University of California Postprints

Similar documents
Segmented Power Generator Modules of Bi 2 Te 3 and ErAs:InGaAlAs Embedded with ErAs Nanoparticles

Transient Harman Measurement of the Cross-plane ZT of InGaAs/InGaAlAs Superlattices with Embedded ErAs Nanoparticles

Segmented Power Generator Modules of Bi 2 Te 3 and ErAs:InGaAlAs Embedded with ErAs Nanoparticles

Transient Harman Measurement of the Cross-plane ZT of InGaAs/InGaAlAs Superlattices with Embedded ErAs Nanoparticles

TRANSIENT THERMAL CHARACTERIZATION OF ERAS/IN 0.53 GA 0.47 AS THERMOELECTRIC MODULE

Seebeck Enhancement Through Miniband Conduction in III V Semiconductor Superlattices at Low Temperatures

Thermoelectricity: From Atoms to Systems

Thermionic power generation at high temperatures using SiGe/ Si superlattices

Quantum Simulation of Nanocrystalline Composite Thermoelectric. Properties

Computational Study of the Electronic Performance of Cross-Plane Superlattice Peltier Devices

Thermal conductivity of symmetrically strained Si/Ge superlattices

Measurement of Seebeck coefficient perpendicular to SiGe superlattice

Thermoelectric power generators are clean, noise free

Thermoelectricity: From Atoms to Systems

MOLECULAR DYNAMICS SIMULATION OF THERMAL CONDUCTIVITY OF NANOCRYSTALLINE COMPOSITE FILMS

Nashville, Tennessee b Department of Mechanical Engineering, Vanderbilt University,

QUANTUM SIMULATION OF NANOCRYSTALLINE COMPOSITE THERMOELECTRIC PROPERTIES

eterostrueture Integrated Thermionic Refrigeration

Exact Solution of a Constrained. Optimization Problem in Thermoelectric Cooling

THERMAL TRANSIENT CHARACTEROZATION OF PACKAGED THIN FILM MICROCOOLERS. Kazuhiko Fukutani, Rajeev Singh and Ali Shakouri

University of California Postprints

Nanoscale Thermal Transport and Microrefrigerators on a Chip

Increased Phonon Scattering by Nanograins and Point Defects in Nanostructured Silicon with a Low Concentration of Germanium

Micro-coolers fabricated as a component in an integrated circuit

Comparison of solid-state thermionic refrigeration with thermoelectric refrigeration

Very high thermoelectric power factor in a Fe 3 O 4 /SiO 2 /p-type Si(100) heterostructure

M R S Internet Journal of Nitride Semiconductor Research

3-1-2 GaSb Quantum Cascade Laser

Electronic and thermoelectric transport in semiconductor and metallic superlattices

Exact Solution of a Constrained. Optimization Problem in Thermoelectric Cooling

SUPPLEMENTARY INFORMATION

Lecture 9: Metal-semiconductor junctions

THE EFFICIENCY of thermoelectric materials is usually

Supplementary Information for On-chip cooling by superlattice based thin-film thermoelectrics

Models for the electronic transport properties of thermoelectric materials. Lars Corbijn van Willenswaard

ac ballistic transport in a two-dimensional electron gas measured in GaAs/ AlGaAs heterostructures

Material Optimization for Heterostructure Integrated Thermionic Coolers

Simulation and Analysis of an Integrated Device to Simultaneously Characterize Thermal and Thermoelectric Properties

Modeling of Thermoelectric Properties of Semi-Conductor Thin Films With Quantum and Scattering Effects

Electron Energy, E E = 0. Free electron. 3s Band 2p Band Overlapping energy bands. 3p 3s 2p 2s. 2s Band. Electrons. 1s ATOM SOLID.

The lattice thermal conductivity of a semiconductor nanowire

Quantum Modeling of Thermoelectric Properties of Si/Ge/Si Superlattices

Homework Week 3: Nanoscale and macroscale characterization Thermoelectricity: From Atoms to Systems

Thermal conductivity of semiconductor superlattices: Experimental study of interface scattering

Asymmetrical heating behavior of doped Si channels in bulk silicon and in silicon-on-insulator under high current stress

Quantum Dot Superlattice Thermoelectric Materials and Devices. T. C. Harman, P. J. Taylor, M. P. Walsh, and B. E. LaForge. Supplementary Material

Measurement of Thin Film Isotropic and Anisotropic Thermal Conductivity Using 3ω and Thermoreflectance Imaging

Pseudomorphic HEMT quantum well AlGaAs/InGaAs/GaAs with AlAs:δ-Si donor layer

Short Time Transient Behavior of SiGe-based Microrefrigerators.

