ABSTRACT Title of dissertation: THE LOW TEMPERATURE MAGNETOCONDUCTANCE OF DELTA-DOPED SILICON Dan Francis Sullivan, Doctor of Philosophy, 2005 Dissertation directed by: Dr. Bruce E. Kane Department of Physics The low temperature electronic transport properties of delta-doped silicon are established here, in an e?ort to determine the technological limitations facing the construction of semiconductor wafers with monolayer doping profiles. Both phosphorous (Si:P) and boron (Si:B) delta-doped silicon wafers prepared by molecular beam epitaxy have been investigated. Devices fabricated from samples of these wafers were cooled to approximately 130 mK in a dilution refrigerator, where four wire resistance measurements were performed in the presence of magnetic fields both perpendicular and parallel to the plane of the devices. The data obtained was interpreted using the theory of weak localization. For Si:P samples, the e?ective thickness of the delta-doped region can be inferred by comparing the magnetoconductance signal in parallel magnetic fields to that of perpendicular fields. Using this technique on several di?erent samples which were annealed at 850 ? C, an estimate of the di?usivity of phosphorus in silicon at this temperature is established. In addition, it is shown that the primary mechanism for the loss of quantum mechanical phase coherence in these samples at low temperatures is the electron-electron interaction, and provide evidence for lattice defects enhancing this dephasing. For Si:B samples, the spin-orbit interaction is shown to dominate the dephasing, and the temperature dependence of this interaction is found to be substantially di?erent from that observed in Si:P devices. Based on these results, I conclude that specific epitaxial growth techniques must be followed if monolayer doping profiles are ever to be achieved in silicon. In addition, because this weak localization technique for measuring the thickness of delta-doped regions is highly sensitive and non-destructive, it may ultimately prove to be the most e?ective method yet established for probing ultra-thin doping profiles. Finally, several theoretical desiderata are established, which would enable a more accurate interpretation of experimental data. THE LOW TEMPERATURE MAGNETOCONDUCTANCE OF DELTA-DOPED SILICON by Dan Francis Sullivan Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2005 Advisory Committee: Professor Frederick C. Wellstood, Chair Doctor Bruce E. Kane Professor Chia-Hung Yang Professor James R. Anderson Adjunct Professor Keith C. Schwab Professor Thomas D. Cohen c?Copyright by Dan Francis Sullivan 2005 To Amanda and Rosalee. ii Acknowledgments It is unfortunately impossible to acknowledge every individual who has contributed in one way or another to this dissertation. Accordingly the people who are mentioned below are those who have figured most prominently in these e?orts. I would first like to thank my dissertation advisor, Dr. Bruce Kane, for supporting me financially through the five plus years I worked for him. I also appreciate him giving me the opportunity to work on this project essentially independently, and for the many conversations we had which helped to deepen my understanding of delta-doping in silicon. Dr. Phillip E. Thompson was a great help in preparing the epitaxially grown delta-doped wafers which were investigated in this dissertation. This research would not have been possible without his assistance, and I am grateful for the service he provided. Professor Richard Webb, one of the pioneers in low temperature experiments involving weak localization, graciously provided extensive experimental consultations. His insights and recommendations have been incorporated throughout this work. Although he was originally slated to be the committee chair for this dissertation, an unfortunate set of circumstances prevented him from being able to do so. His presence is sorely missed, and I wish him the best of luck in his new post at the University of South Carolina. I was blessed to have an outstanding group of coworkers at the Laboratory for iii Physical Sciences (LPS), where this research was carried out. These people include Carlos ?el chingon? Sanchez (whose electrical engineering background proved to be an invaluable resource), the ?nanomechanics boys? Matt LaHaye and Olivier ?Buu-yah? Buu (for numerous political conversations and excursions to the Town Hall), Kenton Brown (for helium loans and the use of his L A T E Xbook in writing this document), Keith Schwab (for numerous conversations), and Luyan Sun. I especially thank Carlos, Matt, and Olivier for making the trek out to Utah for Amanda?s and my wedding. I would also like to acknowledge several members of the sta? and management at LPS for their aid. The skilled machining and design aid provided by J.B. Dottellis helped me to fix a broken dilution refrigerator and rewire it so that my experiments could be performed more e?ciently. Greg Latini and Mark Thornton were always available to help with computer and software related issues, not to mention interesting lunchtime conversations. Geo? Burdge, the director of LPS when I arrived, provided many an adventure on Stampede, and introduced me to Bob Putnam and Better Mousetrap, both much-needed diversions from the graduate school grind. Also assisting me with escape from certain ?imperial entanglements? were Ron Andre, Bernadette Preston, and Marc Manheimer. I appreciate their common sense, loyalty, and unity. Finally, I would like to thank those members of my family who have been a constant source of support to me throughout this process. My Mom deserves special praise, for everything she has done for me in my life, especially teaching me how to work. I also thank my Dad, who introduced me to science through iv mathematics, and who will no doubt be pleased with the continuing evolution of the ?new and improved version? he so often cites. My wife Amanda has truly been phenomenal with her patience and support throughout the long and arduous struggle that is graduate school. It?s finally over, honey. Lastly, and perhaps most of all, I thank my beautiful daughter Rosalee, for being so impossibly perfect. You are the light of my life. v Contents 1 Introduction 1 1.1 Background ................................ 1 1.2 Definition ................................. 2 1.3 ThicknessDetermination......................... 3 1.4 Motivation................................. 7 1.5 Prospectus................................. 9 2Theory 11 2.1 BandStructure .............................. 11 2.1.1 GeneralConsiderationsforSi .................. 12 2.1.2 DensityFunctionalTheory.................... 18 2.1.3 ?-LayerBandStructure...................... 20 2.1.4 Perspective ............................ 36 2.2 WeakLocalization ............................ 40 2.2.1 HistoricalDevelopment...................... 40 2.2.2 Quantum Conductivity . . . . . . . . . . . . . . . . . . . . . . 42 2.2.3 Calculation of ??(H)....................... 47 vi 2.2.4 T Phenomenology ........................ 52 2.3 BerryPhase................................ 55 2.3.1 2DEGBerryPhase........................ 56 3 Fabrication 63 3.1 ?-layerRequirements........................... 63 3.1.1 GeneralConsiderations...................... 64 3.1.2 PreviousWork .......................... 65 3.2 ?-dopedWaferPreparation........................ 72 3.3 DeviceFabrication ............................ 79 3.4 DeviceTesting............................... 87 4Measurement 89 4.1 Background ................................ 89 4.2 The Two-Dimensional Conductivity Tensor . . . . . . . . . . . . . . . 93 4.2.1 HallBarDevices ......................... 95 4.2.2 vanderPauwDevices ...................... 97 4.3 Measurement ............................... 99 4.3.1 Instrumentation..........................101 4.3.2 Grounding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 4.3.3 ExperimentalProcedure .....................109 4.4 MeasurementErrors ...........................120 4.4.1 E ..................................121 4.4.2 H .................................126 vii 4.4.3 J ..................................130 4.4.4 T..................................132 4.4.5 ? ..................................133 5 Results 135 5.1 Si:P ?-Layers ...............................135 5.1.1 Di?usivityMeasurements.....................136 5.1.2 DensityDependence .......................177 5.1.3 FinalRemarksandSummary ..................185 5.2 Si:B ?-layers................................192 5.2.1 SamplePreparation........................193 5.2.2 Theory...............................195 5.2.3 Results...............................196 6 Conclusion 203 6.1 FutureDirections.............................203 6.2 Summary .................................205 6.3 Conclusion.................................206 A Elementary Transport Theory 208 A.1 Classical Conductivity . . . . . . . . . . . . . . . . . . . . . . . . . . 208 A.2 HallE?ect.................................210 A.3 MeanFreePath..............................212 A.4 Mobility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214 viii A.5 DensityofStates .............................215 A.5.1 IsotropicMass...........................216 A.5.2 AnisotropicMass .........................218 B Data Analysis Routines 221 B.1 ??(H ? ) ..................................221 B.1.1 WL1p3...............................222 B.1.2 deltaGWL1p3...........................223 B.1.3 perpfitWL1p3...........................224 B.1.4 deltaLp ..............................225 B.2 vanderPauwTheorem..........................226 B.3 SIMS T ..................................228 C Dilution Refrigerator Modifications 231 C.1 Wiring...................................232 C.2 Repair...................................236 ix Chapter 1 Introduction This introductory chapter provides a comprehensive framework for understanding the data and theoretical analysis presented throughout this dissertation. A brief outline of the material I will cover is included at end of this chapter. 1.1 Background The importance of semiconductor physics in everyday life, and that of silicon, in particular, can hardly be overstated. Since the invention of the transistor by Bardeen, Brattain, and Shockley in 1948, semiconductor technology has steadily advanced, with transistors simultaneously becoming both faster and smaller. These devices have found applications in an endless list of areas, including basic scientific research, personal computers, and all aspects of communication and information technology. Indeed, the myriad of roles the internet now plays in our 1 lives would be impossible without the computational power provided by microprocessors, with billions of transistors integrated onto a single silicon wafer. State-of-the-art transistors now (2005) have dimensions of less than 100 nm, and the trend towards ever smaller devices, with a concurrent doubling of the transistor density on a single wafer every two years (?Moore?s law?), is expected to continue for at least another decade. There are, however, physical limits to this scaling. As the dimensions of these devices become smaller and smaller, the atomic nature of matter can no longer be ignored, and quantum mechanical phenomena, such as charging energies and tunneling events, become manifest. Ultimately, the ability to control doping profiles in semiconductors with atomic precision, will determine how small functional devices, such as transistors, can be built. The technique known as delta-doping of semiconductors represents the frontier in this e?ort [1], and is the subject of research in this dissertation. 1.2 Definition A single delta-doped region in a semiconductor is referred to as a delta-layer (?-layer), and is simply an extremely thin layer of impurity atoms in an otherwise pure crystal. Ideally, the dopants occupy a single atomic layer (monolayer) in the lattice. A schematic representation of this situation is shown in Figure 1.1. Figure 1.1 (a) shows that the impurity atoms (dopants) are confined to a narrow, planar region of the crystal, ?sandwiched? between two layers nominally devoid of 2 dopants. Figure 1.1 (b) shows a plot of the doping density, n(z), as a function of vertical position in the crystal. We see that n(z) = 0 everywhere except at the position z 0 , hence the name ?-layer implies a ?Dirac delta? doping distribution, ?(z -z 0 ). Mathematically, this ideal delta-doping profile n(z) is represented by n(z) = n 2D ?(z?z 0 ) (1.1) where n 2D is the two-dimensional density of atoms in the ?-layer. The relevant consideration for the work in this thesis (?-layers in Si) is that an actual physical ?-layer, composed of real atoms, must possess a finite width, or have a thickness, T , associated with it. Although the physical limit of T is a monolayer, all the samples investigated in this work, and thus far reported in the literature, far exceed this lower bound. Nonetheless, very thin (T?5nm)?-layers have been produced in a variety of semiconductors, and the ultimate limits in preparing these structures are still being explored. 1.3 Thickness Determination One of the central questions I will address in this thesis is to quantify precisely how two-dimensional a given ?-doped sample is (that is, measuring T ). The way I approach this problem is through a weak localization analysis of low temperature, four-wire conductance measurements (?), performed on ?-doped samples in the presence of a magnetic field (H); hence the term magnetoconductance, ?(H). Both perpendicular (?(H ? )) and parallel (?(H bardbl )) magnetic field configurations play an important role in this T measurement, as does as the temperature, T. The 3 (a) (b) Figure 1.1: (a) Schematic representation of a ?-doped wafer, and (b), a plot of the doping density versus vertical position in the crystal for an ideal ?-layer. The doping density is zero everywhere except at the position z 0 . theory of weak localization will be discussed in Chapter 2, but for now it su?ces to say that such a measurement of T amounts to a comparison of the e?ective cross sectional areas subtended by the ?