf (105.5KB, pdf)

References

  1. Alaie S., Baboly M.G., Jiang Y.B., Rempe S., Anjum D.H., Chaieb S., Donovan B.F., Giri A., Szwejkowski C.J., Gaskins J.T., et al. Reduction and increase in thermal conductivity of Si irradiated with Ga+ via focused ion beam. ACS Appl. Mater. Interfaces. 2018;10:37679–37684. doi: 10.1021/acsami.8b11949. [DOI] [PubMed] [Google Scholar]
  2. Anufriev R., Ramiere A., Maire J., Nomura M. Heat guiding and focusing using ballistic phonon transport in phononic nanostructures. Nat. Commun. 2017;8:15505. doi: 10.1038/ncomms15505. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Anufriev R., Gluchko S., Volz S., Nomura M. Probing ballistic thermal conduction in segmented silicon nanowires. Nanoscale. 2019;11:13407–13414. doi: 10.1039/C9NR03863A. [DOI] [PubMed] [Google Scholar]
  4. Anufriev R., Ordonez-Miranda J., Nomura M. Measurement of the phonon mean free path spectrum in silicon membranes at different temperatures using arrays of nanoslits. Phys. Rev. B. 2020;101:115301. doi: 10.1103/PhysRevB.101.115301. [DOI] [Google Scholar]
  5. Biswas K., He J., Blum I.D., Wu C.-I., Hogan T.P., Seidman D.N., Dravid V.P., Kanatzidis M.G. High-performance bulk thermoelectrics with all-scale hierarchical architectures. Nature. 2012;489:414–418. doi: 10.1038/nature11439. [DOI] [PubMed] [Google Scholar]
  6. Bux S.K., Blair R.G., Gogna P.K., Lee H., Chen G., Dresselhaus M.S., Kaner R.B., Fleurial J.-P. Nanostructured bulk silicon as an effective thermoelectric material. Adv. Funct. Mater. 2009;19:2445–2452. doi: 10.1002/adfm.200900250. [DOI] [Google Scholar]
  7. Cahill D.G. Thermal conductivity measurement from 30 to 750 K: the 3ω method. Rev. Sci. Instrum. 1990;61:802–808. doi: 10.1063/1.1141498. [DOI] [Google Scholar]
  8. Chen G. Oxford university press; 2005. Nanoscale Energy Transport and Conversion: A Parallel Treatment of Electrons, Molecules, Phonons, and Photons. [Google Scholar]
  9. Esfarjani K., Chen G., Stokes H.T. Heat transport in silicon from first-principles calculations. Phys. Rev. B. 2011;84:085204. [Google Scholar]
  10. Gluchko S., Anufriev R., Yanagisawa R., Volz S., Nomura M. On the reduction and rectification of thermal conduction using phononic crystals with pacman-shaped holes. Appl. Phys. Lett. 2019;114:023102. doi: 10.1063/1.5079931. [DOI] [Google Scholar]
  11. Graczykowski B., El Sachat A., Reparaz J.S., Sledzinska M., Wagner M.R., Chávez-Ángel E., Wu Y., Volz S., Wu Y., Alzina F., Sotomayor Torres C.M. Thermal conductivity and air-mediated losses in periodic porous silicon membranes at high temperatures. Nat. Commun. 2017;8:415. doi: 10.1038/s41467-017-00115-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Gunst T., Markussen T., Jauho A.-P., Brandbyge M. Thermoelectric properties of finite graphene antidot lattices. Phys. Rev. B. 2011;84:155449. [Google Scholar]
  13. Hao Q. General effective medium formulation for thermal analysis of a polycrystal—the influence of partially specular phonon transmission across grain boundaries. J. Appl. Phys. 2014;116:034305. doi: 10.1063/1.4890362. [DOI] [Google Scholar]
  14. Hao Q., Garg J. A review on phonon transport within polycrystalline materials. ES Mater. Manuf. 2021;14:36–50. doi: 10.30919/esmm5f480. [DOI] [Google Scholar]
  15. Hao Q., Xiao Y. Electron Monte Carlo simulations of nanoporous Si thin films—the influence of pore-edge charges. J. Appl. Phys. 2019;125:064301. [Google Scholar]