Research to Improve Photovoltaic (PV) Cell Efficiency by Hybrid Combination of PV and Thermoelectric Cell Elements.

Nanostructured Thermoelectric Materials: From Superlattices to Nanocomposites

Model of transport properties of thermoelectric nanocomposite materials

Surfaces, Interfaces, and Layered Devices

Minimum superlattice thermal conductivity from molecular dynamics

OPTI510R: Photonics. Khanh Kieu College of Optical Sciences, University of Arizona Meinel building R.626

Simulation and Optimization of an In-plane Thermal Conductivity Measurement Structure for Silicon Nanostructures

White Rose Research Online URL for this paper:

Quantification of Trap State Densities at High-k/III-V Interfaces

Comparison of Ultra-Thin InAs and InGaAs Quantum Wells and Ultra-Thin-Body Surface-Channel MOSFETs

GeSi Quantum Dot Superlattices

Traps in MOCVD n-gan Studied by Deep Level Transient Spectroscopy and Minority Carrier Transient Spectroscopy

Anisotropic spin splitting in InGaAs wire structures

arxiv: v1 [cond-mat.mtrl-sci] 18 Sep 2013

Nanoelectronic Thermoelectric Energy Generation

Enhancement of the thermoelectric figure of merit in p- and n-type Cu/Bi-Te/Cu composites

Electrical properties of InSb quantum wells remotely doped with Si

Influence of Dimensionality on Thermoelectric Device Performance

MSE 310/ECE 340: Electrical Properties of Materials Fall 2014 Department of Materials Science and Engineering Boise State University

Al/Ti/4H SiC Schottky barrier diodes with inhomogeneous barrier heights

Monte Carlo Based Calculation of Electron Transport Properties in Bulk InAs, AlAs and InAlAs

Thermoelectric Oxide Materials For Electric Power Generation

Magnon-drag thermopile

Carrier Transport Mechanisms of a-gaas/ n-si Heterojunctions

Thermoelectric and electrical properties of Si-doped InSb thin films. University, Japan

Cross-Sectional Scanning Tunneling Microscopy and Spectroscopy of Semimetallic ErAs Nanostructures Embedded in GaAs

CME 300 Properties of Materials. ANSWERS: Homework 9 November 26, As atoms approach each other in the solid state the quantized energy states:

SUPPLEMENTARY INFORMATION

Thermal characterization of Au-Si multilayer using 3- omega method

Thermal conductivity of GaAs/Ge nanostructures

Supplementary Information Interfacial Engineering of Semiconductor Superconductor Junctions for High Performance Micro-Coolers

Electron leakage effects on GaN-based light-emitting diodes

Current mechanisms Exam January 27, 2012

Controlling Graphene Ultrafast Hot Carrier Response from Metal-like. to Semiconductor-like by Electrostatic Gating

Copyright 2010 Yee Kan Koh

Simulation of GaN-based Light-Emitting Devices

Temperature Dependent Optical Band Gap Measurements of III-V films by Low Temperature Photoluminescence Spectroscopy

Comparison of the 3ω method and time-domain

NONLINEAR TRANSITIONS IN SINGLE, DOUBLE, AND TRIPLE δ-doped GaAs STRUCTURES

Reduction of Thermal Conductivity by Nanoscale 3D Phononic Crystal

Semiconductor Nanowires: Motivation

Kinetic Monte Carlo simulation of semiconductor quantum dot growth

DISTRIBUTION OF POTENTIAL BARRIER HEIGHT LOCAL VALUES AT Al-SiO 2 AND Si-SiO 2 INTERFACES OF THE METAL-OXIDE-SEMICONDUCTOR (MOS) STRUCTURES

Chris G. Van de Walle

Section 3: Nanoscience and Thermoelectric Energy by Alexis R. Abramson

Hall Effect Measurements on New Thermoelectric Materials

Physics of Semiconductors

Quantum and Non-local Transport Models in Crosslight Device Simulators. Copyright 2008 Crosslight Software Inc.