-layer in H ? and H bardbl . To illustrate this idea, consider the right circular cylinder of radius R and height T ,sketchedin Figure 1.2. When viewed from ?above? (corresponding to the H ? configuration), the observed cross sectional area is simply that of a circle: A ? = ?R 2 . If the cylinder is rotated and viewed from the side (corresponding to H bardbl ), the cross section becomes rectangular, with area A bardbl =2RT . Solving these two equations for T gives T =A bardbl radicalbig ?/4A ? . Clearly, an accurate estimate of T requires a similar accuracy in determining the areas A ? and A bardbl . The discussion of weak localization 4 (a) (b) Figure 1.2: A weak localization measurement of the ?-layer thickness,T , corresponds to a comparison of two e?ective cross sectional areas in perpendicular and parallel magnetic fields. In (a) (corresponding to H ? ), the appropriate e?ective area is A ? = ?R 2 . In (b) (corresponding to H bardbl ), the e?ective area is A bardbl =2RT .Fromthese relations one obtains T =A bardbl radicalbig ?/4A ? . in Chapter 2 will make clear how this is done, as well as taking into account some of the details neglected in this simplified example. To see how these considerations are involved in ?(H) measurements, typical weak localization spectra for a phosphorous-doped silicon (Si:P) ?-layer sample are shown in Figure 1.3 (these measurements were taken at the base temperature of the dilution refrigerator used to cool the samples, T similarequal 130 mK). The upper curve corresponds to H ? data, the lower to H bardbl data. The di?erence in the response of the conductance to H between the two configurations is obviously well-resolved. In general, ?(H ? )isalwaysgreaterthan?(H bardbl ), and the more of a contrast that exists between the two configurations, the thinner is the ?-layer in the device. Again, by exploiting the asymmetry between ?(H ? )and?(H bardbl ) and applying the 5 Figure 1.3: Typical weak localization data (points), with theoretical fits (solid lines). The upper curve corresponds to ?(H ? ) data, the lower to ?(H bardbl ) data. The greater the di?erence between ?(H ? )and?(H bardbl ), the thinner is T . appropriate theory to estimate the relevant cross sectional areas, a measurement of T is obtained. Now that the reader has some feel for what will be presented in this dissertation, some justification for carrying out this work will be discussed. 6 1.4 Motivation This dissertation research was originally motivated by an interest in quantum computing (QC), specifically the proposal of my advisor, Bruce Kane, [4]. The idea was to use individual P donors in crystalline Si (c-Si) as quantum bits (?qubits?). The requirements and implications of QC have been elaborated upon extensively elsewhere, and will not be discussed in any detail here, except to say that such an achievement would render certain formidable computations (e.g. factoring large numbers [5]) tractable. Before QC can be achieved, however, many fundamental and practical questions must be addressed. One is how accurately can P donors can be placed in a Si crystal? This issue turns out to be of high importance for QC [6]. Various groups have made e?orts to investigate the issue of precise P donor placement in c-Si through techniques such as scanning tunneling microscope positioning [7] and ion implantation [9], but perhaps the most readily available system for study are ?-layers grown by molecular beam epitaxy (MBE), as used in this thesis. Although other techniques for constructing ?-layers have been demonstrated (such as chemical vapor deposition (CVD) [8]), MBE o?ers many advantages, which will be presented in Chapter 3. QC is far from the only application for ?-layer structures, however. They are widely used in charge coupled devices (CCDs) as highly e?cient detectors for photons and low-energy fluxes in particle physics [10]. Another application is in resonant interband tunnel diodes (RITDs), which dissipate little power and are 7 therefore desirable for use in integrated circuits [12]. A possible future use of ?-layers is in the construction of single electron transistors (SETs), which have a host of potential applications, including use as a current standard in metrology [13] and ultra-sensitive charge detection [14]. Ialsonotethat?-layers are of interest from a purely theoretical perspective. The electronic system associated with a ?-distribution is inherently two-dimensional, and some of the most striking discoveries in solid state research over the past 20 years stem from the physics of associated with two-dimensional electron gases (2DEGs). One of the most prominent is von Klitzing?s discovery in 1980 [15] of the quantum Hall e?ect (QHE) in an Si MOSFET (Metal Oxide Semiconductor Field E?ect Transistor). This e?ect now forms the basis for a resistance standard [16]. Another important discovery was of that the fractional quantum Hall e?ect (FQHE), first observed by St?ormer and Tsui in 1984 [17] and explained theoretically by Laughlin in 1985 [18]. In the FQHE, charge carriers with a fractional electronic charge (e.g. 1/3, 2/3, 1/5, etc.) are observed in very high mobility (typically GaAs) heterostructures at low temperatures and high magnetic fields. ?-layers are a member of this exotic 2DEG family, but display carrier mobilities many orders of magnitude lower than their ballistic transport, high mobility MOSFET and GaAs cousins, and will consequently display rather di?erent physical characteristics. 8 1.5 Prospectus Chapter 2 will present those theoretical considerations most pertinent to this work. Specifically, the band structure of ?-layers in Si will be discussed, along with the impact this structure has on magnetoconductance measurements. Since this thesis hinges on interpreting such data using the theory of weak localization, a systematic treatment of this phenomenon is covered. Lastly, the subtle e?ect of a Berry phase and its influence on interpreting T using weak localization will be explored. Chapter 3 is concerned with the samples used in my experiments. I include a brief discussion of previous e?orts in preparing Si:P ?-layers, the MBE conditions necessary for constructing a ?-doped wafer, and the appropriate methods used in device fabrication. The presentation of the device fabrication process is rather ?hands-on?, and is meant to serve as a reference recipe to be followed in order to produce high quality devices for measurement. Chapter 4 is perhaps the most important of the entire dissertation, and deals with the measurement techniques and equipment which were used to obtain my weak localization data. Since this is an experimental thesis, Chapter 4 will be somewhat more detailed in its exposition than the other chapters, and explains the precise conditions under which my magnetoconductance measurements were performed. In Chapter 5 I present the results and analysis of my experiments, along with a discussion of the various systematic errors involved in the measurements. One 9 set of measurements is concerned with annealing several di?erent devices constructed from the same Si:P ?-doped wafer, extracting T from each device as described in Chapter 2, and using this information to determine the di?usion coe?cient of Si:P at T = 850 ? C. Other experiments involve weak localization T measurements performed on devices constructed from different wafers with varying carrier densities, n. A brief account of measurements on a device constructed from a boron-doped silicon (Si:B) ?-layer is also presented, showing the dramatic influence of spin-orbit scattering has on ?(H). Chapter 6 summarizes my results and suggests new directions for future investigation. The appendices cover ancillary points relevant my work, including a discussion of elementary transport theory, the programming code I used in data analysis, and the modifications and improvements I made to the dilution refrigerator I used. The works of the numerous authors who helped to lay the foundations for this research are listed in the bibliography. 10 Chapter 2 Theory This chapter presents three theoretical aspects crucial to understanding the physics of delta-doped silicon. The first section is on the electronic band structure of a ?-layer system in Si, and how this structure influences the interpretation of magnetoconductance data. The second is an overview of the theory of weak localization and its use in probing ?-layer structures. Lastly, the phenomenon of a Berry or geometric phase is discussed, and I describe how this phase manifested itself in the magnetoconductance measurements I performed. 2.1 Band Structure This section will describe the energy characteristics of electrons in Si, culminating with a theory particular to Si:P ?-layers. 11 2.1.1 General Considerations for Si I will outline the main results for the band structure of Si here to establish specific numbers and notions pertinent to my work. These results are reasonably well-known, and are taken primarily from [19, 20, 21]. The term ?band structure? refers to specifying the energy of an electron (or hole) in a given crystalline potential as a function of either wavevector k, position r,orboth.TheE(k) representation is often referred to as a dispersion relation, and in essence quantifies how the kinetic energy of an electron in a crystal depends on its direction of propagation. Since thorough discussions of this topic are in standard texts on solid state physics [2, 3], I merely summarize here. Qualitatively, when in the presence of a periodic crystalline potential, solutions of the Schr?odinger equation (H? =E?) have a range of energies over which no solutions exist. This is the origin of the electrical characteristics of solids, and, for the case of interest here, produces the energy (or band) gap, E g ,of crystalline Si (c-Si). To appreciate how this comes about, consider an isolated Si atom, with electronic configuration 1s 2 2s 2 2p 6 3s 2 3p 2 (i.e. four valence electrons outside a closed Ne shell). Combining many such Si atoms into their lowest energy state produces a covalently bonded crystal, identical in structure to that of C atoms in diamond. In this crystal, each atom occupies the center of a tetrahedron, with nearest neighbor atoms positioned at the vertices of the tetrahedron, as depicted in Figure 2.1 (a). This is equivalent to a lattice constructed from two face-centered cubic (fcc) lattices (Figure 2.1 (b)), one copy of which is displaced 12 (a) (b) Figure 2.1: Crystal structure of Si. (a) tetrahedral structure, where each Si atom occupies the center of a tetrahedron. (b) fcc lattice. The c-Si lattice consists of two fcc lattices: one at the at the origin, the other 1/4 of the distance along the main body diagonal. from the other 1/4 of the distance along the main cubic diagonal. To provide the appropriate length scale, x-ray crystallography has determined that the lattice constant for c-Si is similarequal 0.543 nm. A generic semiconducting energy band diagram (in position space) is shown schematically in Figure 2.2. The states below the gap comprise the valence band (E v ), those above the gap the conduction band (E c ), and states with energies forbidden by the Schr?odinger equation lie in E g . In order for an electron (hole) to be transported through the crystal under the application of an arbitrarily small electric field E, it must have an energy greater (less) than E c (E v ). Of course, c-Si is an indirect gap semiconductor, meaning that the maximum of E v and the minimum of E c occur at di?erent points of k-space. Therefore the actual E(k) 13 Figure 2.2: One-dimensional semiconductor energy band diagram (energy vs. posi- tion, E(x)). States with energies forbidden by the Schr?odinger equation lie in the energy gap, which for Si at T = 0 is E g =1.17eV.Thek-space diagram [20] is anisotropic,andE g varies with k. band diagram for Si is much more complicated than that of Figure 2.2. Numerous measurements have determined that the maximum of E v occurs at the center of the Brillouin zone (the ??? point), whereas the minimum of E c occurs along the < 100 > directions of k-space, approximately 3/4 of the distance from the center of the Brillouin zone to its boundary (the ??? points). Due to the symmetry of the lattice, there is a six?fold degeneracy associated with the minimum of E c in Si, as depicted in Figure 2.3. The shape of these six conduction band valleys (or pockets) can be approximated by an ellipsoid whose eccentricity is specified by the e?ective mass tensor of the electrons. Cyclotron resonance measurements have established [22] that two e?ective masses are 14 Figure 2.3: Si energy ellipsoids showing the six-fold conduction band degeneracy. The curvature near the ends of the ellipsoids corresponds to the transverse e?ective mass m ? = 0.190m e , while the curvature near the middle of the ellipsoids corre- sponds to the longitudinal e?ective mass, m bardbl = 0.916m e appropriate for characterizing these Si ellipsoids. Taking the center of a single pocket to be the k-space origin, E(k) takes the form E(k)= planckover2pi1 2 k 2 x 2m bardbl + planckover2pi1 2 (k 2 y + k 2 z ) 2m ? (2.1) with analogous expressions holding for the other symmetry-related pockets of Figure 2.3. Since E ? (1/m), a lighter mass corresponds to a more sharply curved portion of the energy ellipsoid. The results of [22] are m bardbl = 0.916m e and m ? = 0.190m e ,wherem e is the free electron mass (m e similarequal 9.11x10 ?31 kg). As the labels suggest, m bardbl corresponds to a k parallel to the major (long) axis of the ellipsoid, while m ? corresponds to k perpendicular to the major axis. Pure Si, being an intrinsic semiconductor, becomes an insulator at su?ciently low temperatures, with E g = 1.17 eV at T = 0. In order to obtain a Si sample 15 which conducts even at T = 0, the crystal must be doped to a level greater than the metal-insulator transition (MIT) for the particular impurity element involved, defining the critical density, n c . This in turn establishes a limit between the two-dimensional dopant density n in a ?