High thermal conductivity materials are useful for thermal management applications and fundamental studies of phonon transport. Past measurements of several ultrahigh thermal conductivity materials were not able to obtain the peak thermal conductivity, which is expected to appear at a low temperature and contains insight into the competition between extrinsic phonon-defect and phonon-boundary scattering with intrinsic phonon–phonon processes. Here, we report a peak thermal conductivity measurement method based on differential Wheatstone bridge measurement of the small temperature drop between two thin film resistance thermometers patterned directly on the sample. With the use of a mesoscale silicon bar sample as the calibration standard, this method is able to obtain results that agree with past measurements of large bulk silicon crystals at high temperatures and first-principles calculation results that account for additional phonon-boundary scattering in the sample. The agreement demonstrates the accuracy of this measurement method for peak thermal conductivity measurements of high thermal conductivity materials.

Thermal management has become a major challenge for microelectronic devices because the high-density electrical power is dissipated as Joule heating.1–3 High operating temperatures at transistor hot spots can lead to performance degradation and thermomechanical failures in devices. Advanced thermal management approaches play an essential role in future progress of microelectronics. As an effort to find potential solutions to the mounting thermal management challenges, ultrahigh thermal conductivity materials, including diamond,4,5 graphitic materials,6–8 and semiconducting zinc blende boron arsenide (BAs),9 have been studied for heterogeneous integration with microelectronic devices.

Accurate thermal transport measurements are essential in the study of ultrahigh thermal conductivity materials. Several steady-state and transient measurement methods have been established for measuring the thermal conductivity of bulk crystals, thin films, and nanostructures. The standard guarded hot plate method10 is suitable for samples with a large cross section. The accuracy of this measurement method has been improved using a guard and numerical modeling to minimize edge heat losses and precisely control the temperature profile. For a high thermal conductivity sample, however, the thermal resistance across the thickness of the plate sample can be small compared to the contact thermal resistance between the thermometer and each end of the plate. In addition, it is a significant challenge to grow ultrahigh thermal conductivity single crystals with a sufficiently large size for guarded hot plate measurements.

Small-size, high thermal conductivity bulk crystals can be prepared into a rod sample for steady-state thermal conductivity measurement. In this approach,11,12 one end of the sample rod is attached to a heater and a heat flux sensor made of a reference rod material with known thermal conductivity, while the other end is attached to a heat sink. The temperature drops along the heat flux sensor and the sample rod are measured with the use of either differential thermocouples or resistance thermometers attached to the sample with either thermal grease or pressure contact. Due to the presence of heat loss into these surface-mounted thermometers with electrical lead wires and the contact thermal resistance between the thermometers and sample, the thermometer temperature is different from the sample temperature. This difference can become appreciable compared to the small temperature drop along a high thermal conductivity crystal rod sample. The resultant measurement error is especially important at a low temperature where the peak thermal conductivity appears as a result of the interplay between intrinsic phonon–phonon scattering processes and extrinsic phonon-defect and phonon-boundary scattering processes.13 Although the peak thermal conductivity contains valuable insight into the competition between these intrinsic and extrinsic phonon scattering mechanisms, a prior steady-state measurement of small BAs rod samples was not able to obtain the peak thermal conductivity because of the uncertainty due to this measurement error as well as the decreased Seebeck coefficient of the differential thermocouples at low temperatures.14 

Among the transient thermal transport measurement methods, laser flash measurement apparatus is commercially available for thermal diffusivity measurement of relatively large disk-shaped samples. In this apparatus, transient optical heating is applied to one side of the sample disk, while the temperature evolution of the other side is monitored using an optical non-contact measurement technique, such as infrared thermometry. It is often necessary to coat the two sample surfaces with a black graphite spray to confine the optical heating and temperature measurement on the sample surfaces. Although the transient measurement method is relatively insensitive to contact thermal resistance compared to the steady-state guarded hot plate measurement of a high thermal conductivity sample, the contact thermal resistance between the graphite spray and the sample can still become a source of error in laser flash measurement of a high thermal conductivity thin sample. In addition, the specific heat of the sample needs to be measured separately and used with the measured thermal diffusivity to obtain the thermal conductivity. Accurate measurement of the specific heat of small-size samples is a challenge itself at both high and low temperatures.

The time-domain thermal reflectance (TDTR) measurement technique has allowed for the extraction of both the sample effusivity and the thermal interface conductance of a thin film metal transducer coated on a bulk or thin film sample.15,16 The specific heat of the sample needs to be determined from separate measurements or a theory for the conversion of the effusivity into the thermal conductivity. Because the thermal penetration depth of the ultrafast laser heating is usually smaller than tens of micrometers, the TDTR method probes thermal transport in the surface layer of a bulk sample. In high thermal conductivity crystals at low temperatures near the peak thermal conductivity, phonons with mean free paths longer than the penetration depth traverse the temperature gradient ballistically and do not contribute to the thermal conductivity measured by the TDTR.17,18 Consequently, TDTR measurements have not obtained the peak thermal conductivity of high thermal conductivity crystals including diamond, BAs, and boron phosphide (BP).14,19–22

Besides these four commonly used measurement techniques, several other thermal transport measurement techniques have also been reported. As one example, a thermal mirage measurement technique was used for earlier measurements of the thermal conductivity of diamond samples.23 When the sample is heated by a laser beam, a temporarily varying gradient in the index of refraction in both the sample and the surrounding air is produced and can be detected from the measured deflection of a probe beam at a grazing incidence. This method yields the thermal diffusivity. The sensitivity and accuracy of this measurement method decrease at low temperatures because of the long thermal penetration depth in the sample, decreasing diffusivity of the air, diffusivity mismatch between the sample and the metal coating on top, and decreasing interface thermal conductance.24 Due to the small temperature rise in the ultrahigh thermal conductivity sample and the limited temperature sensitivity, this and other measurement methods have not been able to obtain the peak thermal conductivity of high-quality diamond and other ultrahigh thermal conductivity materials.