Lecture 3: Heterostructures, Quasielectric Fields, and Quantum Structures

Midterm I - Solutions

1. Binary III-V compounds 2 p From which atoms are the 16 binary III-V compounds formed?...column III B, Al, Ga and In...column V N, P, As and Sb...

Transcription:

University of California Postprints Year 2007 Paper 2192 Cross-plane Seebeck coefficient of ErAs : InGaAs/InGaAlAs superlattices G H. Zeng, JMO Zide, W Kim, J E. Bowers, A C. Gossard, Z X. Bian, Y Zhang, A Shakouri, S L. Singer, and A Majumdar, Cross-plane Seebeck coefficient of ErAs : InGaAs/InGaAlAs superlattices (2007). Journal of Applied Physics. 101 (3), Article 034502. Postprint available free at: http://repositories.cdlib.org/postprints/2192 Posted at the escholarship Repository, University of California. http://repositories.cdlib.org/postprints/2192

Cross-plane Seebeck coefficient of ErAs : InGaAs/InGaAlAs superlattices Abstract We characterize cross-plane and in-plane Seebeck coefficients for ErAs:InGaAs/InGaAlAs superlattices with different carrier concentrations using test patterns integrated with microheaters. The microheater creates a local temperature difference, and the cross-plane Seebeck coefficients of the superlattices are determined by a combination of experimental measurements and finite element simulations. The cross-plane Seebeck coefficients are compared to the in-plane Seebeck coefficients and a significant increase in the cross-plane Seebeck coefficient over the in-plane Seebeck coefficient is observed. Differences between cross-plane and inplane Seebeck coefficients decrease as the carrier concentration increases, which is indicative of heterostructure thermionic emission in the cross-plane direction. (c) 2007 American Institute of Physics.

JOURNAL OF APPLIED PHYSICS 101, 034502 2007 Cross-plane Seebeck coefficient of ErAs: InGaAs/ InGaAlAs superlattices Gehong Zeng a Department of Electrical and Computer Engineering, University of California, Santa Barbara, California 93106 Joshua M. O. Zide Materials Department, University of California, Santa Barbara, California 93106 Woochul Kim Department of Mechanical Engineering, University of California, Berkeley, California 94720 John E. Bowers Department of Electrical and Computer Engineering, University of California, Santa Barbara, California 93106 Arthur C. Gossard Materials Department, University of California, Santa Barbara, California 93106 Zhixi Bian, Yan Zhang, and Ali Shakouri Electrical Engineering Department, University of California, Santa Cruz, California 95064 Suzanne L. Singer and Arun Majumdar Department of Mechanical Engineering, University of California, Berkeley, California 94720 Received 13 July 2006; accepted 29 November 2006; published online 6 February 2007 We characterize cross-plane and in-plane Seebeck coefficients for ErAs: InGaAs/ InGaAlAs superlattices with different carrier concentrations using test patterns integrated with microheaters. The microheater creates a local temperature difference, and the cross-plane Seebeck coefficients of the superlattices are determined by a combination of experimental measurements and finite element simulations. The cross-plane Seebeck coefficients are compared to the in-plane Seebeck coefficients and a significant increase in the cross-plane Seebeck coefficient over the in-plane Seebeck coefficient is observed. Differences between cross-plane and in-plane Seebeck coefficients decrease as the carrier concentration increases, which is indicative of heterostructure thermionic emission in the cross-plane direction. 2007 American Institute of Physics. DOI: 10.1063/1.2433751 I. INTRODUCTION Thermoelectric materials can be used for solid state refrigeration and power generation devices. The performance of thermoelectric energy conversion devices depends on the thermoelectric figure of merit ZT of a material, which is defined as ZT=S 2 T/k, where S,, k, and T are the Seebeck coefficient, electrical conductivity, thermal conductivity, and absolute temperature, respectively. It has been proposed that semiconductor superlattices are a promising thermoelectric material because of i enhancement in the power factor S 2 due to the quantum confinement effect, 1 ii thermal conductivity reduction due to phonon interface scattering, 2 and iii increase in Seebeck coefficient due to thermionic emission. 3 5 So far only the thermal conductivity reduction has been demonstrated 6,7 conclusively, and much of the recent increase in ZT out of the superlattices 8,9 comes mainly from the thermal conductivity reduction. 10 However, ZT enhancement due to the power factor increase in a superlattice has not yet been quantified mainly because of the challenge in measuring the cross-plane Seebeck coefficient and crossplane electrical conductivity. Due to the nonisotropic superlattice structure, the in-plane and cross-plane Seebeck coefficients are usually different. The measurement of in-plane The ErAs: InGaAs/ InGaAlAs superlattice samples were grown using a Varian Gen II molecular beam epitaxy system on 100 n-type InP substrates with a buffer layer of 50 nm InGaAs. 14 16 The samples consist of an InGaAs/InGaAlAs superlattice structure with semimetallic ErAs particles randomly distributed through InGaAs layers. The ErAs particles act as a n-type dopant in the InGaAs, 14 and Hall measurea Electronic mail: gehong@ece.ucsb.edu Seebeck coefficients is straightforward: a temperature difference is built up across a test sample plate, and by measuring the output voltage V at different temperatures T, the Seebeck coefficient of the material can be determined by V/ T. In contrast, the measurement of the cross-plane Seebeck coefficient is challenging because it is difficult to determine the voltage and temperature drop across a thin film superlattice layer only a few microns thick. 11 13 In this paper, we report a method for experimental characterization of cross-plane Seebeck coefficients of superlattice samples by using device test patterns combined with three-dimensional 3D thermal simulations. Superlattice structures are designed to increase the cross-plane Seebeck coefficient by using heterostructure thermionic emission. 3 5 An explanation of the cross-plane Seebeck coefficient enhancement over inplane Seebeck coefficient is provided. II. MATERIAL STRUCTURES 0021-8979/2007/101 3 /034502/5/$23.00 101, 034502-1 2007 American Institute of Physics