-layer and its thickness T .Ifn< n c T ,the sample will not conduct at low temperatures, and cannot be investigated using weak localization techniques. The two Si dopants studied here, P and B, have n c = 3.7x10 18 /cm 3 and 4.0x10 18 /cm 3 , respectively, while their crystalline ionization energies are 0.045 eV below E c for Si:P, and 0.045 eV above E v for Si:B (since B is a hole donor). Because these energies are near E c and E v respectively, and small compared with most other elements, P and B are known as ?shallow donors? in Si. I note here an interesting relationship which exists between n c and the Si:P Bohr radius, a B , which I shall refer to again in Chapter 5. First, at low temperatures it is legitimate to consider an isolated P atom in Si as an ?e?ective Hydrogen atom? [23], since P has one more valence electron than Si, and the remainder of the crystal is insulating. Furthermore, the other Si atoms in the crystal act as a renormalized vacuum with dielectric constant ?.ForSi,? = 11.9 epsilon1 0 ,whereepsilon1 0 is the permittivity of vacuum (in SI units epsilon1 0 = 8.85x10 ?12 C 2 /N?m 2 ). Relating the kinetic energy (KE) to the potential energy (PE) of this system through the virial theorem (i.e. KE = PE/2) [24], the following relationship is obtained for a B : a B = 4??planckover2pi1 2 me 2 . (2.2) Using m ? for m, I obtain a B similarequal 3.23 nm, quite large compared to the Si lattice 16 Figure 2.4: Sketch of the relationship between the conductivity of a sample and doping density. If the doping level is such that the mean donor d spacing is less than 2a B , the ?Bohr diameter?, the electron clouds (open circles) have considerable overlap, and the sample will conduct at T = 0 (i.e. it is metallic). If d > 2a B ,the sample will insulate, and when d = 2a B , the MIT occurs. spacing of 0.543 nm. The use of m ? is only an approximation, and obviously equation (2.2) implies that a lighter electron will on average lie a greater distance from the nucleus. The mean donor spacing at the MIT is given by taking the cube root of n c for Si:P, for which I obtain d similarequal 6.47 nm. Therefore the relationship d = 2a B is nearly satisfied, to which a simple physical interpretation can be ascribed. If each P donor is thought of as a ?conducting ball? of diameter 2a B , then Si:P samples will become insulating when these balls no longer touch. Thus, from an atomic point of view, the better a sample conducts, the greater is the electronic overlap bewtween donors. A depiction of this situation is shown in Figure 2.4. All of the samples studied in this work were metallic. 17 2.1.2 Density Functional Theory The task of calculating the band structure for an arbitrarily doped, three-dimensional Si sample from first principles is, in general, intractable. Nevertheless, accurate numerical techniques have been developed to overcome this problem. For crystalline samples the most important result is Bloch?s theorem, which states that the appropriate electronic wave functions can be expressed as the product of a plane wave and a periodic function, whose spatial period matches that of the crystal: ? k (r)=e ik?r ? n (k,r)(? n (k,r) is the lattice periodic function). The label n is the band index, and simply arranges the bands in order of increasing energy. Thus determining the band structure of a given crystal reduces to finding the functions ? n (k,r). Many techniques have been developed which are in excellent agreement with experimental results for the energy gap, magnitude of carrier e?ective masses, etc., including Hartree-Fock calculations, density functional theory (DFT), and Monte Carlo and Green?s function methods. I present a brief outline of DFT here. The other methods are discussed in [20] and the refernces therein. DFT, first introduced by Hohenburg and Kohn in 1964 [25], asserts that the ground state energy of a many-body (N-electron) wave function can be transformed into N one ?electron equations in which the electrons are subjected to an e?ective local potential, V eff (r). Since the properties of a homogeneous electron gas are well-known over a wide range of densities [26], characteristics of inhomogeneous systems can be calculated by ?integrating over? a given 18 distribution, using the V eff (r) appropriate for a homogeneous gas. This DFT approach is known as the ?local density approximation? (LDA), and requires that the density of electrons per unit volume, n(r), be a su?ciently slowly-varying function of position. Precisely what constitutes ?su?ciently slowly-varying? is a problem-specific question, which is addressed in [25]. Once the potential of interest has been specified, DFT requires the self ?consistent solution of two coupled partial di?erential equations. These are the Schr?odinger equation (H? i =E i ? i ) for the subband wave functions and energies, and Poisson?s equation (? 2 V(r)=-?(r)/?), relating the electrostatic potential V(r) to the charge density ? and dielectric constant ?. To proceed, I first note Hohenberg and Kohn?s result [25] for the ground-state energy of an inhomogeneous, interacting electron gas in a static potential, v(r): E = e integraldisplay v(r)n(r)dr + parenleftbigg e 2 8?? parenrightbiggintegraldisplayintegraldisplay n(r)n(r prime )drdr prime |r?r prime | + G[n]. (2.3) In this equation G[n] is a universal functional of the density, G[n] = T[n] + E xc [n], where T[n] is the kinetic energy and E xc [n] the exchange-correlation energy for a system of non?interacting electrons. In the LDA, E xc [n]= integraltext n(r)epsilon1 xc (n(r))dr, where epsilon1 xc is the exchange-correlation energy per electron,givenn(r). In this LDA limit, the ith electron?s Schr?odinger equation can be cast in the form [27] { ?planckover2pi1 2 ? 2 2m +[?(r)+? xc (n(r))]}? i (r)=E i ? i (r). (2.4) Thus we have for the e?ective potential V eff (r)=?(r)+? xc (n(r)) (i.e. an 19 ?e?ective? Poisson equation). The two contributions to V eff (r)aregivenby ?(r)=ev(r)+ parenleftbigg e 2 4?? parenrightbiggintegraldisplay n(r prime )dr prime |r?r prime | (2.5) and ? xc = d dn (nepsilon1 xc (n(r))). (2.6) Finally, n(r) is given by the sum of the contributions to the density from the individual electrons: n(r)= N summationdisplay i=1 |? i (r)| 2 . (2.7) The pieces are now in place to self-consistently calculate a crystal?s band structure. The typical procedure is that some initial n(r) is assumed, and ? and ? xc are calculated according to equations (2.5) and (2.6). Placing these quantities in the Schr?odinger equation (2.4), the ? i ?s are determined. These wave functions are then used in (2.7) to obtain a new estimate for n(r). By iterating this procedure many times, a solution is (hopefully) converged to rapidly. 2.1.3 ?-Layer Band Structure I now present a calculation for the T = 0 band structure of ?-doped Si:P. The theory associated with this problem is discussed in Chapters 1 and 2 of [1]. A more general numerical calculation for Si:P ?-layers which I shall summarize subsequently has been performed in [31]. The approach I take in this calculation is to assume a planar doping distribution, and find the associated electronic potential energy. I then substitute this potential into the Schr?odinger equation, and solve it for the energy eigenvalues 20 Figure 2.5: Planar charge distribution for modeling delta-doped Si:P. ? is the charge density per unit area. E i and subband wave functions ? i . Finally, I determine the density of states, ?(E), of the system, and use ?(E) to find the Fermi energy, E F , as a function of doping density by demanding charge neutrality. Energy Eigenstates As discussed in [28], the wavefunction for a general 2DEG can be approximated by the product of a plane wave (e ik?x )lyingintheplanez=0,and a subband wavefunction, ? i (z), perpendicular to the plane. This section establishes the nature of the ? i . Referring to Figure 2.5, I assume that a planar charge distribution, with charge density per unit area ?, lies in the plane z = 0. Such a distribution acts as 21 a ?quantum well?, providing a confinement potential for the donor electrons. I further assume that ? is proportional to the doping density n, and take ? = en for the remainder of this calculation (e is the electron charge, e = 1.602x10 ?19 C). A more general approach would be to take ? = ?en, and examine the dependence of the results on ?. The electric field E produced by ? is determined by applying Gauss?s law in the standard ?pillbox? form to this system, which yields a constant field in the +/-?z direction: E = ? 2? ?z. (2.8) Such a constant field leads to an electrostatic potential which is linear in z. Therefore the potential energy, V(z), of an electron a distance +/- z from the planez=0is V(z) = eE|z| = e? 2? |z|. (2.9) The use of the absolute value of z in equation (2.9) reflects the fact that the energy of an electron in this potential is a symmetric function of z: E(z) = E(-z). Given the potential energy defined by (2.9), the appropriate one-dimensional Schr?odinger equation for subband wave function ? i (z) with energy eigenvalue E i becomes parenleftbigg ?planckover2pi1 2 2m z parenrightbigg d 2 ? i dz 2 + parenleftbigg e? 2? parenrightbigg |z|? i = E i ? i . (2.10) To solve equation (2.10), I note that the di?erential operator d 2 /dz 2 - |z| has been studied since the 19th century, and the solutions to d 2 f dz 2 ?|z|f = ??f (2.11) 22 are the so-called Airy functions, Ai(z) and Bi(z), with eigenvalues ? [32]. An obvious physical boundary condition which the ? i (z) must satisfy is that ? i (z) ? 0as|z|??, which excludes the Bi(z) as possible wave function candidates. A second boundary condition is imposed on the z = 0 derivative of the ? i (z) by the triangular symmetry of the potential (2.9). If the eigenvalues of (2.11) are indexed in increasing order, then the even index eigenvalues (0,2,4,...) are given by the negatives of the zeros of the derivative of Ai(z), while the odd index eigenvalues (1,3,5,...) correspond to the zeros of the Ai(z) itself [32]. Similarly, the even or odd parity of a particular ? i (z) is determined by the parity of its index, with the even valued eigenfunctions satisfying d? i /dz = 0 at z = 0. To cast equation (2.10) in the form of (2.11), I rescale z according to x = Az, and substitute back into (2.10) to obtain: d 2 ? i dx 2 ? parenleftbigg em z ? ?planckover2pi1 2 A 3 parenrightbigg |x|? i = ? parenleftbigg 2m z E i A 2 planckover2pi1 2 parenrightbigg ? i . (2.12) Choosing A = (em z ?/?planckover2pi1 2 ) 1/3 , the left hand side of equation (2.12) is identical in form to that of (2.11), and therefore the right hand sides also equal each other. Thus the energy eigenvalues for the subband wave functions are given by E i = ? i parenleftbigg e?planckover2pi1 ? ? 8m z parenrightbigg 2/3 = ? i parenleftbigg e 2 planckover2pi1n ? ? 8m z parenrightbigg 2/3 . (2.13) Since electrons in Si have two possible values for m z (m z =m ? and m z =m bardbl ), a discrete spectrum of states exist for each of these masses. The complete energy spectrum for this system is shown plotted in Figure 2.6. In Table 2.1, the 10 lowest eigenvalues of equation (2.11) are listed [29], as well as the associated (rescaled) subband energies for either mass. Since E ? n 2/3 , all of the energies in 23 Figure 2.6: Energy spectrum of states corresponding to m ? = 0.19m e and m bardbl = 0.916m e masses for Si. The energy of the various states have been rescaled such that E 0 =1,whereE 0 is the lowest available energy of the system (corresponding to ? = 1.0187 and m = m bardbl in equation (2.13)). Figure 2.6 and Table 2.1 have been scaled relative to E 0 , the lowest energy subband (i.e. E 0 ? 1). From equation (2.13), it is clear that E 0 has m z =m bardbl , ? = 1.0187, while the next lowest subband has m z =m ? , ? = 1.0187, the next lowest m z =m bardbl , ? = 2.3381, etc. A plot of the probability density |? i (z)| 2 for i = 0,1,2,3 is shown in Figure 2.7. Each ? i (z) satisfies integraltext ? ?? |? 2 i |(z)dz = 1, and for each increment in the index, an additional ?lobe? appears in the wave function profile. The boundary conditions require that the origin be shifted for each ? i (z), i.e. ? 0 (z) ? Ai(z-? 0 ), ? 1 (z) ? Ai(z-? 1 ), ? 2 (z) ? Ai(z-? 2 ), etc [32]. 24 Table 2.1: Table showing eigenvalues of equation (2.11) and corresponding (scaled) energy states for Si masses. index ? E(m bardbl ) E(m ? ) 0 1.0187 1.00 1.68 1 2.3381 2.30 3.86 2 3.2482 3.19 5.36 3 4.0879 4.01 6.75 4 4.8201 4.73 7.95 5 5.5206 5.42 9.11 6 6.1633 6.05 10.17 7 6.7867 6.66 11.20 8 7.3722 7.24 12.16 9 7.9441 7.80 13.11 25 Figure 2.7: Probability densities |? i (z)| 2 of the subband wavefunctions in ?-doped Si:P. As the band index increases, the wave functions become more extended. The ? i have been o?set from each other vertically. 26 Density of States In this section I calculate the density of states, ?(E), for an Si:P ?-doped system. The motivation here is to understand the interplay between the three-dimensional Si energy ellipsoids of Figure 2.3, and the two-dimensional doping plane of Figure 2.5. Appendix A discusses the physical meaning of ?(E) (which has dimensions of 1/(energy)(area)), and calculates ?(E) for isotropic (circular dispersion relation) and anisotropic (elliptical dispersion relation) two-dimensional electron systems. As shown in Appendix A, ? 1 (E) for an isotropic, two-dimensional, single subband electron system, with e?ective mass m ? , is a constant, given by ? 1 (E)= m ? ?planckover2pi1 2 . (2.14) For an elliptical dispersion relation in two dimensions, two masses, m 1 and m 2 ,are required to specify ? 2 (E): ? 2 (E)= ? m 1 m 2 ?planckover2pi1 2 . (2.15) The two-fold spin degeneracy of the electrons is included in both (2.14) and (2.15). Figure 2.8 shows why the two relations (2.14) and (2.15) are necessary for modeling Si:P band structure. This figure is essentially the result of projecting the ellipsoids of Figure 2.3 onto the doping plane of Figure 2.5. The cross section of the ellipsoids perpendicular to the plane is circular, corresponding to ? 1 (E), while the cross section of the in-plane ellipsoids is elliptical, corresponding to ? 