  16. Hao Q., Xiao Y. Periodic nanoslot patterns as an effective approach to improving the thermoelectric performance of thin films. Phys. Rev. Appl. 2020;13:064020. doi: 10.1103/PhysRevApplied.13.064020. [DOI] [Google Scholar]
  17. Hao Q., Chen G., Jeng M.-S. Frequency-dependent Monte Carlo simulations of phonon transport in two-dimensional porous silicon with aligned pores. J. Appl. Phys. 2009;106:114321. doi: 10.1063/1.3266169. [DOI] [Google Scholar]
  18. Hao Q., Zhu G., Joshi G., Wang X., Minnich A., Ren Z., Chen G. Theoretical studies on the thermoelectric figure of merit of nanograined bulk silicon. Appl. Phys. Lett. 2010;97:063109. [Google Scholar]
  19. Hao Q., Xiao Y., Zhao H. Characteristic length of phonon transport within periodic nanoporous thin films and two-dimensional materials. J. Appl. Phys. 2016;120:065101. doi: 10.1063/1.4959984. [DOI] [Google Scholar]
  20. Hao Q., Xu D., Lu N., Zhao H. High-throughput ZT predictions of nanoporous bulk materials as next-generation thermoelectric materials: a material genome approach. Phys. Rev. B. 2016;93:205206. doi: 10.1103/PhysRevB.93.205206. [DOI] [Google Scholar]
  21. Hao Q., Xiao Y., Zhao H. Analytical model for phonon transport analysis of periodic bulk nanoporous structures. Appl. Therm. Eng. 2017;111:1409–1416. doi: 10.1016/j.applthermaleng.2016.06.075. [DOI] [Google Scholar]
  22. Hao Q., Zhao H., Xu D. Thermoelectric studies of nanoporous thin films with adjusted pore-edge charges. J. Appl. Phys. 2017;121:094308. [Google Scholar]
  23. Hao Q., Xu D., Zhao H., Xiao Y., Medina F.J. Thermal studies of nanoporous Si films with pitches on the order of 100 nm—comparison between different pore-drilling techniques. Sci. Rep. 2018;8:9056. doi: 10.1038/s41598-018-26872-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Hao Q., Xu D., Zhao H., Xiao Y., Medina F.J. Thermal studies of nanoporous Si films with pitches on the order of 100 nm —comparison between different pore-drilling techniques. Sci. Rep. 2018;8:9056. doi: 10.1038/s41598-018-26872-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Hao Q., Zhao H., Xiao Y., Kronenfeld M.B. Electrothermal studies of GaN-based high electron mobility transistors with improved thermal designs. Int. J. Heat Mass Tran. 2018;116:496–506. doi: 10.1016/j.ijheatmasstransfer.2017.09.048. [DOI] [Google Scholar]
  26. Hao Q., Xiao Y., Chen Q. Determining phonon mean free path spectrum by ballistic phonon resistance within a nanoslot-patterned thin film. Materials Today Physics. 2019;10:100126. doi: 10.1016/j.mtphys.2019.100126. [DOI] [Google Scholar]
  27. Hao Q., Xiao Y., Wang S. Two-step modification of phonon mean free paths for thermal conductivity predictions of thin-film-based nanostructures. Int. J. Heat Mass Tran. 2020;153:119636. doi: 10.1016/j.ijheatmasstransfer.2020.119636. [DOI] [Google Scholar]
  28. Heron J.-S., Bera C., Fournier T., Mingo N., Bourgeois O. Blocking phonons via nanoscale geometrical design. Phys. Rev. B. 2010;82:155458. doi: 10.1103/PhysRevB.82.155458. [DOI] [Google Scholar]
  29. Huang C.-L., Lin Z.-Z., Feng Y.-H., Zhang X.-X., Wang G. Thermal conductivity prediction of 2- dimensional square-pore metallic nanoporous materials with kinetic method approach. Int. J. Therm. Sci. 2017;112:263–269. doi: 10.1016/j.ijthermalsci.2016.09.033. [DOI] [Google Scholar]