034502-2 Zeng et al. J. Appl. Phys. 101, 034502 2007 FIG. 2. The combined Seebeck coefficients for the four superlattice samples and the Seebeck coefficients of their InP substrates were measured using device test patterns. FIG. 1. Color online The diagram of the test pattern device structure: a integrated microheaters on top of the superlattice mesa and b integrated microheaters on the InP substrate. ments in a van der Pauw geometry show that without codoping, these randomly distributed ErAs particles provide an electron concentration around 2 10 18 cm 3. 16 The superlattice structure consisted of 70 periods of 10 nm InAlGaAs and 20 nm InGaAs. The silicon codoping values for the four superlattice samples were 0, 2 10 18, 4 10 18, and 8 10 18 cm 3, respectively. Both ErAs and Si are electron donors and the total doping values for each sample are 2 10 18,4 10 18,6 10 18, and 1 10 19 cm 3, respectively. The 10 nm InGaAlAs layer consists of a digital alloy of 60% InGaAs and 40% InAlAs, and 0.3% erbium is randomly distributed in the 20 nm n-ingaas layer in addition to silicon codoping. The reason for choosing the combination of these materials to increase ZT is described herein. InGaAlAs is known to have a band offset of E=0.2 ev with respect to ErAs:InGaAs, which could increase the Seebeck coefficient of the sample by approximately E/qT, where q is the charge of an electron. The role of ErAs:InGaAs is to reduce the thermal conductivity below that of InGaAs or InGaAlAs. It has been shown previously that incorporating ErAs nanoparticles 1 5 nm in diameter effectively reduces the thermal conductivity below the InGaAs alloy because ErAs nanoparticles effectively scatter long and midwavelength phonons, while atomic substitution in the InGaAs alloy scatter short wavelength phonons. 17 The hypothesis is that the combination of a large Seebeck coefficient, low thermal conductivity, and possibly high electrical conductivity could potentially increase ZT. III. DEVICE FABRICATION AND MEASUREMENTS Figure 1 shows the test pattern structure with integrated microheaters on both the top of the superlattice mesa a and on the InP substrate b. A3 m ErAs:InGaAs/InGaAlAs superlattice mesa was first formed by reactive ion etching, and then the surface was passivated with 300 nm of Si 3 N 4. Contact windows were opened on the Si 3 N 4, and a contact metal layer of Ni/GeAu/Ni/Au was deposited via electron beam evaporation. This metal layer is to form Ohmic contact to the semiconductor and the probe electrode as well. Si 3 N 4 and SiO 2 double insulating layers of 300 nm thick in total were deposited on top of the contact metal layer, and microheaters of Ti/Au were then patterned on top of the insulating layers. This same pattern was formed on top of the superlattice mesa as well as on the InP substrate. The patterns on top of the superlattice mesa are used for the combined Seebeck coefficient measurement of superlattice and InP substrate, S com, while those on the InP substrate are used for the Seebeck coefficient measurement of the InP substrate, S InP. When a current flows through the microheater on the top of the InP substrate, a local temperature difference is built up, and the thermoelectric voltage, V InP, and the temperature difference, T InP, between the top contact and the ground contact are simultaneously measured. The InP substrate Seebeck coefficient S InP is the slope of the linearly fitted V InP vs T InP, which is expressed as S InP = V InP / T InP. Similarly, by using a test pattern on the top of superlattice mesa and a ground electrode on the InP substrate, the temperature difference T com versus output voltage V com can be obtained. The measurement linear fitting value S com = V com / T com is a combined Seebeck coefficient, which can be expressed as S com = S SL T SL T com + S InP T InP T com, where T com is the temperature drop across the superlattice and InP substrate, T SL and T InP are the temperature drop across the superlattice and InP substrate, and S SL is the crossplane Seebeck coefficient of the superlattice. The measurement results of the InP substrate and combined Seebeck coefficients for the four samples using device test patterns are shown in Fig. 2. Because there is no direct way to measure the temperature at the interface of the superlattice and InP substrate, a 3D thermal model was built, by which we could estimate the ratio of the temperature drop across the superlattice and the 1