2 (E). Combining these two contributons, and ignoring, for the moment, the quantized 27 Figure 2.8: The two-dimensional density of states for Si (100). The energy ellipsoids of Figure 2.3 have been projected into the k z = 0 plane, superimposing Si band structure on the doping distribution of Figure 2.5. Since equation (2.10) is concerned with motion perpendicular to this plane, the m bardbl subbands correspond to the circular area(s), the m ? subbands to the ellipses. 28 nature of the energy states in this system, ?(E) for ?-doped Si:P is given by: ?(E)= 2m ? ?planckover2pi1 2 + 4 ? m ? m bardbl ?planckover2pi1 2 = ? bardbl + ? ? . (2.16) As suggested by the second equality, the first term in (2.16) corresponds to electrons with an m bardbl mass, the second to the m ? mass. The prefactors of each reflect the degeneracies of the ellipses in Figure 2.8. I remark here that a counterintuitive distinction exists between the transport mass (m t )andthedensity of states mass (m ? ). For m t =m bardbl ,m ? =m ? , while for m t =m ? ,m ? = ? m ? m bardbl . To arrive at the final expression I shall use for ?(E) in Si:P ?-layers, I now take into account the discrete nature of the energy states calculated in the previous section. By plotting the energies in Table 2.1 in ascending order on a line (i.e. turning Figure 2.6 ?on it?s side?), it is (hopefully) clear that a given subband becomes occupied only if the most energetic electrons in the system (i.e. those at the Fermi energy, E F ) exceed the subband?s energy eigenvalue. The lowest energy subband (E 0 in Table 2.1) corresponds to m = m bardbl , the second lowest (E 1 = 1.68) to m = m ? , the next two lowest (E 2 =2.30andE 3 = 3.19) to m = m bardbl ,etc.Thus for E 0 ? E F ? E 1 , only a single subband is occupied, and ?(E) = ? bardbl .ForE 1 ? E F ? E 2 , two subbands are occupied, one for each mass, and ?(E) = ? bardbl + ? ? .ForE 2 ? E F ? E 3 , three subbands are occupied (two for ? bardbl ,onefor? ? )and?(E) = 2? bardbl + ? ? . Continuing in this way, one can build up the ?staircase? of the density of states for Si:P ?-layers, shown in Figure 2.9 (the vertical axis is scaled such that ? bardbl = 1). Comparing the two terms in equation (2.16), we see that for ? bardbl =1,? ? = 4.39, and consequently it is easy to determine which steps in Figure 2.9 correspond 29 Figure 2.9: Staircase density of states ?(E) for ?-doped Si:P. The short vertical steps correspond to states with m = m bardbl , large vertical steps correspond to states with m =m ? . 30 to m ? or m bardbl states. The states with short (unit) steps correspond to m bardbl ,those with large steps to m ? states. With the step-like nature of ?(E) in Si:P ?-layers established, the final form I shall use in this dissertation is: ?(E)= summationdisplay i ? i ?(E F ?E i ). (2.17) Here ?(x) is the Heaviside step function, defined as ?(x) = 0 for x < 0and?(x) =1forx? 0. ? i is either ? bardbl or ? ? , depending on whether E i corresponds to an m bardbl or an m ? state. Fermi Energy The final aspect of this calculation will be to determine the Fermi energy, E F , of an Si:P ?-layer for arbitrary doping density, n. For a single subband system, the following relation holds between n and ?(E) [2, 3]: n = integraldisplay E max E min ?(E)f(E)dE. (2.18) f(E) is the Fermi-Dirac distribution function. The integral (2.18) gives the number of electrons per unit area with energies between E min (the lowest allowed energy in the band) and E max . Since I am considering T = 0 in this computation, f(E) = 1 for E ? E F , and f(E) = 0 for E > E F . Furthermore, having seen that ?(E) is constant for a 2DEG with a single subband, (2.18) becomes n = ?(E max ?E min )=?(E F ?E min ) (2.19) since the integral is cut-o? for E max > E F . 31 I now use (2.19) to determine the electron density at which the first subband is fully occupied. By setting ? = ? bardbl ,E min =E 0 and E F =E 1 in (2.19), and using equation (2.13) for the energies, I find that the first subband is filled when n = 3.39x10 12 /cm 2 . Thus for densities greater than this, at least two subbands will be occupied. All of the samples I investigated in this work had densities greater than this value. An analogous computation can be carried out for the second subband (i.e. for what n are the first two subbands filled?). In this case, n is given by: n = ? bardbl (E 2 ?E 0 )+? ? (E 2 ?E 1 ). (2.20) Substituting in the appropriate quantities I obtain n = 6.46x10 14 /cm 2 ,which exceeds the highest density sample I investigated (n similarequal 1.5x10 14 /cm 2 ) by roughly a factor of 4. Thus I conclude that all of the samples I measured in this work had two subbands occupied: one with m z =m bardbl , the other with m z =m ? . A relevant consideration here are the physical limits imposed on n. Obviously, n can be made arbitrarily small (zero doping limit), but in order to investigate a sample using weak localization it must have n ? n c = 2.39x10 12 /cm 2 (corresponding to the two-dimensional MIT for Si:P). There is also an upper limit to n: it can be no greater a monolayer Si (100) surface (1 ML = 6.78x10 14 /cm 2 ). In other words, for n > 1 ML, the doping distribution is no longer ??-like? since the impurity atoms must begin to inhabit other lattice planes. Therefore, for truly planar Si:P doping distributions, a maximum of three subbands are occupied at T =0. 32 The results of this calculation for 10 11 /cm 2 ? n ? 10 15 /cm 2 are displayed in Figure 2.10. Plotted are the two lowest energy subbands for the m bardbl and m ? masses, E F , and the densities of the samples I have measured (labeled A, B, C, D). E F was found by specifying n, calculating the necessary E i from (2.13), substituting these into (2.19) or (2.20), and then solving the resulting expression for E F . This ensures that the system obeys charge neutrality. Also labeled are the densities corresponding to the MIT and a ML doping level. According to this plot, all of the samples I measured had two subbands occupied. Calculational Limitations To illustrate some theoretical limitations of my calculational approach, I now compare my results for the band structure of ?-doped Si:P to those reported by Qian et al. in [31]. First, it should be clear that the potential energy approximation I used (2.9) cannot be correct, since V(z) increases indefinitely with |z|. In reality, if one assumes charge neutrality, the electric field of the donor nuclei must eventually be ?screened? by the conduction electrons. In other words, su?ciently far from the ?-layer, V(z) should be flat, corresponding to zero electric field. Accounting for this ?band-bending? (and other factors) is done in [31]. These authors use a more general Schr?odinger equation, approximating the ?-doped system by [H 0 + V dop + V H + V xc ]?(r)=E?(r) (2.21) Here H 0 is the potential associated with a 1000 ML Si lattice, V dop is the potential 33 Figure 2.10: Results of calculation. E 0 and E 2 correspond to the two lowest energy subbands with with m z =m bardbl ,E 1 and E 3 to the lowest energy subbands with m z =m ? . The dashed line represents the Fermi energy, E F .TheE 1 subband starts to become occupied for a density n > 3.39x10 12 /cm 2 ,theE 2 subband for n > 6.64x10 14 /cm 2 . For a 1 ML doping level (n = 6.78x10 14 /cm 2 ), three subbands are occupied (two with m z =m bardbl ,onewithm z =m ? ). The approximate MIT density for ?-doped Si:P is shown, as well as the densities of the four samples measured in this work (A,B,C,D), discussed in Chapter 5. All of the samples measured in this work had two subbands populated: one with m z =m ? , the other with m z =m bardbl . 34 due to the P dopants, V H is the Hartree potential due to the electron-electron interaction, and V xc the exchange-correlation potential of the electron gas. They assume a planar distribution of donors, in which the P atoms lie in a c(4x2) Si (100) surface reconstruction (see Chapter 5 of [20]). They ignore both spin-orbit scattering (which is weak in Si:P [30]), and ?inter-pocket? scattering between the various energy ellipsoids in Figure 2.3. The band structure for this system is solved by by expanding ?(r) in a planar Wannier orbital (PWO) basis, which transforms the di?erential equation (2.21) into a finite-dimensional eigenvalue problem, enabling the appropriate eigenvalues and eigenvectors to be found. Since a PWO is the product of a two-dimensional plane wave and a one-dimensional Gaussian function, products of such PWOs have convenient computational properties. Two classes of subbands result, labeled the ??? and ??? bands. Two ? bands are found two exist, lying approximately 20 meV apart near the ? point of the Brillouin zone, which correspond to the circles of Figure 2.8. The four equivalent ? bands correspond to the in-plane ellipsoids of Figure 2.8. The dispersion relations for these subbands are E i (k)=E (?) i + k 2 m ? ? (2.22) and E i (k)=E (?) i + (k x ?k ? ) 2 m ? bardbl + (k y ) 2 m ? ? . (2.23) The e?ective masses are determined to be m ? ? = 0.211m e and m ? bardbl = 0.95m e ,close to the values of m ? and m bardbl for ?normal? Si. Should the need arise, I will use these masses in subsequent calculations, and I define m ? t ,thetransport e?ective mass 35 [34], by m ? t =2m ? ? m ? bardbl /(m ? ? +m ? bardbl ) similarequal 3.15x10 ?31 kg. Although [31] provides a more general calculation than I have performed, the results of both approaches are compatible. As an experimentalist, this is heartening, since it suggests a strong understanding of the fundamental physics governing ?-doped Si:P. 2.1.4 Perspective The significance of subband structure in the context of 4-wire measurements can be appreciated by the fact that such a measurement actually sums the contribution of each subband to the total conductance signal. For example, when measuring the Hall resistance to obtain the carrier density n (see Appendix A), this quantity can be considered as the sum of the carrier densities n i in each individual subband. Therefore, if there are M independent subbands populated in a given device, n is given by n= M summationdisplay i=1 n i . (2.24) Similar considerations apply to the measured conductance ?, assuming independent (decoupled) subbands. Labeling the ith subband?s contribution by ? i , and using the Drude expression for each component (see Appendix A), we arrive at ? = M summationdisplay i=1 ? i = M summationdisplay i=1 parenleftbigg n i e 2 ? i m i parenrightbigg . (2.25) Since I used ?(H) to probe the thickness T of ?-layers (see Chapter 5), it is important to understand the contribution of each ? i to T . In the calculation I performed above, a perfectly planar distribution was assumed. However, Figure 36 2.7 clearly shows that even for a perfectly flat ?-layer, the spatial extent of the electronic distribution depends on which subband is considered. To establish a consistent definition for T , I first define the thickness of the ith subband, T i ,in terms of its mean-square variance in the ?z direction: T i =2 radicalBig < z 2 i > ? < z i > 2 (2.26) where = integraltext ? ?? f(z)? ? i (z)? i (z)dz. Assuming only the i = 0 (lowest energy) subband is occupied, this reduces to [1] T =2 radicalbigg 7 5 parenleftbigg 4planckover2pi1 2 18emE parenrightbigg 1/3 , (2.27) and evaluating this with ?prototypical? device parameters (n = 10 14 /cm 2 ,m=m ? t = 3.15x10 ?31 kg), I obtain Tsimilarequal1.1 nm. Since E?n, as in (2.27), in general for this approximation T?n ?1/3 . Thus even for a single subband and a perfectly planar distribution, di?erent 2DEG thicknesses are obtained depending on the doping density n. I note that although it is completely reasonable from a physical perspective that T of a 2DEG should vary with n, the assumed thickness of the donor distribution does not change with doping density. In fact, the donors in my calculation as well as [31] have been assumed to possess zero spatial extent; only the electrons are ?spread out?. I will neglect this e?ect in what follows, and assume equality between the thickness of the 2DEG, as inferred from ?(H) measurements, and the donor distribution itself. This is reasonable if the donors are distributed in a non-planar fashion, which is the case for all of the samples I examined. 37 From equations (2.24) and (2.25), one might be tempted to define T as T = summationtext T i . However, given this definition, higher energy subbands, although generally less populated, would be weighted equally in the sum as the lower energy bands, whence T would be artificially enhanced. A more appropriate definition should take into account the fact that the T i do not contribute uniformly to T : they should be weighted by the subband population. Given this consideration, we are finally in a position to give a reasonable definition of T . Using the fact that each subband contributes to n through (2.24), ? through (2.25), and T through (2.26) I define the ?total 2DEG thickness signal? T by T = M summationdisplay i=1 parenleftBig n i n parenrightBig T i . (2.28) Since this is an experimental dissertation, it would have been preferable to determine the subband populations n i through an actual measurement, instead of resorting to approximate theoretical calculations. Unfortunately, being able to resolve the contribution of each subband to the total conductance ? requires capabilities beyond those discussed in this thesis. For example, in order to infer subband occupations due to Shubnikov de-Haas oscillations in the magnetoresistance [2, 3], one needs to apply a field |H| of su?cient strength to render the cyclotron frequency ? c at least as large as the elastic scattering rate 1/? e .Since? c =e|H|/m and 1/? e =v F /lscript e , this puts a lower bound on |H| of |H| ? planckover2pi1k F /elscript e . For our prototypical device n similarequal 10 14 /cm 2 and lscript e similarequal 25 nm, which yields |H|similarequal65 T. This field is over 5 times stronger than the maximum H (12 T) the cryomagnetic system used in these measurements could attain at 4.2 K. 38 Accordingly the subband occupations could not be experimentally determined. To conclude this section, I now clarify the model for the subband structure I will use for the remainder of this dissertation. What has been omitted from the work [1, 31] I have discussed are any physical e?ects which could lead to a broadening of the ?