  30. Jain A., Yu Y.-J., McGaughey A.J.H. Phonon transport in periodic silicon nanoporous films with feature sizes greater than 100 nm. Phys. Rev. B. 2013;87:195301. doi: 10.1103/PhysRevB.87.195301. [DOI] [Google Scholar]
  31. Kargar F., Penilla E.H., Aytan E., Lewis J.S., Garay J.E., Balandin A.A. Acoustic phonon spectrum engineering in bulk crystals via incorporation of dopant atoms. Appl. Phys. Lett. 2018;112:191902. doi: 10.1063/1.5030558. [DOI] [Google Scholar]
  32. Kashiwagi M., Liao Y., Ju S., Miura A., Konishi S., Shiga T., Kodama T., Shiomi J. Scalable multi-nanostructured silicon for room-temperature thermoelectrics. ACS Appl. Energy Mater. 2019;2:7083–7091. doi: 10.1021/acsaem.9b00893. [DOI] [Google Scholar]
  33. Kraemer D., Poudel B., Feng H.-P., Caylor J.C., Yu B., Yan X., Ma Y., Wang X., Wang D., Muto A., et al. High-performance flat-panel solar thermoelectric generators with high thermal concentration. Nat. Mater. 2011;10:532–538. doi: 10.1038/nmat3013. http://www.nature.com/nmat/journal/v10/n7/abs/nmat3013.html#supplementary-information [DOI] [PubMed] [Google Scholar]
  34. Lan Y., Poudel B., Ma Y., Wang D., Dresselhaus M.S., Chen G., Ren Z. Structure study of bulk nanograined thermoelectric bismuth antimony telluride. Nano Lett. 2009;9:1419–1422. doi: 10.1021/nl803235n. [DOI] [PubMed] [Google Scholar]
  35. Lee J., Lee W., Wehmeyer G., Dhuey S., Olynick D.L., Cabrini S., Dames C., Urban J.J., Yang P. Investigation of phonon coherence and backscattering using silicon nanomeshes. Nat. Commun. 2017;8:14054. doi: 10.1038/ncomms14054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Lim J., Wang H.-T., Tang J., Andrews S.C., So H., Lee J., Lee D.H., Russell T.P., Yang P. Simultaneous thermoelectric property measurement and incoherent phonon transport in holey silicon. ACS Nano. 2016;10:124–132. doi: 10.1021/acsnano.5b05385. [DOI] [PubMed] [Google Scholar]
  37. Lu L., Yi W., Zhang D.L. 3ω method for specific heat and thermal conductivity measurements. Rev. Sci. Instrum. 2001;72:2996–3003. doi: 10.1063/1.1378340. [DOI] [Google Scholar]
  38. Ma D., Ding H., Meng H., Feng L., Wu Y., Shiomi J., Yang N. Nano-cross-junction effect on phonon transport in silicon nanowire cages. Phys. Rev. B. 2016;94:165434. [Google Scholar]
  39. Maire J., Anufriev R., Yanagisawa R., Ramiere A., Volz S., Nomura M. Heat conduction tuning by wave nature of phonons. Sci. Adv. 2017;3:e1700027. doi: 10.1126/sciadv.1700027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Maire J., Anufriev R., Hori T., Shiomi J., Volz S., Nomura M. Thermal conductivity reduction in silicon fishbone nanowires. Sci. Rep. 2018;8:4452. doi: 10.1038/s41598-018-22509-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Maldovan M. Phonon wave interference and thermal bandgap materials. Nat. Mater. 2015;14:667–674. doi: 10.1038/nmat4308. [DOI] [PubMed] [Google Scholar]
  42. Malhotra A., Maldovan M. Surface scattering controlled heat conduction in semiconductor thin films. J. Appl. Phys. 2016;120:204305. doi: 10.1063/1.4968542. [DOI] [Google Scholar]
  43. Marconnet A.M., Kodama T., Asheghi M., Goodson K.E. Phonon conduction in periodically porous silicon nanobridges. Nanoscale Microscale Thermophys. Eng. 2012;16:199–219. [Google Scholar]
  44. Marconnet A.M., Asheghi M., Goodson K.E. From the casimir limit to phononic crystals: 20 Years of phonon transport studies using silicon-on-insulator technology. J. Heat Transfer. 2013;135 doi: 10.1115/1.4023577. [DOI] [Google Scholar]