034502-3 Zeng et al. J. Appl. Phys. 101, 034502 2007 FIG. 3. The lapped InP substrates of the four superlattice samples were measured for the calibration of the results measured using device test patterns, and the Seebeck coefficients were obtained from the slope of the linear fit of the measured V vs T. InP substrate, T SL / T InP. The thermal conductivity of each individual layer the ErAs:InGaAs/InGaAlAs superlattice, 18 the InP substrate, and the combined insulating layer of Si 3 N 4 and SiO 2 was measured at room temperature using the 3 technique 19 and used as an input to the simulation. The simulation results show that the temperature difference ratio of T SL / T InP for the device test patterns is close to 1, therefore by using Eq. 1, the cross-plane Seebeck coefficients for the four ErAs: InGaAs/ InGaAlAs superlattice samples with carrier concentrations of 2 10 18, 4 10 18, 6 10 18, and 1 10 19 cm 3 can be calculated as 405, 326, 265, and 232 V/K, respectively. The above measurements assumed that the temperature measured on top of the microheaters was the same as the temperature at the interface of the superlattice surface and contact metal, which should be corrected through calibration. As the transverse and longitudinal masses of the conduction band Gamma valley minima of InP substrate are almost the same, the Seebeck coefficients of the InP substrate along the in-plane and cross-plane directions are supposed to be very similar. 20,21 The calibrations were carried out by removing all the superlattice structures of the four samples and then measuring the Seebeck coefficients of their InP substrates separately. The results were compared with those measured using device test patterns. Each superlattice sample was cut into 5 20 mm 2 in strip size and lapped down about 40 m from the superlattice surface to totally remove the superlattice structures. The indium electrode was patterned at each end of the strip sample. The sample was placed across two separate thermoelectric cooler stages at two different temperatures. The output voltage V and the temperature difference T were measured simultaneously with a multimeter and two thermocouples. Figure 3 shows the linear curve fitting of the measured V vs T, and the Seebeck coefficient values of the four InP substrates for the four superlattice samples with the carrier concentrations of 2 10 18,4 10 18,6 10 18, and 1 10 19 cm 3 are 204, 221, 207, and 200 V/K, respectively. The slight doping difference in the four different InP substrate samples could result in minor differences among the InP substrate Seebeck coefficients. By comparing the InP substrate Seebeck coefficients measured using test patterns with the direct measurements using the lapped InP substrate samples, we find that the test pattern measurement results of the four InP substrates are about 20%, 23%, 22%, and 20% less than those of the direct measurements, respectively. The local temperature drops are mainly across the three sections: i the insulating Si 3 N 4 and SiO 2 layers between the contact metal and the microheater, T SiN, ii ErAs:InGaAs/InGaAlAs superlattice layer, T SL, and iii InP substrate, T InP. Finite element thermal simulations show that the ratio of T SiN / InP is about 1/5 for the InP substrate Seebeck coefficient test pattern shown in Fig. 1 b, and the ratio of T SiN / T SL / T InP is about 1/4.5/4.5 for the combined Seebeck coefficient test pattern shown in Fig. 1 a. When taking the ratio of temperature drop in these three sections into account, the overestimation for T InP in the InP substrate test pattern and T com in the superlattice test pattern is around 20% and 10%, respectively. It was the overestimation of T InP that produced the measurement differences between those using InP test patterns and those using direct measurement methods. Therefore, the cross-plane Seebeck coefficients of the four superlattices with carrier concentrations of 2 10 18, 4 10 18, 6 10 18, and 1 10 19 cm 3 are 427, 330, 266, and 235 V/K, respectively, after calibration using Eq. 1 based on the V com and T com measured on test patterns, the Seebeck coefficients of the four InP substrates measured with direct measurement methods, and ratio of temperature drop T SiN / T SL / T InP calculated with finite element thermal simulations. As our material combination of InGaAs wells with In- GaAlAs barriers are designed to enhance the cross-plane Seebeck coefficient by using heterostructure thermionic emission, 3 5 the in-plane and cross-plane Seebeck coefficients are expected to be different. In order to find out this difference, we grew the second set of four 30-period samples of the same superlattice structures and doping concentrations on semi-insulating InP substrates for in-plane Seebeck measurements. The in-plane Seebeck coefficients were measured with the direct measurement method similar to the calibration measurements of those lapped InP substrates described above. The in-plane Seebeck coefficients of 183, 120, 100, and 70 V/K were measured for the four samples with carrier concentrations of 2 10 18, 4 10 18, 6 10 18, and 1 10 19 cm 3, respectively. The measurement results of the in-plane and cross-plane Seebeck coefficients for the two sets of eight samples with different carrier concentrations are shown in Fig. 4. IV. DISCUSSION The absolute value of the Seebeck coefficient of a bulk material decreases as the carrier concentration increases and the Fermi level moves deeper into the conduction band. 22 This conclusion can be reached from the formula of Seebeck coefficient for a bulk material,