-layer band structure. This is significant, since if the energy levels of two subbands were broadened to the extent that their distributions overlaped considerably, they would then behave as as one effective subband. Some of the physical processes which could lead to subband broadening include inter-subband scattering, non-zero temperature, defects, a finite thickness of the ?-layer, fluctuations of T in a given sample, impurties, and surface e?ects. All of these play a role in ?randomizing? the band structure of the ?-layer system. The issue of inter-subband scattering has been discussed theoretically in [33, 56], and experimentally in [34] (the measurements of [34] were on Si:Sb doped ?-layers). The results of [34] were that analysis with single subband models proved su?cient for modeling their data. Since my samples and those of [34] were both low mobility, ?-doped Si samples (? similarequal 100 cm 2 /V?s), it seems reasonable to conclude that a similar approach applies here. In fact, single subband models were used for the data which will be dicussed in Chapter 5, and very good theoretical fits to the data were obtained, which experimentally justifies this approach. Thus I shall henceforth assume that a single, e?ective subband is responsible for the transport characteristics observed in this work. With this single subband model in mind, I now discuss the theory of weak localization, and the application of this theory to ?-layer systems. 39 Figure 2.11: Longitudinal resistance of a ?-doped sample at T similarequal 130 mK with a perpendicular magnetic field H applied. The sample displays a negative magnetoresistance, characteristic of the phenomenon of weak localization. 2.2 Weak Localization 2.2.1 Historical Development Before stating any theoretical results, let us first examine magnetoresistance data from an Si:P ?-doped device taken in the presence of a perpendicular magnetic field H at T similarequal 130 mK, shown in Figure 2.11. We see that the resistance of the sample displays a maximum at H =0,and 40 decreases monotonically for |H| negationslash= 0. Such negative magnetoresistance had been observed in two-dimensional systems (such as depletion layers in MOSFETs) long before it was given a solid theoretical interpretation in the early 1980s. The questions I will address here are what causes this negative magnetoresistance to occur,andhowcanitbeusedtoprobea?-layer? Before doing so, however, a brief digression into the historical evolution of weak localization is appropriate. The modern theory of weak localization starts with Anderson?s observation [35] that a conductor which is su?ciently disordered will in fact insulate at T = 0. Subsequent diagrammatic calculations carried out by Langer and Neal [36] showed that the so-called ?maximally crossed? diagrams which contribute to the T = 0 conductivity are divergent, and as we shall see, these are precsiely the diagrams used for carrying out computations of the weak localization correction to the conductivity. The scaling theory of localization, first espoused by Thouless [37], furthered our understanding of localization behavior by subdividing a conductor into phase coherent units, and investigating the e?ect boundary conditions between these units had on the macroscopic conductance (analogous to the overlap viewpoint of Figure 2.4). This led to the theoretical discoverybyAbrahamset al. in 1979 that all two-dimensional systems insulate at T = 0 [38]. The importance of the electron-electron interaction in understanding the behavior of disordered conductors was elaborated upon by many workers in the early 1980?s, including seminal theoretical contributions by Altshuler [39] and Lee [40], among others, while experiments validating these theories have been demonstrated in numerous experiments, beginning with Dolan and Osherho? [41]. Thorough presentations of 41 the theory of weak localization with references to the original literature can be found in the review articles by Bergmann [43] and Chakravarty and Schmid [44], as well as in [39, 40]. A more recent treatment can be found in the paper by Aleiner, et al. [45]. 2.2.2 Quantum Conductivity Often referred to as the ?quantum correction? to the classical conductivity, the weak localization data shown in Figure 2.11 can be expressed as ?(H)=? 0 + ??(H) (2.29) where ? 0 is the Drude (H = 0) conductivity and ??(H) is the weak localization (H negationslash= 0) contribution to the magnetoconductance. Here ?(H) = 1/R a50 (H), and R a50 is the ohms/square resistivity of the device (see Chapter 4). This section will explore the nature of ??(H), including its quantum mechanical origin, magnitude and sign, etc. As it turns out, the data of Figure 2.11, and therefore ??(H), can be understood by invoking a particular length scale into the problem. In addition to the mean free path, lscript e (Appendix A), the other length scale which can be inferred from such data is the dephasing length, lscript ? . The physics behind lscript ? is in some sense much more involved than that of lscript e , but is nevertheless conceptually easy to understand. Whereas lscript e is the average distance an electron can travel in a solid before undergoing an elastic (energy-conserving) collision, lscript ? is the corresponding distance for an inelastic (nonconservative) collison. Since the time evolution of a 42 quantum state with energy E is proportional to the phase e ?iEt/planckover2pi1 , any change in the state?s energy will alter the evolution of the phase of the wave function. Given the fact that quantum mechanical probability distributions are extremely sensitive to phase relationships between state kets (i.e. matrix elements), measureable quantities including magnetoconductance are strongly influenced by these same phase correlations. The phase-breaking mechanisms in solids are the electron-electron, electron-phonon and spin-flip interactions [44]. As we shall see in Chapter 5, the electron-electron interaction is the most important for Si:P ?-layer physics. To proceed with our explanation of Figure 2.11, consider Figure 2.12 (a), which shows a particle (electron) located at the initial position r A at time t = 0. I now ask, what is the probability that it will be found at position r B at some later time t? If we label the probability amplitude of an arbitrary path satisfying the specified conditions by A ? , elementary quantum mechanics dictates that the desired probability, P A?B , is given by the following expression [47]: P A?B = parenleftBigg summationdisplay ? A ? ? parenrightBiggparenleftBigg summationdisplay ? A ? parenrightBigg (2.30) i.e. the squared-modulus of the sum over all appropriate amplitudes A ? . Carrying out the prescribed computation, I obtain P A?B = summationdisplay ? |A ? | 2 + summationdisplay ?negationslash=? A ? ? A ? . (2.31) The first sum represents the classical probability of propagating from r A to r B along the various paths ?. The second sum represents the quantum mechanical interference between arbitrary pairs of paths, labeled by ? and ?.Formostpairs 43 (a) (b) Figure 2.12: Paths from A to B. (a) The probability of propagating from A to B is given by summing the quantum mechanical amplitudes (equation (2.30)) for all appropriate trajectories. (b) When A = B, there is an enhanced probability for a particle to return to A (remain localized), and the resistance is maximal at H = 0 (?coherent backscattering?). For H negationslash= 0, a phase shift proportional to the flux through the loop is induced between oppositley directed paths, and the constructive interference is reduced, making the sample less resistive. 44 of paths, the phase accumulated along the separate trajectories is incoherent (oscillating near a given point r B on the order of the Fermi wavelength, ? F ); thus the interference terms typically average to zero. However, if r A = r B ,thatis,we consider the closed trajectories of Figure 2.12 (b), the situation changes dramatically. Now each closed loop can be paired with a ?time-reversed? counterpart, and whether a given loop is traversed clockwise (?forward in time?) or counterclockwise (?backward in time?) has no influence on the phase accumulation around the path. To explicitly illustrate this, let A 1 and A 2 represent the amplitudes for propagation around a given loop, corresponding respectively to clockwise and counterclockwise trajectories. Denoting the contribution of these two paths to P A?B by p,wehave p = |A 1 + A 2 | 2 =(A 1 + A 2 )(A 1 + A 2 ) ? =2|A 1 | 2 +2Re(A 1 A 2 )=4|A 1 | 2 . (2.32) Ihaveused|A 1 | 2 = |A 2 | 2 in (2.32) due to the fact these trajectories are identical except for the direction of traversal. Note that if the contribution of the interference terms A 1 A ? 2 and A ? 1 A 2 to p had been neglected, we would just have the classical probability p =2|A 1 | 2 , one-half of that predicted by (2.32). Thus we see that these time-reversed paths are coherent (in phase), interfere constructively, and contribute to a higher probability of returning to r A after a time t. These closed loop paths, collectively referred to as ?Cooperons?, are the source of the weak localization correction to the conductivity. Suppose now that a uniform magnetic field H is ?turned on? in the vicinity of these paths, and identify A as the vector potential corresponding to H in the 45 usual way [46] (i.e. H = ??A). The presence of A breaks the time-reversal symmetry of the problem, and causes the wavefunction of a charged particle to acquire an additional phase ?? as it propagates around a loop, given by [47] ?? = parenleftBig e planckover2pi1 parenrightBig contintegraldisplay A?dl (2.33) where dl is an infinitesimal length element tangent to the path. Since we are considering here closed paths, Stokes? theorem ( contintegraltext A?dl = contintegraltext (??A)?da = contintegraltext H?da) allows us to rewrite this expression for ?? as ?? = parenleftBig e planckover2pi1 parenrightBig ? (2.34) where ? is the magnetic flux enclosed by the path. This path-dependent phase accumulated by a charged particle in a magnetic field shows that, ultimately, the phenomenon of weak localization is a particular manifestation of the Aharonov-Bohm e?ect [48]. Accordingly, whether the charge traverses the loop clockwise or counterclockwise a?ects the sign of the phase acquired, and the amplitudes A i in (2.30) must be transformed according to A i ? e i?? A i ,as required by gauge invariance. Taking this into account an re-computing p,Iobtain p =2|A 1 | 2 (1 + cos(2??)). (2.35) Thus, ?H negationslash=0, the constructive interference which contributed to the ?enhanced backscattering? at H = 0 is lost, and it becomes less probable for the charge to return to its initial position. Hence the application of H renders the electrons in the sample ?less localized?, which makes the device less resistive, and we are finally able to appreciate the physical origin of the negative magnetoresistance 46 signal shown in Figure 2.11. In addition, the symmetry of the signal about H =0 can be understood by the ? cos(2??) dependence of p on H. 2.2.3 Calculation of ??(H) Although a start-to-finish calculation of ??(H) is too lengthy to be shown here, a brief sketch of the standard computational procedure is nonetheless appropriate, in order to appreciate the origin of the expressions which will be used for fitting data. The references cited earlier can be consulted for details. As was shown in the previous section, Cooperons are responsible for the resistance maximum at H =0.Thustocalculate?(H) for a particular system, one must merely add up (integrate) the contribution of each Cooperon to obtain the total conductance signal. Several equivalent approaches include calculating ?(H) using the Kubo-Greenwood formula [49], as done in [43], employing Abrikosov?s result [50] for the residual resistance of a normal metal, as in [39], or through the use of linear response theory [51]. Since weak localization stems from the interference of electron waves as they propagate through an impurity distribution, it should come as no surprise that all of these methods employ impurity averaged Green?s functions [50]. The interference between the time reversed paths which comprise the Cooperon can be conveniently represented using Feynman diagrams (a pedagogical introduction to the use of Feynman diagrams in many-body problems can be found in [52]). A representation of the appropriate diagrams for use in weak localization 47 calculations in k-space are shown in Figure 2.13 (a). Here a solid line corresponds to a single-particle propagator (Green?s function), a wavy line to the amplitude for scattering o? of an impurity, and the impurity itself by an open circle. Note that the order of impurity scattering between the upper propagator line (?forward in time?) and lower line (?backwards in time?) is reversed, as one would expect for the previously explained origin for the interference manifested by the Cooperon. Although only two scattering events are depicted here, the term ?maximally crossed? diagrams is apparent from the figure, and simply refers to switching the order of the scattering sequence. If the lower propagator line is reversed, taking care to preserve the scattering sequence, the diagram is ?uncrossed?, and the interference between the two paths can be expressed by the ?ladder diagrams? of Figure 2.13 (b). Taking into account all possible scattering sequences by summing the obvious geometric series suggested by Figure 2.13 (b), an integral equation is obtained for the Cooperon, commonly referred to as the Dyson equation.The diagrammatic representation for Dyson?s equation is shown in Figure 2.13 (c). A few final comments about calculations of ??(H) will now be made. First, it is important to note that not all closed paths in the sample contribute to the weak localization signal: there is a cut-o? parameter, namely lscript ? ,which defines the average size of the phase coherent region in the sample. Furthermore, lscript ? is temperature dependent. In general, this dependence is logarithmic on the temperature, raised to a power p: lscript ? ? ln(T ?p ) [40]. In other words, lscript ? should diverge as T ? 