Here, we report a steady-state method that is designed for peak thermal conductivity measurements of high thermal conductivity rod samples. This method is based on the deposition and patterning of thin film resistive heaters and thermometers directly on the sample surface to control the interface thermal resistance between the sample and thermometers. The electrical lead wires for the heater and thermometers are arranged to be outside of the sample region of interest to eliminate the heat loss through the thermometers and the error caused by the interface thermal resistance. Heat loss from the heater into the lead wires is obtained directly from two measurements with different numbers of lead wires attached to the heater. Moreover, a differential Wheatstone bridge circuit is employed to measure the small temperature drop between two resistive thermometers. Meanwhile, the steady-state heating is stepped according to a sinusoidal pattern to allow Fast Fourier Transform (FFT) analysis of the data to improve the signal to noise ratio. A silicon sample is used as the calibration standard to investigate the accuracy of this method. A comparison between the obtained measurement results and literature reports and theoretical calculations is used to evaluate the effectiveness of this method for peak thermal conductivity measurements.

Instead of attaching thermocouples to the sample surface with thermal grease or glue, thin film Au/Cr serpentine resistive thermometers are deposited and patterned on the sample surface that is covered by a thin silicon nitride (SiNx) dielectric film, as shown in Figs. 1(a)1(f). A thin film Au/Cr serpentine resistive heater is patterned on the suspended end of the sample. The other end of the sample is joined with silver epoxy to a copper heat sink. Because of the use of microfabrication to pattern the heater and thermometers directly and precisely on the sample bar, this approach is particularly useful for mesoscopic samples with a lateral dimension in the range between hundreds of micrometers and several millimeters.

FIG. 1.

(a) Schematic illustration of the sample configuration. The arrows show the sensing current (i) flow path in the two thermometer lines. (b) Photograph of a patterned silicon bar sample assembled with a copper heat sink and a printed circuit board (PCB). Labels indicate the locations of the heater, thermometers 1 and 2. Scanning electron micrographs (SEMs) of (c) and (d) the patterned thin film Au/Cr serpentine resistive heater, where the gray particles on the metal pads are the residues of the removed Al:Si bonding wires, (e) the patterned thin film Au/Cr resistance thermometer 1, and (f) the patterned thin film Au/Cr resistance thermometer 2. (g) Thermal resistance circuit representation of the measured sample when only radiation loss from the thermometer lines is accounted for.

FIG. 1.

(a) Schematic illustration of the sample configuration. The arrows show the sensing current (i) flow path in the two thermometer lines. (b) Photograph of a patterned silicon bar sample assembled with a copper heat sink and a printed circuit board (PCB). Labels indicate the locations of the heater, thermometers 1 and 2. Scanning electron micrographs (SEMs) of (c) and (d) the patterned thin film Au/Cr serpentine resistive heater, where the gray particles on the metal pads are the residues of the removed Al:Si bonding wires, (e) the patterned thin film Au/Cr resistance thermometer 1, and (f) the patterned thin film Au/Cr resistance thermometer 2. (g) Thermal resistance circuit representation of the measured sample when only radiation loss from the thermometer lines is accounted for.

Close modal

The thermal resistance of the 300-nm-thick SiNx dielectric layer under the thermometer line at environmental temperature T0 = 300 K is calculated to be 91 K W−1 as

(1)

where Ac is the 300 × 2 μm2 cross-sectional area for vertical heat conduction across the dielectric between the thermometer line and the sample and κSiNx is the thermal conductivity of the SiNx film and reported to be about 5.5 W m−1 K−1.25 In addition, the thermal interface resistance (ri) for an evaporated or sputtered thin film is typically on the order of 3 × 10−8 K m2 W−1.26 Thus, the interface resistance of contact area Ac is

(2)

The total contact thermal resistance (Rc) between the thermometer and the sample is ∼141 K W−1 based on

(3)

Similarly, the thermal resistance across the metal layer is 0.7 K W−1 calculated based on

(4)

where κm and tm are the thermal conductivity and thickness of the metal film. In comparison, the resistance of radiation (Rrad) from the surface of the metal thermometer line is about 1.6 × 1016 K W−1 based on

(5)

where ϵm and σ are the emissivity of gold and Stefan–Boltzmann constant, respectively. Hence, the (Rc + Rm) value is about fourteen orders of magnitude smaller than Rrad. Consequently, the metal film temperature is expected to be essentially the same as the sample temperature based on the thermal resistance circuit representation of the sample [Fig. 1(g)].

There are four contacts to the serpentine heater and six contacts to the two thermometers. The sensing current flows between two contacts through the two thermometers that are connected in series. The contacts are connected to a printed circuit board (PCB) with 25.4 μm-diameter bonding wires that are made of 99% Al and 1% Si. The bonding wires for the serpentine heater and two thermometer lines are arranged to be at the two ends of the sample to minimize surface heat loss from the region between the heater and thermometer 2.

The sample assembled on the heat sink and the PCB is loaded into a high vacuum cryostat with two radiation shields to reduce the surface radiation and heat loss to gas molecules. The heat sink is attached directly to the cryostat stage with screws to improve the temperature uniformity between the heat sink and the sample stage. The sample stage temperature is in the range from 350 K to as low as 12 K for these measurements. The measured DC heating current (I) and voltage drop (V) across the serpentine are used to obtain the heating power as Q = IV.