  45. Nomura M., Anufriev R., Zhang Z., Maire J., Guo Y., Yanagisawa R., Volz S. Review of thermal transport in phononic crystals. Materials Today Physics. 2022;22:100613. doi: 10.1016/j.mtphys.2022.100613. [DOI] [Google Scholar]
  46. Oh J., Yoo H., Choi J., Kim J.Y., Lee D.S., Kim M.J., Lee J.-C., Kim W.N., Grossman J.C., Park J.H., et al. Significantly reduced thermal conductivity and enhanced thermoelectric properties of single- and bi-layer graphene nanomeshes with sub-10nm neck-width. Nano Energy. 2017;35:26–35. doi: 10.1016/j.nanoen.2017.03.019. [DOI] [Google Scholar]
  47. Park W., Sohn J., Romano G., Kodama T., Sood A., Katz J.S., Kim B.S.Y., So H., Ahn E.C., Asheghi M., et al. Impact of thermally dead volume on phonon conduction along silicon nanoladders. Nanoscale. 2018;10:11117–11122. doi: 10.1039/C8NR01788C. [DOI] [PubMed] [Google Scholar]
  48. Pei Y., Wang H., Snyder G.J. Band engineering of thermoelectric materials. Adv. Mater. 2012;24:6125–6135. doi: 10.1002/adma.201202919. [DOI] [PubMed] [Google Scholar]
  49. Péraud J.P.M., Hadjiconstantinou N.G. Efficient simulation of multidimensional phonon transport using energy-based variance-reduced Monte Carlo formulations. Phys. Rev. B. 2011;84:205331. doi: 10.1103/PhysRevB.84.205331. [DOI] [Google Scholar]
  50. Poudel B., Hao Q., Ma Y., Lan Y., Minnich A., Yu B., Yan X., Wang D., Muto A., Vashaee D., et al. High-thermoelectric performance of nanostructured bismuth antimony telluride bulk alloys. Science. 2008;320:634–638. doi: 10.1126/science.1156446. [DOI] [PubMed] [Google Scholar]
  51. Prasher R. Predicting the thermal resistance of nanosized constrictions. Nano Lett. 2005;5:2155–2159. doi: 10.1021/nl051710b. [DOI] [PubMed] [Google Scholar]
  52. Qiu B., Tian Z., Vallabhaneni A., Liao B., Mendoza J.M., Restrepo O.D., Ruan X., Chen G. First-principles simulation of electron mean-free-path spectra and thermoelectric properties in silicon. EPL. 2015;109:57006. doi: 10.1209/0295-5075/109/57006. [DOI] [Google Scholar]
  53. Ravichandran N.K., Minnich A.J. Coherent and incoherent thermal transport in nanomeshes. Phys. Rev. B. 2014;89:205432. [Google Scholar]
  54. Roediger P., Wanzenboeck H.D., Waid S., Hochleitner G., Bertagnolli E. Focused-ion-beam-inflicted surface amorphization and gallium implantation—new insights and removal by focused-electron-beam-induced etching. Nanotechnology. 2011;22:235302. doi: 10.1088/0957-4484/22/23/235302. [DOI] [PubMed] [Google Scholar]
  55. Romano G., Grossman J.C. Toward phonon-boundary engineering in nanoporous materials. Appl. Phys. Lett. 2014;105:033116. doi: 10.1063/1.4891362. [DOI] [Google Scholar]
  56. Romano G., Kolpak A.M. Directional phonon suppression function as a tool for the identification of ultralow thermal conductivity materials. Sci. Rep. 2017;7:44379. doi: 10.1038/srep44379. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Rossnagel S. In: Handbook of Thin Film Deposition Processes and Techniques. Second Edition. Seshan K., editor. William Andrew Publishing); 2001. 8 - sputtering and sputter deposition; pp. 319–348. [DOI] [Google Scholar]
  58. Scott E.A., Hattar K., Braun J.L., Rost C.M., Gaskins J.T., Bai T., Wang Y., Ganski C., Goorsky M., Hopkins P.E. Orders of magnitude reduction in the thermal conductivity of polycrystalline diamond through carbon, nitrogen, and oxygen ion implantation. Carbon. 2020;157:97–105. doi: 10.1016/j.carbon.2019.09.076. [DOI] [Google Scholar]