0. This can be understood by the fact that the phase-breaking rate corresponding to the electron-electron interaction varies as 1/? e?e similarequal T 2 /E F , 48 Figure 2.13: Feynman diagrams in k-space for calculating ??(H). In (a), the inter- ference between ?forward? in time electrons (upper lines), and ?backward? in time electrons (lower lines) is accounted for by reversing the scattering sequence under consideration, hence the term ?maximally crossed? diagrams. This is equivalent to the ?ladder diagram? in (b), where the interference is between two forward in time paths, but with k vectors switched. Summing all such diagrams leads to the Dyson equation, represented by (c). See [52] for a discussion of Feynman diagrams in the many-body problem. 49 while that of the electron-phonon interaction varies as 1/? e?ph similarequal T 3 /? 2 D [45] (E F is the Fermi energy, ? D the Debye frequency). For c-Si, ? D similarequal 1 THz [20], whence it is clear that 1/? e?e will dominate at low temperatures. Second, since the phase accumulated around a given path is extremely sensitive to magnetic fields, the device under consideration must be free of magnetic impurities. In addition, if any spin-orbit e?ects are present in the sample, this will also cause a path-dependent phase shift. In fact, the spin-orbit interaction is quite strong in Si:B ?-doped samples, and as we shall see in Chapter 5, produces very di?erent magnetoresistance behavior than that of Figure 2.11. Finally, the e?ective dimensionality of the sample is important, and di?erent corrections apply depending on the geometric properties assumed in the calculation. All of the samples in this work are considered to be e?ectively two-dimensional, but one-dimensional weak localization has also been extensively studied experimentally. Having said all this, it should be apparent that a great deal of theory has been developed to produce useful expressions for fitting experimental data obtained from 2DEGs. Here I simply state that when placed in a perpendicular magnetic field (H ? ), the prediction for the measured change in conductance of a 2DEG system is given by [53]: ??(H ? )=( ?e 2 2? 2 planckover2pi1 ) bracketleftbigg ? parenleftbigg 1 2 + H 1 H ? parenrightbigg ? parenleftbigg 3 2 parenrightbigg ? parenleftbigg 1 2 + H 2 H ? parenrightbigg ? parenleftbigg 1 2 parenrightbigg ? parenleftbigg 1 2 + H 3 H ? parenrightbiggbracketrightbigg (2.36) where ?(x) is the digamma function, the ?logarithmic derivative? of the gamma 50 function. ?(x)= d dx ln(?(x)), and asymptotically approaches the natural logarithm ln(x) for large x. The characteristic fields H 1,2,3 are associated with the transport relaxation mechanisms of impurity potential scattering (H 0 ), inelastic scattering (H i , including electron-electron and electron-phonon interactions), spin scattering (H s ), and spin-orbit scattering (H so ), as follows [43, 53]: H 1 = H 0 + H so + H s (2.37) H 2 = 4 3 H 0 + 2 3 H s + H i H 3 =2H s + H i . In turn, a relaxation time, ? i , can be associated with the H i through H i ? i = planckover2pi1 4eD (2.38) where D is the electronic di?usivity. In two dimensions, D = v 2 F ? e /2. Although in principle knowledge of the carrier e?ective mass would allow one to estimate D fairly accurately, the best way of independently determining D is through measurements at high frequencies [57]. Since this was not done here, I will henceforth equate an e?ective length with H i and ? i via lscript i = radicalbig D? i . (2.39) Fortunately, equation (2.36) can be simplified considerably in the case of Si:P ?-layers, as previous work has shown that both spin and spin-orbit scattering in Si:P are weak [58]. Therefore, the relaxation times associated with these processes are very large, and can be neglected when compared with the electron-electron and 51 electron-phonon interactions. Taking this into account, I arrive at the final expression I will use for modeling ??(H ? ) in Si:P ?-layers: ??(H ? )=( e 2 2? 2 planckover2pi1 )[?( 1 2 + planckover2pi1 4eH ? lscript ? 2 )??( 1 2 + planckover2pi1 2eH ? lscript e 2 )+ln( 2lscript ? 2 lscripte 2 )] (2.40) where lscript ? is the previously defined dephasing length (lscript ? = radicalbig D? ? ), and lscript e the elastic mean free path. If instead the magnetic field is applied parallel to the 2DEG (H bardbl ), the appropriate expression takes the form [54, 55, 56] ??(H bardbl )=( e 2 2? 2 planckover2pi1 )ln(1+?H 2 bardbl ). (2.41) The physical interpretation of the parameters lscript ? and ? in the context of measuring T will become apparent presently. As discussed in Chapter 1, using weak localization to measure T corresponds to a cross section comparison of the 2DEG in H ? and H bardbl configurations. Now that I have discussed the physical origin of the negative magnetoresistance in Figure 2.11 (namely magnetic field dependent quantum interference e?ects), I present a phenomenological approach for determining T . 2.2.4 T Phenomenology This section uses straightforward dimensional analysis to find an expression for T in a ?-layer using weak localization data. Although reasonably simple, this analysis reveals certain features of this approach to measuring T which all models have in common. 52 To proceed, consider the data in Figure 1.3, which shows both ?(H ? ) (corresponding to the upper curve in the figure) and ?(H bardbl )(lowercurve).Fora particular value of magnetic field, say H 0 , the graphs show that the maximum value of ?(H 0 ) is attained when H 0 = H ? , while the minimum occurs for H 0 = H bardbl . Furthermore, both curves are symmetric functions of H (i.e. ?(H)=?(-H)), which suggests that ?(H) can be represented by a Taylor series expansion with only even powers of H: ?(H ?,bardbl )= ? summationdisplay n=0 ? 2n H 2n ?,bardbl . (2.42) Then to lowest non-vanishing order in H,wecanwrite ??(H ? )=? ? H 2 ? (2.43) and ??(H bardbl )=? bardbl H 2 bardbl (this is the lowest order Taylor expansion of equation (2.41)). From our discussion of weak localization and the data of Figure 1.3, it is clear that a greater magnetic flux penetrates the 2DEG when H = H ? (since the conductance is higher), and therefore that we can associate an e?ective area, A eff (or cross section), with ? ? and ? bardbl , as follows. First, I adopt the convention that ??(H) is dimensionless, measured in ?Cooperon units? of e 2 /2? 2 planckover2pi1 =e 2 /?h. Then in order for both sides of equation(s) (2.43) to be dimensionless, ? must have units of 1/T 2 . Because weak localization is due to interference between pairs of paths, the natural units of magnetic flux in 53 this problem are planckover2pi1/2e, and this combination of physical constants has dimensions of T?m 2 . Hence the desired relationship between ? and A eff in this approximation is given by ? = parenleftbigg 2eA eff planckover2pi1 parenrightbigg 2 . (2.44) The ?cross section? argument presented in Chapter 1 is shown again in Figure 2.14. When H = H ? , all directions in the 2DEG should be treated symmetrically, which suggests that a circle is the appropriate cross sectional area to be associated with A ? . Furthermore, the characteristic length scale of the 2DEG in this configuration is lscript ? , and using this (subjectively!) as the diameter of the circle, we have ? ? = parenleftbigg e?lscript 2 ? 2planckover2pi1 parenrightbigg 2 . (2.45) On the other hand, when H = H bardbl , all directions in the device are not equivalent, since, due to the confinement potential, the wavefunction is coherent in the direction perpendicular to the doping plane. In fact, if the sample is to be considered two-dimensional from a quantum mechanical perspective, the condition T = parenleftbigg 1 A parenrightbiggintegraldisplay T (x,y)dxdy (2.48) should be used, where A is the total area of the device. To account for how the T (x,y) distribution influences ??(H), I introduce an extremely useful construct which captures the essential statistical information. I define the ?thickness autocorrelation function?, C T (r) as follows: C T (r)=< T (0)T (r) >. (2.49) The bracket < ? > implies ?averaging over all values of the origin?, and r represents the distance from the origin to the point of interest. In other words, 58 equation (2.49) quantifies the average correlation between the thickness at an arbitrary point in the sample (taken to be the origin) with all other points a planar distance |r| away. Using the Wiener-Khintchine theorem,thiscanbeshown to be equivalent to the following expression: C T (r)= integraldisplay ? ?? T 2 (r)e ik?r dk (2.50) i.e. C T (r) is the Fourier transform of the absolute square of T .Atomicforce microscope (AFM) images of surfaces have established that C T (r)takesthe following form for random Gaussian surfaces: C T (r)=? 2 ? e ?(r/lscript c ) 2H (2.51) Where ? ? is the rms surface roughness, lscript c the correlation length,andH the so-called ?scaling exponent? (usually Hsimilarequal1). Although such a surface model is probably of limited utility in ?-layer physics, we can nonetheless identify lscript c as the appropriate statistical quantity for taking into account how the ?roughness? of the ?-layer influences the magnetoresistance. A precise estimate of the ?-layer thickness will depend on the relationship between lscript e , lscript ? ,andlscript c . For example, as discussed in [62], if the inequality lscript c lessmuch lscript e lessmuch lscript ? holds, the situation is analogous to ?homogeneous broadening? in nuclear magnetic resonance, and the weak localization line shape (?(H)) is unaltered, as sketched in Figure 2.16 (a). In the ?inhomogeneous broadening? limit (Figure 2.16 (b)), lscript e lessmuch lscript ? lessmuch lscript c , and the line shape is altered, showing a change of sign from the ideal line shape (zero roughness limit) as a function of H ? . I apologize for my lack of artistry in crafting this figure. 59 (a) (b) Figure 2.16: NMR analogy for parallel field weak localization measurements based on the length scales (lscript c , lscript e , lscript ? ) characterizing the ?-layer. The dashed curves rep- resent ?(H)forzero parallel field, while the solid curves represent a finite (and constant) parallel field. The ?homogeneous broadening? limit corresponds to (a), applicable when lscript c lessmuch lscript e lessmuch lscript ? . The ?inhomogeneous broadening? limit corresponds to (b), and is applicable when lscript e lessmuch lscript ? lessmuch lscript c . Previous measurements, not reported in this thesis, suggest that the samples studied here are homogeneously broadened. 60 I will consider only two models for lscript c in this thesis. In the first case, I completely ignore any e?ects of lscript c on ?(H), and estimate an ?uncorrelated? thickness, T : T = planckover2pi1 ? ? 2elscript ? . (2.52) This corresponds to the cross section argument presented earlier. In the second case, I will assume homogeneous broadening of the line shape (lscript c lessmuch lscript e lessmuch lscript ? ), and estimate a ?correlated? thickness, T c : T c = parenleftbigg 1 4? parenrightbigg 1/4 bracketleftBigg parenleftbigg planckover2pi1 elscript ? parenrightbigg 2 parenleftbigg lscript e ? lscript c parenrightbigg bracketrightBigg 1/2 (2.53) Since independently measuring lscript c is presently impossible (due to the fact that the ?-layer lies beneath the device surface), all estimates of T in this dissertation will use the uncorrelated estimate, equation (2.52). The lone exception to this will be when I discuss Si:P di?usivity measurements in Chapter 5, and model the data using (2.53), to be consistent with the published results. Regardless of the theoretical limitations facing a given thickness model, it seems safe to assume that the devices measured in this work were homogeneously broadened. This is suggested by previous measurements, not reported in this thesis, which more closely mimicked the homogeneous broadening case of Figure 2.16 (a). In addition, for our prototypical device lscript e similarequal 25 nm and lscript ? similarequal 100 nm, while AFM images on device surfaces have shown lscript c similarequal 5 nm. Although lscript c of the surface and that of the ?-layer are undoubtedly distinct, the likelihood that lscript c of the ?-layer exceeds 100 nm seems remote. A natural question to ask is what actually causes the fluctuations in T ?It 61 seems likely that this disorder stems from two sources. The first is the intrinsic roughness of Si surfaces. A perfectly flat atomic surface does not exist, and for Si in particular the quality of the surface depends heavily on the temperature of the substrate during growth. Because of the technological importance of Si, such surface features have been extensively studied and characterized. According to the experts who actually grow ?-doped samples, this source of disorder is likely to be the limiting factor to achieving monolayer ?-doped samples, once other, less intrinsic issues, are overcome (assuming they can be). The second source of fluctuations in T is from the phenomenon of segregation. This will be discussed more fully in Chapter 3, but for now it su?ces to define segregation as the tendency of donors to migrate towards the sample surface during epitaxial growth. In other words, the donors do not remain on the surface where they were deposited, which leads to a broadening of the ?-layer. This tendency is much more pronounced in Si:P than in Si:B, and has recently been a subject of considerable interest. At the time of this writing, it is safe to say that segregation is the dominant contribution to a non-uniform T in Si:P ?-doped samples. 62 Chapter 3 Fabrication The details of how a ?-doped wafer is constructed, and the processing methods used to fabricate devices for experimental investigation, are presented in this chapter. A comparison is drawn with the approaches implemented by other research groups, and the use of SIMS analysis in characterizing the quality of the samples is discussed. The wafers investigated in this dissertation were prepared by Dr. Philip E. Thompson, head of the Si MBE section at NRL. This work would not have been possible without his expertise, and his contribution is greatly appreciated. 