034502-4 Zeng et al. J. Appl. Phys. 101, 034502 2007 FIG. 4. Measurement results of the cross-plane and in-plane Seebeck coefficients for the two sets of eight ErAs: InGaAs/ InGaAlAs superlattice samples. S = 1 qt d E E E F de, 2 d E de where d and E F are the differential electrical conductivity and Fermi energy, respectively. Usually the same principle can be applied to heterostructures. Both the cross-plane and in-plane Seebeck coefficients of superlattices decrease when the carrier concentration increases. Band diagrams of ErAs: InGaAs/ InGaAlAs superlattices were calculated by solving Schrödinger and Poisson equations self-consistently, and the dependence of E F E C on carrier concentration is shown in Fig. 5, where E C of wells for in plane and barriers for cross plane are used, respectively. The enhancement of the Seebeck coefficients in the cross plane over the in-plane direction is obvious in Fig. 4, which results from the energy filtering by potential barriers. When the Fermi level moves deep inside the conduction band of the InGaAlAs barriers, the energy filtering effect is minimal. In this case, the change of the cross-plane Seebeck coefficient with carrier concentrations would be very similar to that of the in-plane Seebeck coefficient. The experimental results support this reasoning: in the doping range of 2 10 18 4 10 18 cm 3, the E F is nearby to E C for the barriers, and the cross-plane Seebeck coefficient value of Eq. 2 changes more quickly with different carrier concentrations. The differences between the cross-plane and in-plane Seebeck coefficients of the samples with effective dopings of 2 10 18 and 4 10 18 cm 3 are 244 and 210 V/K, respectively. When the carrier concentration is over 6 10 18 cm 3, the Fermi level is deep inside the conduction band for both the wells and barriers, and the integration value rate of change of Eq. 2 with carrier concentrations gradually becomes very similar for both the wells in plane and barriers cross plane. Our measurement results show that the differences of in-plane and cross-plane Seebeck coefficients for the two samples with carrier concentrations of 6 10 18 and 1 10 19 cm 3 are very close to each other Fig. 4, and the differences are 166 and 165 V/K, respectively. V. CONCLUSION In summary, a method to measure the cross-plane Seebeck coefficient of a superlattice is described. Two sets of eight ErAs: InGaAs/ InGaAlAs superlattice samples with different carrier concentrations were systematically grown and used in this study. We measured combined Seebeck coefficients of the superlattice and the substrate using patterned microheaters. Then, the cross-plane Seebeck coefficients of the superlattice were deduced with the additional information of the temperature difference ratio, T SL / T InP, determined by finite element thermal simulations. InGaAlAs has a band offset of E=0.2 ev with respect to ErAs:InGaAs. Therefore, electrons see an energy barrier in the cross-plane direction when the Fermi level is below the conduction band minimum of InGaAlAs. Experimental results showed that there is a significant increase of the cross-plane Seebeck coefficient over the in-plane Seebeck coefficients, and both the cross-plane and in-plane Seebeck coefficients of the superlattices decrease as the carrier concentration increases. The differences between the cross-plane and in-plane Seebeck coefficients decrease with the increase of carrier concentrations. This could indicate heterostructure thermionic emission in the cross-plane direction: as the carrier concentration increases, energy differences, which are related to thermionic emission, between the Fermi level and conduction band minimum of InGaAlAs decrease. Therefore, differences between the cross-plane and in-plane Seebeck coefficients should decrease. ACKNOWLEDGMENTS The authors acknowledge useful discussions with Dr. Mihal Gross. This work is supported by the Office of Naval Research Thermionic Energy Conversion Center MURI. FIG. 5. Calculation results of the energy difference of E F E C for the well and barrier of ErAs/ InGaAs superlattice with different carrier concentrations. 1 L. D. Hicks and M. S. Dresselhaus, Phys. Rev. B 47, 12727 1993. 2 M. V. Simkin and G. D. Mahan, Phys. Rev. Lett. 84, 927 2000. 3 A. Shakouri and J. E. Bowers, Appl. Phys. Lett. 71, 1234 1997. 4 D. Vashaee and A. Shakouri, J. Appl. Phys. 95, 1233 2004. 5 D. Vashaee and A. Shakouri, Phys. Rev. Lett. 92, 106103 2004. 6 S. T. Huxtable et al., Appl. Phys. Lett. 80, 1737 2002.