  59. Shi X.-L., Zou J., Chen Z.-G. Advanced thermoelectric design: from materials and structures to devices. Chem. Rev. 2020;120:7399–7515. doi: 10.1021/acs.chemrev.0c00026. [DOI] [PubMed] [Google Scholar]
  60. Slack G. In: CRC Handbook of Thermoelectrics. Rowe D.M., editor. CRC Press; 1995. New materials and performance limits for thermoelectric cooling; pp. 407–440. [Google Scholar]
  61. Song D., Chen G. Thermal conductivity of periodic microporous silicon films. Appl. Phys. Lett. 2004;84:687–689. [Google Scholar]
  62. Tamura M., Shukuri S., Moniwa M., Default M. Focused ion beam gallium implantation into silicon. Appl. Phys. A. 1986;39:183–190. [Google Scholar]
  63. Tang G.H., Bi C., Fu B. Thermal conduction in nano-porous silicon thin film. J. Appl. Phys. 2013;114:184302. doi: 10.1063/1.4829913. [DOI] [Google Scholar]
  64. Tang J., Wang H.-T., Lee D.H., Fardy M., Huo Z., Russell T.P., Yang P. Holey silicon as an efficient thermoelectric material. Nano Lett. 2010;10:4279–4283. doi: 10.1021/nl102931z. [DOI] [PubMed] [Google Scholar]
  65. Verdier M., Anufriev R., Ramiere A., Termentzidis K., Lacroix D. Thermal conductivity of phononic membranes with aligned and staggered lattices of holes at room and low temperatures. Phys. Rev. B. 2017;95:205438. doi: 10.1103/PhysRevB.95.205438. [DOI] [Google Scholar]
  66. Vineis C.J., Shakouri A., Majumdar A., Kanatzidis M.G. Nanostructured thermoelectrics: big efficiency gains from small features. Adv. Mater. 2010;22:3970–3980. doi: 10.1002/adma.201000839. [DOI] [PubMed] [Google Scholar]
  67. Wagner M.R., Graczykowski B., Reparaz J.S., El Sachat A., Sledzinska M., Alzina F., Sotomayor Torres C.M. Two-dimensional phononic crystals: disorder matters. Nano Lett. 2016;16:5661–5668. doi: 10.1021/acs.nanolett.6b02305. [DOI] [PubMed] [Google Scholar]
  68. Wan C., Wang Y., Wang N., Norimatsu W., Kusunoki M., Koumoto K. Development of novel thermoelectric materials by reduction of lattice thermal conductivity. Sci. Technol. Adv. Mater. 2010;11:044306. doi: 10.1088/1468-6996/11/4/044306. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Wang S., Xu D., Gurunathan R., Snyder G.J., Hao Q. Thermal studies of individual Si/Ge heterojunctions — the influence of the alloy layer on the heterojunction. Journal of Materiomics. 2020;6:248–255. doi: 10.1016/j.jmat.2020.02.013. [DOI] [Google Scholar]
  70. Wang Z., Alaniz J.E., Jang W., Garay J.E., Dames C. Thermal conductivity of nanocrystalline silicon: importance of grain size and frequency-dependent mean free paths. Nano Lett. 2011;11:2206–2213. doi: 10.1021/nl1045395. [DOI] [PubMed] [Google Scholar]
  71. Winter M.R., Clarke D.R. Oxide materials with low thermal conductivity. J. Am. Ceram. Soc. 2007;90:533–540. doi: 10.1111/j.1551-2916.2006.01410.x. [DOI] [Google Scholar]
  72. Winter N., Becton M., Zhang L., Wang X. Effects of pore design on mechanical properties of nanoporous silicon. Acta Mater. 2017;124:127–136. doi: 10.1016/j.actamat.2016.11.006. [DOI] [Google Scholar]
  73. Won Y., Cho J., Agonafer D., Asheghi M., Goodson K.E. IEEE); 2013. Cooling Limits for GaN HEMT Technology; pp. 1–5. [Google Scholar]
  74. Wood C. Materials for thermoelectric energy conversion. Rep. Prog. Phys. 1988;51:459–539. [Google Scholar]