3.1 ?-layer Requirements Referring to Figure 1.1, it should be obvious that ?-layers do not occur in Nature: they must somehow be built. As mentioned in the Introduction, several methods of ?-layer construction have been demonstrated, including ion 63 implantation [9], chemical vapor deposition (CVD) [8], and molecular beam epitaxy (MBE), as in this and many other works. I shall first outline the physical characteristics all ?-layers should have. 3.1.1 General Considerations Without question, MBE is by far the most powerful technique for growing ?-doped samples. What makes MBE so elegant is that, under suitable conditions, the lattice planes are grown atom by atom.Thistypeofcontrolispreciselywhat is sought in a ?-layer sample, where ideally donor placement in the c-Si lattice lies in a single lattice plane (i.e. in a monolayer). The basic idea behind MBE is straightforward. A substrate is placed in the MBE chamber to serve as the ?base? for the heterostructure to be grown. Various ?e?usion cells? (specialized ovens) are distributed throughout the chamber, which act as sources for the di?erent materials to be grown into the sample. The growth rate of a given material depends on the temperature of its cell, as well as the temperature of the substrate. Thus, in a loose sense, MBE is similar to the common technique of metallic evaporation. The MBE chamber is held at ultra-high vacuum (UHV), and therefore contamination by unwanted impurities is kept to a minimum. Regardless of whether MBE or another approach is used, there are essentially only two factors which characterize the quality of a ?-doped sample. One, of course, is the thickness, T , and the smaller is T , the better. The other aspect is multifaceted, namely the defect density inherent in a particular method of 64 preparation. Obviously a sample with a lower defect density is ?better?, and there are many ways of quantifying this. Some examples include an asymmetry of the ?-layer (which, due to segregation, tends to be much worse in Si:P than Si:B), lattice defects (such as vacancies and interstitials), and impurities which ?pollute? the ?-layer in a fashion other than intended. While quantifying T is relatively straightforward in that it reduces to a length measurement, the matter of defect specification is much more di?cult. For instance, in the case of MBE grown layers, the temperature sequence incorporated during the various stages of device processing has a profound impact on the electrical characteristics a device will display. The details of this will be discussed in Chapter 5 with respect to how the resistivity of a device is influenced by a high temperature (850 ? C) anneal. At present, eliminating dopant segregation is a major area of research in the production of Si:P ?-layers. This phenomenon was described in Chapter 2 as the tendency of the donors to migrate toward the surface of the sample during growth, leading to a broader ?-layer. To see how other workers have addressed this issue, and to contrast their approaches with those used in the course of this research, I briefly digress to discuss the published results. 3.1.2 Previous Work Three papers [64, 65, 66] on Si:P ?-layers are summarized here. The basic idea behind the MBE growth techniques used by all workers who have reported results on Si:P ?-layers, is illustrated by the ?naive? MBE growth sequence shown in 65 Figure 3.1: A naive approach to MBE ?-layer growth. A minimum of three steps are necessary to isolate the ?-doped region from the (100) Si substrate: 1) bu?er growth (or annealing), to prepare a clean surface, 2) ?-doping, in which the donors are deposited, and 3) ?cap? growth, which isolates the ?-layer and keeps unwanted impurities out of sample. Layer widths are not to scale. Figure 3.1. In this approximation, there are only three growth steps to perform. A clean substrate of high-resistivity p-type (100) Si is used as the ?base? for the rest of the heterostructure. The reason for using a p-type (100) wafer is to better isolate the n-type ?-layer from the substrate. The first step is to prepare a pristine surface for deposition of the P atoms, either by growing a so-called ?bu?er? layer of pure Si, or by annealing the sample at high temperatures. In the second step, ?-doping takes place by depositing P atoms (in one form or another) for a specified amount of time. This controls the doping density, n. The last step in the process is to encapsulate the ?-layer by growing pure Si (the ?cap? layer) on top of the ?-doped region. This serves to isolate the P donors from the environment. 66 T.-C. Shen, et al., at Utah State University, were the first to report results for a Si:P ?-doped wafer [64]. This work focused on the surface morphology of the sample during the various stages of growth, rather than end-result electrical characteristics. The preparation sequence employed by this group was more elaborate than the ?naive? approach described above, and consisted of the following steps: 1. Start with a high-resistivity (0.1 ??cm) Si(100) substrate. 2. ?Flash? anneal the sample at T = 1500 K to prepare the surface for P deposition. 3. Expose the substrate to a ?dose? of phosphine (PH 3 ) gas at room temperature (RT). The P density in the ?-layer is controlled by the pressure and duration of the exposure. 4. Anneal at T = 510 K to dissociate the PH 3 and adsorb the P atoms into the substrate. 5. Deposit 0.4 nm of Si at T = 428 K. 6. Anneal at T = 773 K. 7. Deposit the cap layer, an additional 6.3 nm of Si at T = 468 K. 8. Anneal at T = 673 K. Presumably, the annealing steps were employed to reduce the defect density in the sample. This is born out by the STM (scanning tunneling microscope) analysis 67 they performed of the growth stages, which showed low surface roughness and c(2x2) structures, characteristic of high-quality Si surfaces. For the two samples discussed in [64], transport measurements yielded for one sample n = 1.3x10 14 /cm 2 , with a mobility of ? =13cm 2 /V?s, while for the other sample they found n = 8.8x10 13 /cm 2 ,and? =38cm 2 /V?s. No measure of T in these samples was reported, nor were values for lscript ? , although they were able to observe weak localization in devices fabricated from these wafers. A similar (and independent) approach to Si:P ?-layer construction is reported in [65], obtained by L. Oberbeck, et al., at the University of New South Wales (UNSW) in Australia. For comparison with [64], I also enumerate their preparation sequence: 1. Start with a high-resistivity (? 1??cm) Si(100) substrate. 2. ?Flash? anneal the sample at T = 1150 ? Cfor? 30 seconds to prepare the surface. 3. Expose the substrate to PH 3 at RT. 4. Anneal at T = 550 ? Cfor5minutes. 5. Deposit 24 nm of Si at T = 250 ? C. These workers performed a slightly more thorough characterization of their samples by performing a SIMS (Secondary Ion Mass Spectrometry) analysis, in addition to STM and weak localization measurements. This led to identification of the annealing step after the PH 3 dose as the key element to limiting segregation. 68 Their explanation of this e?ect is that as the P atoms are adsorbed into the Si crystal, they become bound by three covalent bonds, instead of just one as when the P atom resides on the surface. Transport measurements yielded n = (2.0 +/- 0.4)x10 14 /cm 2 ,and? similarequal 140 cm 2 /V?s. They also observed weak localization, but did not report values for lscript ? . The final work I shall discuss was performed by K.E.J. Goh, et al.,alsoat UNSW [66]. This is by far the most complete study on segregation in Si:P ?-doped samples reported to date. Here a systematic study of the substrate temperature on P segregation into the cap layer was performed, with di?erent samples prepared at RT, 250, 400, and 600 ? C. This paper does report their weak localization results, which serves as an important piece of the puzzle for optimizing ?-layers grown using phosphine. The largest value of lscript ? (and therefore the ?best? from a quantum transport perspective) was obtained for the sample whose cap layer was grown at T = 250 ? C, and for this sample the corresponding device parameters were determined to be ? ? 61 cm 2 /V?s, n = 1.64x10 14 /cm 2 ,andlscript ? =72nm.Their transport measurements were performed at T = 4.2 K. Since the research in [65, 66] relied heavily on the use of SIMS to assess the quality of their process, a brief discussion of that technique is in order. SIMS SIMS is a very powerful technique for acquiring profiles of elemental concentrations versus depth in a given material sample. An example of a SIMS profile from an Si:P ?-doped sample is shown in Figure 3.2. The way such a profile 69 is obtained is to bombard the sample with a high-energy (? keV) primary ion beam, which sputters matter away and leaves a ?crater? behind. The ejected material consists of both neutral atoms and secondary ions. The secondary ions are separated from the other matter electrostatically, then directed into a mass spectrometer tuned to detect the particular elemental mass of interest. The primary beam is rastered in order to create a flat-bottomed crater, but the ions are acquired only from the center of this crater, which ensures that each data point comes from a unique depth. The intensity of the secondary ion signal is proportional to the ion concentration at a given depth, whose scale is quantified by the use of reference samples and surface profilometer measurements. By correlating the crater depth with the secondary ion intensity, the desired depth profile is constructed. It is also worth noting that SIMS can distinguish the signals between di?erent isotopes of a given element, which has been used to explain the large surface signal in Figure 3.2. According to [65, 66], this signal is due not to P (with atomic number Z = 15), but to ?false counts? induced by the simultaneous detection of Si (Z = 14) and H (Z =1). The detailed procedures incorporated in a given SIMS analysis are rather extensive, and must be optimized for the particular application involved. The doping profile shown in Figure 3.2 was performed by the analytical group at Charles Evans and Associates, and used a quadrupole SIMS instrument with Cs as the primary ions. The incident ion energy was 1 keV, and the angle of incidence was 60 ? . The P detection limits for this analysis were 4x10 15 /cm 3 . The ?straggle? of the doping distribution toward the surface is a clear manifestation of 70 Figure 3.2: SIMS data for a Si:P ?-doped sample, with a large amount of segregation apparent (the region between 10 and 50 nm). The signal peak at the surface of the sample is likely due to Si and H (total protons = 14 + 1 = 15) ?masquerading? as P (15 protons) [65, 66]. 71 segregation. Although this SIMS analysis has yielded a detailed cross section of the doping profile in the sample, several limitations of this technique are worth pointing out. First, T at only one point in the crystal is sampled, and examining many areas would require a substantial amount of time (and money). Second, the technique is destructive. By boring a crater in the sample through ionic bombardment, the P atoms in the ?-layer are to some extent forced further into the crystal. This often manifests itself in the profile appearing steeper near the surface of the sample than on the ?backside?, and thus leads to an erroneous estimate of T . Third, little to no information in regard to electrical characteristics or defect densities can be inferred from SIMS. Fourth, and perhaps most importantly, the spatial resolution of SIMS is limited. Determining the precise resolution limits of SIMS for Si:P ?-layers is presently an unresolved question, but it is safe to say that sub-monolayer resolution is not possible using SIMS. In contrast, weak localization measurements for determining T su?er from none of these limitations. Chapter 5 will explore this more fully. 3.2 ?-doped Wafer Preparation With the necessary background in place, I now discuss the preparation sequence for the ?-doped wafers used in this work. I shall first comment on some particular aspects of our approach. First, the primary di?erence between the samples used here versus those of 72 Figure 3.3: ?Old? ?-doping technique. In an e?ort to minimize P segregation, a sequence of bu?er layers are incorporated in the growth as the substrate temperature is reduced. Bu?er 1 corresponds to 170 nm of Si grown at T = 650 ? C, bu?er 2 to 18 nm of Si grown as T was reduced to 400 ? C,andbu?er3to12nmofSiwithT held at 400 ? C. The ?-layer and cap layer (50 nm of Si) were grown with T = 400 ? C. No trap layer was implemented, and the layer thicknesses are not to scale. 73 [64, 65, 66] is that the P was obtained through thermal decomposition (?burning?) of GaP. The Ga was separated from the P by passing the GaP ?fumes? through a B filter, apparently producing a very pure P source. Second, the substrates were always high-resistivity p-type Si(100) wafers. As mentioned earlier, a p-type substrate more e?ectively isolates the n-type ?-layer from the rest of the sample, but more importantly, a ?-layer in Si can only be grown along the (100) lattice direction [63]. Finally, typical vacuum levels during growth were ? 5x10 ?9 Torr, and the wafers were rotated at ?30 rpm during growth to promote a uniform doping density across the wafer. An actual growth sequence for a ?-doped sample is depicted in Figure 3.3. The major di?erence between this recipe and the ?naive? one described earlier is the temperature sequence the substrate is taken through in the course of growth. After placing the substrate in the MBE chamber, the wafer?s temperature is raised to T = 650 ? C and a 170 nm bu?er layer of pure Si is grown. Next, the substrate temperature was reduced to T = 400 ? C over the course of 3 minutes, then held at T = 400 ? C for the remainder of the growth process. The total Si growth during the time the sample has its temperature reduced and stabilized is ? 