Instead of measuring the electrical resistances (R1 and R2) of the two thermometer lines to obtain their temperature rises and the temperature drop as the small difference between two large values, we use the differential Wheatstone bridge circuit shown in Fig. 2 to measure the temperature drop between them directly. In this circuit, the reference resistor (Rr) is significantly larger than the sum of all other resistors in series so as to limit the sensing current (i) in the two resistance thermometers as

(6)

where Rl is the resistance of the lead wires including the Al:Si bonding wires, RL is the 50 Ω built-in output impedance of a SRS 830 lock-in amplifier, VL is the voltage output of the lock-in amplifier, and Vr is the voltage drop measured by a lock-in amplifier across Rr. Here, the leakage current through the lock-in amplifier is ignored because its 10 MΩ input impedance is much larger than Rr = 100 k Ω.

FIG. 2.

Electrical circuit for differential bridge measurement of the temperate drop between two thermometers with the electrical resistances R1 and R2. VL is the sinusoidal voltage output of a SRS 830 lock-in amplifier. RL is the built-in output impedance of the lock-in amplifier. Rr is a precision resistor for controlling the sensing current. Four lock-in amplifiers are used to measure the voltage drop Vr across Rr, V1 across R1, V2 across R2, and the difference (VS) between the voltage drops across R1 and R2 that are connected to two differential amplifiers with output impedance being R01 and R02. Rl represents the sum of the lead wire resistance including the resistance of the bonding wires for thermometers 1 and 2. The bridge is made of three precision resistors (R1a, R1b, and R2a) and a potentiometer (R2b).

FIG. 2.

Electrical circuit for differential bridge measurement of the temperate drop between two thermometers with the electrical resistances R1 and R2. VL is the sinusoidal voltage output of a SRS 830 lock-in amplifier. RL is the built-in output impedance of the lock-in amplifier. Rr is a precision resistor for controlling the sensing current. Four lock-in amplifiers are used to measure the voltage drop Vr across Rr, V1 across R1, V2 across R2, and the difference (VS) between the voltage drops across R1 and R2 that are connected to two differential amplifiers with output impedance being R01 and R02. Rl represents the sum of the lead wire resistance including the resistance of the bonding wires for thermometers 1 and 2. The bridge is made of three precision resistors (R1a, R1b, and R2a) and a potentiometer (R2b).

Close modal

In Fig. 2, the bridge circuit voltage signal is related to the electric resistances of the two thermometers according to

(7)

where g1 and g2 are the gains of the two SR560 voltage preamplifiers, which are used to measure the voltage drops (V1 and V2) across the two thermometers, R01 and R02 are the built-in output impendence of the two preamplifiers, R1a, R1b, and R2a are three precision resistors, and R2b is a potentiometer to be adjusted to minimize the following zero-point offset of the Wheatstone bridge output prior to the application of a heating current,

(8)

Here, Vs,0 and i0 are the Vs and i values and R1,0 and R2,0 are the resistances of the two thermometer lines at Q = 0.

Under steady-state Joule heating in the serpentine heater, the temperature rise (ΔT1 and ΔT2) of the two resistance thermometers leads to the following voltage signal:

(9)
(10)
(11)

where Δiii0, ΔR1R1R1,0, and ΔR2R2R2,0. In addition, α1dR1,0R1,0dT and α2dR2,0R2,0dT are the measured temperature coefficients of resistances (TCR) of the two thermometers. Thus, the temperature difference between the two thermometers can be obtained as

(12)

Due to the much larger Rr than ΔR1 and ΔR2, Δii0Vs,0 is much smaller than ΔVs. The use of the Wheatstone bridge to nullify Vs,0 helps to make Vs,0ii0α2θ2 much smaller than ΔVs. When α1 and α2 are close to each other for the two thermometer lines made of the same metal, the second term proportional to α1α2 is small so that ΔVs becomes the dominant signal source in the δθ expression. The resistance of the second thermometer adjacent to the heat sink is measured with the use of a lock-in amplifier to obtain θ2. Meanwhile, the heater temperature rise θh is obtained from the change in the four-probe resistance obtained from the measured DC current–voltage (IV) curve.27,28 Joule heating in the Al:Si bonding wires of the heater can be determined based on the voltage drop (Vw) measured across the Al:Si bonding wires with a voltage preamplifier in a four-probe configuration.

Based on the measured temperatures, the thermal conductivity of the sample can be obtained from a radiative fin model, in which the heat loss to the bonding wires of the heater and from the sample surface are accounted for. In this model, the total heat flow in the sample under the heater toward thermometer 1 is

(13)

where Qend is calculated as the radiation heat loss from the suspended end surface of the heater segment and Qloss,i is the heat loss into each of the four bonding wires. To evaluate the heat loss into the bonding wires, we treat the bonding wire as a fin with known temperatures Th and Tl at two ends (x = 0 and x = l). The steady-state heat diffusion equation for such a fin is29 

(14)

where κw, Aw, P, h, and q̇ are the thermal conductivity, cross-sectional area, perimeter, surface heat transfer coefficient, and Joule heating density of the Al:Si bonding wire. For the two bonding wires with Joule heating, q̇=VwIlAw, whereas q̇=0 for the other two wires without Joule heating. The solution of this fin equation yields the heat loss from the sample to the bonding wire as

(15)

where θiTiT0q̇AhP for i being h or l, fin parameter m=hP/κA with h=4ϵσT3, ϵ being the emissivity, and T0 being the sample stage temperature that is assumed to be the same as the temperature of the inner radiation shield mounted on the sample stage.30 The heat loss difference between two identical bonding wires with and without internal Joule heating in the wire is thus

(16)

In this model, the thermal conductivity of the bonding wire can be evaluated according to Wiedemann–Franz law from the measured electrical resistance (Rw) as

(17)

where Lo is the Lorenz number. The surface emissivity of the aluminum wire can be calculated from its resistivity (ρw) as31 

(18)

where the wavelength λ is chosen as λ=bT0 with b ≈ 2898 µm K based on Wien’s displacement law.