  75. Xia M., Cheng Z., Han J., Zheng M., Sow C.-H., Thong J.T.L., Zhang S., Li B. Gallium ion implantation greatly reduces thermal conductivity and enhances electronic one of ZnO nanowires. AIP Adv. 2014;4:057128. doi: 10.1063/1.4880240. [DOI] [Google Scholar]
  76. Xiao Y., Chen Q., Ma D., Yang N., Hao Q. Phonon transport within periodic porous structures - from classical phonon size effects to wave effects. ES Mater. Manuf. 2019;5:2–18. doi: 10.30919/esmm5f237. [DOI] [Google Scholar]
  77. Xiao Y., Xu D., Medina F.J., Wang S., Hao Q. Thermal studies of nanoporous thin films with added periodic nanopores—a new approach to evaluate the importance of phononic effects. Materials Today Physics. 2020;12:100179. doi: 10.1016/j.mtphys.2020.100179. [DOI] [Google Scholar]
  78. Xiao Y., Chen Q., Hao Q. Inverse thermal design of nanoporous thin films for thermal cloaking. Materials Today Physics. 2021;21:100477. doi: 10.1016/j.mtphys.2021.100477. [DOI] [Google Scholar]
  79. Xiao Y., Hao Q. Nanoslot patterns for enhanced thermal anisotropy of Si thin films. Int. J. Heat Mass Tran. 2021;170:120944. doi: 10.1016/j.ijheatmasstransfer.2021.120944. [DOI] [Google Scholar]
  80. Xu D., Tang S., Du X., Hao Q. Detecting the major charge-carrier scattering mechanism in graphene antidot lattices. Carbon. 2019;144:601–607. doi: 10.1016/j.carbon.2018.12.080. [DOI] [Google Scholar]
  81. Yang J., Stabler F.R. Automotive applications of thermoelectric materials. J. Electron. Mater. 2009;38:1245–1251. doi: 10.1007/s11664-009-0680-z. [DOI] [Google Scholar]
  82. Yang L., Yang N., Li B. Extreme low thermal conductivity in nanoscale 3D Si phononic crystal with spherical pores. Nano Lett. 2014;14:1734–1738. doi: 10.1021/nl403750s. [DOI] [PubMed] [Google Scholar]
  83. Yang L., Zhang Q., Wei Z., Cui Z., Zhao Y., Xu T.T., Yang J., Li D. Kink as a new degree of freedom to tune the thermal conductivity of Si nanoribbons. J. Appl. Phys. 2019;126:155103. doi: 10.1063/1.5119727. [DOI] [Google Scholar]
  84. Yang L., Zhao Y., Zhang Q., Yang J., Li D. Thermal transport through fishbone silicon nanoribbons: unraveling the role of Sharvin resistance. Nanoscale. 2019;11:8196–8203. doi: 10.1039/C9NR01855G. [DOI] [PubMed] [Google Scholar]
  85. Yu J.-K., Mitrovic S., Tham D., Varghese J., Heath J.R. Reduction of thermal conductivity in phononic nanomesh structures. Nat. Nanotechnol. 2010;5:718–721. doi: 10.1038/nnano.2010.149. [DOI] [PubMed] [Google Scholar]
  86. Zhao Y., Liu D., Chen J., Zhu L., Belianinov A., Ovchinnikova O.S., Unocic R.R., Burch M.J., Kim S., Hao H., et al. Engineering the thermal conductivity along an individual silicon nanowire by selective helium ion irradiation. Nat. Commun. 2017;8:15919. doi: 10.1038/ncomms15919. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Zhao Y., Yang L., Liu C., Zhang Q., Chen Y., Yang J., Li D. Kink effects on thermal transport in silicon nanowires. Int. J. Heat Mass Tran. 2019;137:573–578. doi: 10.1016/j.ijheatmasstransfer.2019.03.104. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Document S1. Figure S1
mmc1.pdf (105.5KB, pdf)

Data Availability Statement

Any additional information required to re-analyze the data reported in this paper is available from the lead contact upon request.

The phonon Monte Carlo simulation code is unavailable for now.


Articles from iScience are provided here courtesy of Elsevier

RESOURCES