034502-5 Zeng et al. J. Appl. Phys. 101, 034502 2007 7 R. Venkatasubramanian, Phys. Rev. B 61, 3091 2000. 8 T. C. Harman, P. J. Taylor, M. P. Walsh, and B. E. LaForge, Science 297, 2229 2002. 9 R. Venkatasubramanian, E. Siivola, T. Colpitts, and B. O Quinn, Nature London 413, 597 2001. 10 G. Chen, A. Narayanaswamy, and C. Dames, Superlattices Microstruct. 35, 161 2004. 11 B. Yang, J. L. Liu, K. L. Wang, and G. Chen, Appl. Phys. Lett. 80, 1758 2002. 12 B. Yang, W. L. Liu, J. L. Liu, K. L. Wang, and G. Chen, Appl. Phys. Lett. 81, 3588 2002. 13 Y. Zhang, G. Zeng, R. Singh, J. Christofferson, E. Croke, J. E. Bowers, and A. Shakouri, in Measurement of Seebeck coefficient perpendicular to SiGe superlattice, 21st International Conference on Thermoelectronics, 25 29 August 2002, pp. 329 332. 14 D. C. Driscoll, M. Hanson, C. Kadow, and A. C. Gossard, Appl. Phys. Lett. 78, 1703 2001. 15 D. C. Driscoll, M. P. Hanson, E. Mueller, and A. C. Gossard, J. Cryst. Growth 251, 243 2003. 16 J. M. Zide, D. O. Klenov, S. Stemmer, A. C. Gossard, G. Zeng, J. E. Bowers, D. Vashaee, and A. Shakouri, Appl. Phys. Lett. 87, 112102 2005. 17 W. Kim, J. Zide, A. Gossard, D. Klenov, S. Stemmer, A. Shakouri, and A. Majumdar, Phys. Rev. Lett. 96, 045901 2006. 18 W. Kim et al., Appl. Phys. Lett. 88, 242107 2006. 19 D. G. Cahill, Rev. Sci. Instrum. 61, 802 1990. 20 S. Adachi, Physical Properties of III V Semiconductor Compounds: InP, InAs, GaAs, GaP, InGaAs, and InGaAsP Wiley, New York, 1992. 21 I. Kudman and E. F. Steigmeier, Phys. Rev. 133, A1665 1964. 22 G. Chen and A. Shakouri, ASME Trans. J. Heat Transfer 124, 242 2002.