The heat flow across the SiNx film under the surface heater is much larger than that under each thermometer line. At low temperatures, the thermal resistance for heat flow from the surface serpentine heater across the SiNx film into the high thermal conductivity sample is appreciable compared to the thermal resistance of the sample. Thus, the temperature rise (θb) in the sample right underneath the serpentine heater is lower than the measured heater temperature rise (θh). In particular, there exists a three-dimensional (3D) temperature distribution in the sample in the vicinity of the surface heater. Because the surface radiation thermal resistance is many orders of magnitudes larger than the thermal resistance of the SiNx film, in comparison, the temperature distribution in the bar sample is expected to become one-dimensional (1D) at a region sufficiently far away from both the surface heater and the copper heat sink. For our samples, the distance between each of the two thermometer lines and the surface heater or the copper heat sink is about ten times the sample thickness. As such, 1D temperature distribution is established in the segment between the two thermometers to yield the same temperature of each thermometer line and the sample right beneath the thermometer line.

By treating the segment of the sample from the heater to the second thermometer as a fin, we use the measured temperature rise (θ1 = θ2 + δθ and θ2) at the first and second thermometer lines to calculate29 

(19)

where L1 is the distance between the heater and the first thermometer line, L2 is the distance between the two thermometer lines, and ms is the fin parameter of the sample and calculated with emissivity values for Si32 and SiNx. For the semi-transparent thin film SiNx layer, the emissivity can be calculated based on the real part of the refractive index and the extinction index of the SiNx thin film at each temperature.33,34

The fin heat transfer rate from the heater toward the two thermometers is

(20)

which can be used to obtain the sample thermal conductivity κs. The obtained κs is the weighted average between the silicon bar, the SiNx film, and the metal thermometer layer. Due to the much smaller cross section of the SiNx film and the thin metal film, the difference between κs and the silicon thermal conductivity is well within 0.01%.

As a comparison, a pure heat conduction model is also employed to analyze the measurement data to obtain the sample thermal conductivity as

(21)

where heat loss to the bonding wires is similarly accounted for. To ensure that the heat loss through the bonding wires is small compared to Qin as suggested by the model calculation, in addition, we removed the two bonding wires 3 and 4 that were used to measure the voltage drop of the heater and repeated the measurements to obtain the temperature drop (δθ′) between the two thermometers. Based on the heat conduction model, removal of these two bonding wires reduces the heat loss by

(22)

The measured δθδθδθ is within 2.5% at above 20 K, below which the uncertainty in the measured δθ′ and δθ increases due to reduced TCR as a result of the Kondo effect associated with electron scattering by magnetic impurity in Au.35 

The accuracy of this method is investigated with the use of the silicon bar sample shown in Fig. 1 as the calibration standard. The silicon bar sample was diced from an undoped silicon wafer. A 320-nm-thick SiNx thin film was sputtered on top of the sample for electrical insulation of the resistive thermometer from the semiconducting silicon sample. Both the heater and thermometers were then patterned from a 10-nm-thick chromium adhesion layer and a 100-nm-thick gold film deposited on the sample surface. Scanning electron microscopy (SEM) measurements show that the silicon sample is 624.4 μm wide, 514.4 μm thick, and the length between the two thermometer lines is 5.1 mm.

Figure 3 shows the measurement results of the silicon bar sample. The measured electrical resistance at the sample stage temperature without electrical heating is shown in Fig. 3(a) and can be used to obtain the TCR, which is then used to convert the measured electrical resistance rise at each heating current to the temperature rise. Figures 3(b) and 3(c) show the measured voltage drop of thermometer 2 and the bridge voltage signal as a function of the heating current. The results can be used to obtain the electrical resistance change of thermometer 2 as shown in Fig. 3(b) and temperature rise (θ2) and temperature difference (δθ) shown in Fig. 3(d) at the stage temperature T0 = 35 K, respectively.

FIG. 3.

(a) Measured electrical resistances of thermometers 1 (Re,1) and 2 (Re,2) at different sample stage temperatures (T0). The TCR is obtained as the local slope of the polynomial fit of each set of the measurement data. (b) Measured voltage of thermometer 2 (V2) (left y-axis) and the calculated average electrical resistance increase of thermometer 2 (ΔRe,2) (right y-axis) as a function of the heating current (I). (c) Measured voltage deference (VS) as a function of the heating current with the use of unitary gain for the two differential amplifiers. (d) Calculated average temperature rise of each thermometer (θ1, θ2), and temperature difference (δθ) between two thermometers as a function of heating current at T0 = 35 K. The error bars show the statistical uncertainty of 20 measurements and are smaller than many symbol sizes.

FIG. 3.

(a) Measured electrical resistances of thermometers 1 (Re,1) and 2 (Re,2) at different sample stage temperatures (T0). The TCR is obtained as the local slope of the polynomial fit of each set of the measurement data. (b) Measured voltage of thermometer 2 (V2) (left y-axis) and the calculated average electrical resistance increase of thermometer 2 (ΔRe,2) (right y-axis) as a function of the heating current (I). (c) Measured voltage deference (VS) as a function of the heating current with the use of unitary gain for the two differential amplifiers. (d) Calculated average temperature rise of each thermometer (θ1, θ2), and temperature difference (δθ) between two thermometers as a function of heating current at T0 = 35 K. The error bars show the statistical uncertainty of 20 measurements and are smaller than many symbol sizes.

Close modal

At each temperature, the heating current from a DC current source is stepped to take 16 discrete steady values according to a sinusoidal pattern as shown in Fig. 4(a), which is repeated for 20 cycles. Each discrete and steady current is held for 12 s to allow the thermal response to reach the steady state and be measured. Each cycle takes 192 s to yield an equivalent modulation frequency (fm) of about 0.005 Hz. FFT analysis is used to obtain the modulation amplitudes of the measured I, V, ΔRe,2, and δθ, as shown in Fig. 4. The FFT amplitudes are used to calculate the second harmonic amplitudes of Joule heating (Q), temperature rise of thermometer 2 (θ2), and temperature difference (δθ) with Eq. (12). Figures 4(d)4(f) show clear modulation at the second harmonic component corresponding to the modulation frequency (fm) of the heating current. These second harmonic amplitudes are used in the model calculation of the thermal conductivity. Compared to averaging the measurement results of the 20 cycles of data, this FFT approach can improve the signal to noise ratio when the temperature drop along the sample becomes small near the peak thermal conductivity.28 

FIG. 4.

(a) Discrete steady values of the heating current at a very low equivalent modulation frequency fm ≈ 0.005 Hz. Amplitude spectra of (b) the heating current (I), (c) heater voltage drop (V), (d) heating power (Q), (e) electrical resistance rise of thermometer 2 (ΔRe,2), and (f) temperature difference between two thermometers (δθ) and temperature rises (θ2) in the thermometer 2 as a function of the normalized frequency (f/fm) at T0 = 35 K. The gain is unitary for the two differential amplifiers.

FIG. 4.

(a) Discrete steady values of the heating current at a very low equivalent modulation frequency fm ≈ 0.005 Hz. Amplitude spectra of (b) the heating current (I), (c) heater voltage drop (V), (d) heating power (Q), (e) electrical resistance rise of thermometer 2 (ΔRe,2), and (f) temperature difference between two thermometers (δθ) and temperature rises (θ2) in the thermometer 2 as a function of the normalized frequency (f/fm) at T0 = 35 K. The gain is unitary for the two differential amplifiers.

Close modal

However, both the FFT spectra in Fig. 4(f) and the average temperature rise data in Fig. 3(d) show higher random noise in δθ measured by the Wheatstone bridge with unitary gain set for the two differential amplifiers than in θ2 measured by a lock-in amplifier. For this measurement where the sample stage temperature was highly stable at 35 K, the noise floor in the measured θ2 is as small as about 5 × 10−4 K. With either an increased stage temperature or reduced time for stabilizing the stage temperature, the noise floor in the measured θ2 can exceed 3 × 10−3 K, as shown in Fig. 5. Consequently, the noise floor can exceed 4 × 10−3 K in the calculated (θ1θ2), where both θ1 and θ2 are obtained from the voltage drops (V1 and V2) measured directly across the two thermometer lines. With the gain for the two differential amplifiers increased from 1 to 100, in comparison, the noise floor in δθ measured by the bridge circuit can be reduced to below 2 × 10−3 K. This comparison indicates that the resolution of the differential measurement method can be enhanced to the level of 1 × 10−3 K for measuring the small temperature drop along a high thermal conductivity material, even when the temperature fluctuation in the sample stage makes it impossible to obtain this resolution by measuring θ1 and θ2 directly to obtain (θ1θ2).

FIG. 5.

Comparison between the FFT amplitude spectra of (θ1θ2) and δθ, where (θ1θ2) is calculated from the voltage drops (V1 and V2) measured across the two thermometer lines and δθ is obtained from the Wheatstone bridge measurement. The stage temperature is 285 K. The gain for the two differential amplifiers is 100.

FIG. 5.

Comparison between the FFT amplitude spectra of (θ1θ2) and δθ, where (θ1θ2) is calculated from the voltage drops (V1 and V2) measured across the two thermometer lines and δθ is obtained from the Wheatstone bridge measurement. The stage temperature is 285 K. The gain for the two differential amplifiers is 100.

Close modal

Figure 6 shows the thermal conductivity measurement results based on the radiative fin model and the conduction model. The uncertainty in the thermal conductivity was calculated by a Monte Carlo simulation based on the uncertainties in the various measurements.28 For each measured modulated parameter such as the ΔVs/i term in Eq. (12), the local noise of the FFT amplitude spectrum is treated as a folded normal distribution to calculate the standard deviation of the distribution. The obtained FFT peak amplitudes are perturbed with normally distributed random generator with the corresponding standard deviations in the Monte Carlo calculation of δθ from Eq. (12). While only random uncertainty is shown for δθ and θ2 in Fig. 3(d), additional systematic errors are accounted for by the Monte Carlo simulation. Specifically, the preamplifier gain (g1) is similarly perturbed with a normally distributed random generator to account for the 1% gain error. In addition, the dominant systematic uncertainty source for δθ is the uncertainty in the measured α1α2 that shows a slight dependence on the actual temperatures (T1 and T2) of the two thermometer lines. At each stage temperature T0, T1, and T2 differ from T0 by a small amount depending on the heating level. Thus, in the Monte Carlo calculation, the thermometer temperature is perturbed with a uniformly distributed random generator within this amount to determine α1 and α2 corresponding to the perturbed thermometer temperature. The dimensions are measured several times to obtain the standard deviations of the dimensions. The perturbated values of δθ, Q, and the dimensions are then used to calculate the thermal conductivity based on Eqs. (19) and (20). This process is repeated for 5000 times to achieve independency on the number of Monte Carlo simulations. The uncertainty in the thermal conductivity is calculated based on the standard deviation of the 5000 calculated values of thermal conductivity, at 95% confidence interval.

FIG. 6.

Thermal conductivity of the Si bar sample in comparison with the theoretical calculation and literature results obtained by Inyushkin et al.36 of a silicon sample that is 2.5 mm wide, 2.0 mm thick, and 23 mm long, and obtained by Glassbrenner and Slack37 of a silicon rod with diameter 4.4 mm and length 20 mm, respectively. The silicon sample measured in this work is 624.4 μm wide, 514.4 μm thick, and the length between the two thermometers is 5.1 mm. The error bars for the radiative fin model results show the uncertainty calculated by Monte Carlo simulation at 95% confidence. The measurement results can be fitted by first-principles calculation with the use of Lb = 584 μm and g2 = 2.07 × 10−4.

FIG. 6.

Thermal conductivity of the Si bar sample in comparison with the theoretical calculation and literature results obtained by Inyushkin et al.36 of a silicon sample that is 2.5 mm wide, 2.0 mm thick, and 23 mm long, and obtained by Glassbrenner and Slack37 of a silicon rod with diameter 4.4 mm and length 20 mm, respectively. The silicon sample measured in this work is 624.4 μm wide, 514.4 μm thick, and the length between the two thermometers is 5.1 mm. The error bars for the radiative fin model results show the uncertainty calculated by Monte Carlo simulation at 95% confidence. The measurement results can be fitted by first-principles calculation with the use of Lb = 584 μm and g2 = 2.07 × 10−4.

Close modal

At temperature above 20 K, the difference between the conduction model and the radiation fin model is within 6.2%, which includes less than 3.3% radiation loss from the sample surface and less than 3.1% due to the difference between the fin model calculation of the heat loss into the bonding wires and the measurement with different numbers of bonding wires. At low temperatures, the radiation loss becomes negligible. At temperatures from 400 to 100 K, the differences between the current measurement results and literature results36,37 for larger size silicon samples are within 5% and 10%, respectively. At low temperatures, the increased surface-phonon scattering processes in the smaller size sample measured here lower the peak thermal conductivity and increase the peak temperature compared to the literature reports.

To verify that the lower peak thermal conductivity measured for the mesoscale silicon sample than the larger bulk sample is due to boundary scattering instead of measurement errors, we incorporate appropriate boundary scattering and defect scattering into three-phonon scattering in a theoretical calculation of the lattice thermal conductivity. This first-principles calculation is based on the density functional theory as implemented in Quantum Espresso,38,39 which yields results in good agreement with measurement results of a number of pure crystals without the use of fitting parameters.40,41 Norm-conserving pseudopotentials and the Perdew and Zunger42 exchange-correlation functional were used. An energy cutoff of 70 Ry was taken, which was sufficient to ensure that forces were accurately described. After structural relaxation, the calculated lattice constant was 5.40 Å, close to the measured value of 5.43 Å. Harmonic interatomic force constants (IFCs) were calculated using density functional perturbation theory, from which phonon dispersions were obtained, which showed good agreement with the experiment. Third-order anharmonic IFCs were calculated using a finite displacement approach with a 250 atom supercell. The harmonic and anharmonic IFCs were used to calculate three-phonon scattering rates. Here, a mass disorder model with mass variance parameter, g2,43 was used to describe phonon scattering by the naturally occurring silicon isotope mixture as well as by other point defects. Phonon scattering from sample boundaries was described by an empirical scattering rate formula, τb,λ1=vλ/Lb, where vλ is the magnitude of the phonon velocity in mode λ, while Lb is an averaged sample size. The linearized Peierls–Boltzmann transport equation was solved using our home-grown code to obtain the silicon thermal conductivity. Both g2 and Lb are taken as adjustable parameters and used to give a best fit to the measured thermal conductivity data over the range of temperatures studied. The measurement data can be fitted with the use of Lb = 584 μm and g2 = 2.07 × 10−4. As the Lb value is close to the (wt)1/2 = 567 μm of the silicon sample with width w = 624.4 μm and thickness t = 514.4 μm, diffuse surface scattering can explain the measured lower peak thermal conductivity of the mesoscale silicon sample than literature reports36,37 of larger silicon crystals.44 

These results demonstrate the accuracy and effectiveness of this differential thin film resistance thermometry approach for peak thermal conductivity measurement of high thermal conductivity materials. Direct deposition and patterning of the thin film thermometers on the sample surface allow for judicious arrangement of the electrical lead wires to essentially eliminate both the heat loss from the sample into the thermometers and the corresponding error associated with the interface thermal resistance. Measurements with different numbers of electrical connection wires to the heater directly yield the heat loss into the wires. The small temperature drop along a high thermal conductivity sample is resolved with a combination of Wheatstone bridge measurement of the responses of two resistive thermometers and FFT analysis of the steady-state data obtained at different heating rates. As this method is validated here with a silicon calibration standard, we expect that this method will find increasing use in the ongoing search of high thermal conductivity materials for thermal management, including diamond, BAs, BP, boron nitride (BN),45 BC2N,46 and theta-phase tantalum nitride (θ-TaN).47 

This work was supported by Office of Naval Research MURI Award No. N00014-16-1-2436.

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

2.
A. L.
Moore
and
L.
Shi
,
Mater. Today
17
,
163
(
2014
).
3.
M. M.
Waldrop
,
Nat. News
530
,
144
(
2016
).
4.
P.
Hui
and
H. S.
Tan
,
J. Appl. Phys.
75
,
748
(
1994
).
5.
S.
Jin
and
H.
Mavoori
,
J. Electron. Mater.
27
,
1148
(
1998
).
6.
Z.
Yan
,
G.
Liu
,
J. M.
Khan
, and
A. A.
Balandin
,
Nat. Commun.
3
,
827
(
2012
).
7.
H.
Huang
,
C. H.
Liu
,
Y.
Wu
, and
S.
Fan
,
Adv. Mater.
17
,
1652
(
2005
).
8.
M. J.
Biercuk
,
M. C.
Llaguno
,
M.
Radosavljevic
,
J. K.
Hyun
,
A. T.
Johnson
, and
J. E.
Fischer
,
Appl. Phys. Lett.
80
,
2767
(
2002
).
9.
Y.
Cui
,
Z.
Qin
,
H.
Wu
,
M.
Li
, and
Y.
Hu
,
Nat. Commun.
12
,
1284
(
2021
).
10.
D.
Salmon
,
Meas. Sci. Technol.
12
,
R89
(
2001
).
11.
S. S.
Ballard
,
K. A.
McCarthy
, and
W. C.
Davis
,
Rev. Sci. Instrum.
21
,
905
(
1950
).
12.
R. W.
Powell
,
Proc. Phys. Soc.
48
,
381
(
1936
).
13.
D. G.
Cahill
and
R. O.
Pohl
,
Annu. Rev. Phys. Chem.
39
,
93
(
1988
).
14.
F.
Tian
,
B.
Song
,
X.
Chen
,
N. K.
Ravichandran
,
Y.
Lv
,
K.
Chen
,
S.
Sullivan
,
J.
Kim
,
Y.
Zhou
,
T.-H.
Liu
,
M.
Goni
,
Z.
Ding
,
J.
Sun
,
G. A. G.
Udalamatta Gamage
,
H.
Sun
,
H.
Ziyaee
,
S.
Huyan
,
L.
Deng
,
J.
Zhou
,
A. J.
Schmidt
,
S.
Chen
,
C.-W.
Chu
,
P. Y.
Huang
,
D.
Broido
,
L.
Shi
,
G.
Chen
, and
Z.
Ren
,
Science
361
,
582
(
2018
).
15.
D. G.
Cahill
,
Rev. Sci. Instrum.
75
,
5119
(
2004
).
16.
H.-K.
Lyeo
and
D. G.
Cahill
,
Phys. Rev. B
73
,
144301
(
2006
).
17.
Y. K.
Koh
and
D. G.
Cahill
,
Phys. Rev. B
76
,
075207
(
2007
).
18.
R. B.
Wilson
and
D. G.
Cahill
,
Nat. Commun.
5
,
5075
(
2014
).
19.
S.
Li
,
Q.
Zheng
,
Y.
Lv
,
X.
Liu
,
X.
Wang
,
P. Y.
Huang
,
D. G.
Cahill
, and
B.
Lv
,
Science
361
,
579
(
2018
).
20.
J. S.
Kang
,
M.
Li
,
H.
Wu
,
H.
Nguyen
, and
Y.
Hu
,
Science
361
,
575
(
2018
).
21.
Q.
Zheng
,
S.
Li
,
C.
Li
,
Y.
Lv
,
X.
Liu
,
P. Y.
Huang
,
D. A.
Broido
,
B.
Lv
, and
D. G.
Cahill
,
Adv. Funct. Mater.
28
,
1805116
(
2018
).
22.
J. S.
Kang
,
H.
Wu
, and
Y.
Hu
,
Nano Lett.
17
,
7507
(
2017
).
23.
L.
Wei
,
P. K.
Kuo
,
R. L.
Thomas
,
T. R.
Anthony
, and
W. F.
Banholzer
,
Phys. Rev. Lett.
70
,
3764
(
1993
).
24.
L.
Wei
, “
Thermal property characterization of single-crystal diamond with varying isotopic composition
,” Ph.D. dissertation (
Wayne State University
,
1993
).
25.
L.
Shi
, “
Mesoscopic thermophysical measurements of microstructures and carbon nanotubes
,” Ph.D. dissertation (
University of California
,
Berkeley
,
2001
).
26.
H.-C.
Chien
,
D.-J.
Yao
,
M.-J.
Huang
, and
T.-Y.
Chang
,
Rev. Sci. Instrum.
79
,
054902
(
2008
).
27.
J. H.
Seol
,
A. L.
Moore
,
L.
Shi
,
I.
Jo
, and
Z.
Yao
,
J. Heat Transfer
133
,
022403
(
2010
).
28.
E.
Fleming
,
F.
Du
,
E.
Ou
,
L.
Dai
, and
L.
Shi
,
Carbon
145
,
195
(
2019
).
29.
T. L.
Bergman
,
F. P.
Incropera
,
D. P.
DeWitt
, and
A. S.
Lavine
,
Fundamentals of Heat and Mass Transfer
(
John Wiley & Sons
,
2011
).
30.
A. L.
Moore
and
L.
Shi
,
Meas. Sci. Technol.
22
,
015103
(
2010
).
31.
J. R.
Howell
,
M. P.
Mengüc
,
K.
Daun
, and
R.
Siegel
,
Thermal Radiation Heat Transfer
(
CRC Press
,
2020
).
32.
T.
Satō
,
Jpn. J. Appl. Phys., Part 1
6
,
339
(
1967
).
33.
G.
Cataldo
,
J. A.
Beall
,
H.-M.
Cho
,
B.
McAndrew
,
M. D.
Niemack
, and
E. J.
Wollack
,
Opt. Lett.
37
,
4200
(
2012
).
34.
J.
Kischkat
,
S.
Peters
,
B.
Gruska
,
M.
Semtsiv
,
M.
Chashnikova
,
M.
Klinkmüller
,
O.
Fedosenko
,
S.
Machulik
,
A.
Aleksandrova
,
G.
Monastyrskyi
,
Y.
Flores
, and
W.
Ted Masselink
,
Appl. Opt.
51
,
6789
(
2012
).
35.
P.
Nozières
and
A.
Blandin
,
J. Phys. France
41
,
193
(
1980
).
36.
A. V.
Inyushkin
,
A. N.
Taldenkov
,
A. M.
Gibin
,
A. V.
Gusev
, and
H.-J.
Pohl
,
Phys. Status Solidi C
1
,
2995
(
2004
).
37.
C. J.
Glassbrenner
and
G. A.
Slack
,
Phys. Rev.
134
,
A1058
(
1964
).
38.
P.
Giannozzi
,
S.
Baroni
,
N.
Bonini
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
,
G. L.
Chiarotti
,
M.
Cococcioni
,
I.
Dabo
,
A.
Dal Corso
,
S.
de Gironcoli
,
S.
Fabris
,
G.
Fratesi
,
R.
Gebauer
,
U.
Gerstmann
,
C.
Gougoussis
,
A.
Kokalj
,
M.
Lazzeri
,
L.
Martin-Samos
,
N.
Marzari
,
F.
Mauri
,
R.
Mazzarello
,
S.
Paolini
,
A.
Pasquarello
,
L.
Paulatto
,
C.
Sbraccia
,
S.
Scandolo
,
G.
Sclauzero
,
A. P.
Seitsonen
,
A.
Smogunov
,
P.
Umari
, and
R. M.
Wentzcovitch
,
J. Phys.: Condens. Matter
21
,
395502
(
2009
).
39.
P.
Giannozzi
,
O.
Andreussi
,
T.
Brumme
,
O.
Bunau
,
M.
Buongiorno Nardelli
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
,
M.
Cococcioni
,
N.
Colonna
,
I.
Carnimeo
,
A.
Dal Corso
,
S.
de Gironcoli
,
P.
Delugas
,
R. A.
DiStasio
,
A.
Ferretti
,
A.
Floris
,
G.
Fratesi
,
G.
Fugallo
,
R.
Gebauer
,
U.
Gerstmann
,
F.
Giustino
,
T.
Gorni
,
J.
Jia
,
M.
Kawamura
,
H.-Y.
Ko
,
A.
Kokalj
,
E.
Küçükbenli
,
M.
Lazzeri
,
M.
Marsili
,
N.
Marzari
,
F.
Mauri
,
N. L.
Nguyen
,
H.-V.
Nguyen
,
A.
Otero-de-la-Roza
,
L.
Paulatto
,
S.
Poncé
,
D.
Rocca
,
R.
Sabatini
,
B.
Santra
,
M.
Schlipf
,
A. P.
Seitsonen
,
A.
Smogunov
,
I.
Timrov
,
T.
Thonhauser
,
P.
Umari
,
N.
Vast
,
X.
Wu
, and
S.
Baroni
,
J. Phys.: Condens. Matter
29
,
465901
(
2017
).
40.
D. A.
Broido
,
M.
Malorny
,
G.
Birner
,
N.
Mingo
, and
D. A.
Stewart
,
Appl. Phys. Lett.
91
,
231922
(
2007
).
41.
A.
Ward
,
D. A.
Broido
,
D. A.
Stewart
, and
G.
Deinzer
,
Phys. Rev. B
80
,
125203
(
2009
).
42.
J. P.
Perdew
and
A.
Zunger
,
Phys. Rev. B
23
,
5048
(
1981
).
43.
S.-i.
Tamura
,
Phys. Rev. B
27
,
858
(
1983
).
44.
Z.
Wang
and
N.
Mingo
,
Appl. Phys. Lett.
99
,
101903
(
2011
).
45.
K.
Chen
,
B.
Song
,
N. K.
Ravichandran
,
Q.
Zheng
,
X.
Chen
,
H.
Lee
,
H.
Sun
,
S.
Li
,
G. A. G.
Udalamatta Gamage
,
F.
Tian
,
Z.
Ding
,
Q.
Song
,
A.
Rai
,
H.
Wu
,
P.
Koirala
,
A. J.
Schmidt
,
K.
Watanabe
,
B.
Lv
,
Z.
Ren
,
L.
Shi
,
D. G.
Cahill
,
T.
Taniguchi
,
D.
Broido
, and
G.
Chen
,
Science
367
,
555
(
2020
).
46.
S. N.
Sadeghi
,
S. M. V.
Allaei
,
M.
Zebarjadi
, and
K.
Esfarjani
,
J. Mater. Chem. C
8
,
015705
(
2020
).
47.
A.
Kundu
,
X.
Yang
,
J.
Ma
,
T.
Feng
,
J.
Carrete
,
X.
Ruan
,
G. K. H.
Madsen
, and
W.
Li
,
Phys. Rev. Lett.
126
,
115901
(
2021
).