30 nm (a constant Si growth rate of ? 0.1 (nm/s) is assumed). At this point the P ?-layer itself is grown by opening the shutter to the P source for 200 seconds, which produces a doping density of ?10 14 Patoms/cm 2 . The heterostructure is completed by growing a 50 nm cap layer with the substrate temperature held at T = 400 ? C. The SIMS analysis for the ?-layer produced by this process was shown in Figure 3.2. The most obvious feature of this plot is that the ?-layer itself is not 74 Figure 3.4: Growth sequence of ?-doped sample incorporating both substrate cooling and an Si trap layer. This technique produces the sharpest ?-profiles observed in this work. Bu?er 1 corresponds to 150 nm of Si grown at T = 650 ? C, and bu?er 2 to 18 nm of Si grown as T was reduced to 320 ? C. The ?-layerand4nm?traplayer? were grown with T = 320 ? C, while the 50 nm cap layer was grown at T = 500 ? C. All magnetoconductance data presented in this thesis were taken on samples using this type of growth sequence. Layer thicknesses are again not to scale. very sharp, but rather displays a skewness or ?straggle? towards the surface. This is segregation, and is especially pronounced in Si:P. The low growth temperature used (400 ? C) was intended to minimize this segregation, but clearly was not entirely e?ective. The best growth technique used in this work is shown in Figure 3.4. Two bu?er layers are grown to isolate the ?-layer from the substrate, the first being a layer 150 nm thick grown at T = 650 ? C, the second a 18 nm layer grown as the 75 substrate temperature was reduced from 650 ? C to 320 ? C over the course of three minutes. The ?-layer was then grown at 320 ? C by opening the P shutter for 200 seconds. The next step is the most crucial part of the entire process: 4 nm of undoped Si were grown on top of the delta-layer at 320 ? C in an e?ort to ?trap? the P donors to a narrower region than had been obtained by the methods of Figure 3.3. Incorporating this step was motivated by [65]. All growth (Si and P) was then stopped, and the substrate temperature was raised from 320 ? C to 500 ? C over a 3 minute interval. Finally, a 50 nm cap layer was grown to complete the process. A SIMS analysis from a sample of this wafer is shown in Figure 3.5. Clearly something rather dramatic has occurred. Whereas when using the ?old? growth method the sample showed a great deal of segregation, the ?new? sample shows almost none. In addition, the doping profile is much more symmetric about its maximum doping density level, and in general appears to be much more ??-like?. These are two very significant improvements to the quality of the ?-doping profile, and can be traced back to the implementation of the trap layer in the growth sequence. A heuristic explanation for this e?ect goes somewhat as follows. Since segregation is more pronounced at higher temperatures, it is desirable to grow the ?-layer at as a low a temperature as possible. Unfortunately, the growth of high?quality c-Si has just the opposite temperature dependence: the higher the Si growth temperature, the fewer intrinsic lattice defects (vacancies and interstitials) are produced [20]. It is also well-known that di?usion is enhanced by the presence of such defects in the crystal [85]. Thus we see that growing Si at a higher temperature will reduce the intrinsic di?usivity of P in Si, and 76 Figure 3.5: SIMS analysis of a ?-doped wafer grown with the 4 nm thick trap layer. A much narrower and more symmetric ??-like? profile is obtained, showing the importance of the trap layer for Si:P ?-layer growth. 77 consequently will help to produce a narrower ?-layer because the P di?uses less in the c-Si lattice. These two opposing requirements are balanced by the trap layer. By growing a narrow (? 4 nm) Si trap layer on top of the ?-doped region and then heating the substrate, the P donors do segregate to the surface, but no farther: they are bound to this region. Thus the name ?trap layer? is highly suggestive of its function in that it limits the P segregation to a well-defined portion of the crystal. The cap layer can then be grown at a higher temperature (T = 500 ? C) without being overly concerned with P segregation. It should also be mentioned that a sample with a lower defect density generally has a higher electronic mobility, which is often desirable in nanostructure devices. This insight presents interested researchers with a wide range of potential avenues for investigation. For example, does the ?-layer width correlate strongly with the trap layer width? That is, if the trap layer were grown to be only 1, 2 or 3 nm thick, would this correspondingly produce a narrower ?-profile? Another interesting question is precisely what temperatures and growth rates optimize the process? In other words, can true Dirac-delta (monolayer)dopingprofilesbe achieved in Si:P using MBE? The available parameter space to be explored is enormous, and will not be discussed further here. Instead, we shall content ourselves to simply note that very sharp, ?-like doping distributions can indeed be achieved in MBE grown Si:P ?-layers, provided that a trap layer is incorporated in the growth process. For reference, a step-by-step algorithm to grow a high-quality ?-layer is included below. 78 1. Clean p-type Si(100) substrate (includes NRL clean and 2.5% HF dip to remove native oxide) and place in MBE chamber. 2. Increase substrate temperature to T = 650 ? C for bu?er layer growth. 3. Bu?er layer growth: ? 150 nm grown at T = 650 ? C. Si growth rate is ? 0.1 nm/s. 4. Reduce substrate temperature to T = 320 ? Coverthreeminutesfor?-layer growth. Total Si growth in this interval is ? 18 nm. 5. Grow ?-layer with substrate temperature at T = 320 ? C. A 200 second exposure to the P source produces a doping density of ? 10 14 /cm 2 . 6. Trap layer growth: 4 nm of Si grown at T = 320 ? C. 7. No growth of either Si or P as substrate temperature is increased from T = 320 ? C to T = 500 ? Coverthreeminutes. 8. Cap layer growth: 50 nm of Si grown with T = 500 ? C. Once a ?-doped wafer has been prepared as described, the next step in the experimental investigation is to prepare mesoscopic devices for measurement. This fabrication process will now be described. 3.3 Device Fabrication One of the primary goals of the fabrication process is to keep the sample as free from contaminants as possible. Therefore, a substantial number of the 79 processing steps are devoted to ?cleaning? the device. In addition, the samples are inspected optically throughout the process to be sure that nothing obvious has gone awry. Even if things ?look? fine with the device (i.e. no defects are apparent), the only meaningful check is actual electronic testing. Such testing will be discussed at the end of this chapter. Before any processing is done to a sample of the ?-doped wafer, it is advisable to simply check whether a piece of it conducts at low (liquid helium) temperatures. This is done by ?scribing? a small piece o? of the wafer, placing 4 In solder ?blobs? on the edges of the chip, connecting Au wires to the blobs, attaching the chip to a DIP (dual inline pin) header, and dipping the sample in liquid helium. If thesampledisplaysalowtwo-terminal resistance (R ? 10 k?) between any two contacts, then the enumerated procedure described below can be followed to make an actual device. If this test does not work, i.e. the two-terminal resistances are very high (R ? 1 M?), then obtaining a new sample is in order, rather than going through the laborious process of fabrication only to find, in the end, that the device does not work! Here, then, is a step-by-step algorithm for producing a Hall bar device suitable for measurement. 1. Soak wafer in concentrated HCl to remove any residual ?scum? from previous handling (? 2 hours). 2. Dice wafer into approximately 5 mm x 5 mm pieces for use in actual device fabrication. The DIP header on which completed devices were mounted has 80 awidthof?7 mm, therefore the dies used must be smaller than this. 3. Non-chlorinated (NC) solvent clean. In order, the chip is sprayed with acetone, methanol, and isopropanol alcohols, then ?blow-dried? with N 2 gas. 4. Spin on resist. Typically Futurrex NR7 1500 P negative resist was used. The resist was spun at 5000 rpm for 1 minute. 5. Bake sample in oven at T = 120 ? C for 1 minute to perform initial ?cure? of resist. 6. Perform electron beam lithography using the scanning electron microscope (SEM) on the resist to define a device pattern. The reason e-beam lithography was used is twofold. For one, it is very easy to make changes to a given pattern or design entirely new devices by using the NPGS software which controls the SEM. Photomasks lack this flexibility. Second, by using the SEM, annoying issues associated with photomasks are avoided, such as proper contact between the sample and mask or the sample sticking to the mask. In a nutshell, the use of e-beam lithography makes the fabrication process more reliable. For reference, a typical exposure dose was ? 10 ?C/cm 2 . 7. Post-bake sample in oven at T = 120 ? C for 1 minute for final ?cure? of resist. 8. Develop resist in appropriate solvent (RD6 for Futurrex negative resists). 9. Dip in de-ionized (DI) water to remove solvent from exposed sample surface. 81 10. Etch device pattern into ?-doped sample using RIE (Reactive Ion Etch). An SF 6 plasma was used to etch the Si 20-30 nm below the ?-layer in order to expose an edge of the doping plane for subsequent electrical contact. A ?mesa? structure results from this etch (see Figures 3.6 and 3.7). 11. Clean resist from sample by soaking in a ?piranha? bath ?30 minutes. The piranha bath consists of a mixture of H 2 S0 4 and H 2 O 2 in a 3:1 ratio. 12. Measure the mesa profile to ascertain how much Si the SF 6 etch removed. The ?-layer should nominally lie somewhere near the middle of the ?wall? profile of the mesa. For example, if the chip were from the same wafer as the SIMS analysis of Figure 3.5, the wall should have a height of at least 60-75 nm. 13. Perform 0 2 plasma clean (?descum?) in RIE to further remove residual resist (and possibly other) contaminants. 14. Dip sample in concentrated HF solution to remove the SiO 2 produced by the O 2 plasma of the previous step. 15. Clean ?polymer scum? (CF x , Si-C, Si-F, etc.) o? of sample surface left as a result of the SF 6 etch. This is done by placing the sample in a bath of the ?barracuda? clean, a mixture of NH 4 OH, H 2 0 2 ,andH 2 0 in the ratio 1:1:5 for ? 30 minutes [67]. This barracuda cleaning step is crucial, since the polymer scum otherwise causes contacts to the ?-layer to become insulating at low temperatures, making electrical measurements impossible. 82 16. Dip sample in DI water, and perform NC solvent clean. 17. Spin and bake resist for second layer of lithography as in steps 4 & 5. 18. Perform optical lithography to define ??-contact? pads (see Figure 3.7). The resist exposure is done with a photomask and ultraviolet light on a contact aligner. 19. Post-bake and develop resist as in steps 7 & 8. 20. Dip sample in HF:H 2 O (1:100) solution to remove any oxide which has accumulated at the surface of the ?-layer. 21. Perform first metalization. In a thermal evaporator, deposit ? 2.5 nm of Cr, then enough of the Au-Sb (99:1) alloy to exceed the mesa wall height by 10-20%. The reason Au-Sb is used is to lower the Schottky barrier associated with the metal-semiconductor contact. 22. ?Lift-o?? resist and unused Au-Sb in RR2 solvent. For this step, the RR2 has T = 100 ? C, and the sample is allowed to soak for ?20 minutes. 23. If the RR2 lift-o? still has not removed all of the unused Au-Sb, place the sample in a beaker of DI water and subject the beaker to ultrasonic treatment for 1 minute intervals until the lift-o? is complete. 24. Anneal the sample in a rapid thermal annealer (RTA) for 5 minutes at T = 450 ? C. The ambient atmosphere in the RTA is a combination of N 2 and H 2 gas. The purpose of the annealing is to di?use the Au-Sb alloy into the 83 ?-doped region of the sample and further lower the Schottky barrier. This step is important for establishing an Ohmic electrical connection with the ?-layer (i.e. a connection which displays a linear I-V curve). 25. NC solvent clean. 26. Spin and bake resist for third and final layer of lithography. 27. Perform optical lithography (including development) to define ?bonding pads? which overlap with the ?-contact pads (see Figure 3.7). 28. Perform second metalization. In a thermal evaporator, deposit ? 2.5 nm Cr and ? 100 nm Au. 29. Lift o? using RR2 as in step 22. 30. NC solvent clean. 31. Connect Au wires to bonding pads with a wire bonder. 32. Fasten sample to DIP header using either GE varnish or vacuum grease. 33. Use silver paint to finally connect Au wires to DIP header contacts. The reason for using Ag paint is that the header pins cannot be directly wire bonded to, and the use of a solder such as In generally causes magnetic shielding due to the Meissner e?ect (to be discussed more fully in Chapter 4). Generally speaking, this process yields high quality devices for measurement. Figures 3.6 and 3.7 show the evolution of the sample from 84 Figure 3.6: A ?horizontal profile? of the device fabrication process flow. In a), the device?s dimensions are defined by electron beam lithography. In b) the device pattern is etched into the ?-doped Si with an SF 6 etch. Electrical contacts are made via 2 metalizations as in c), and finally wires bonded are to the metal in d). At this point the device is ready to be mounted on a header and tested at T = 4.2 K. 85 (a) (b) (c) Figure 3.7: Approximate device evolution following process flow sequence of Figure 3.6. First the ?mesa? is defined with electron beam lithography and a dry SF 6 etch, as in (a). Two metalizations follow to make electrical contacts with the ?-layer. The first evaporation is Au-Sb (b), for the actual ??-contacts?, the second (c) is pure Au, which overlaps the Au-Sb contacts and provides a larger surface area (? 250 ?m x 250 ?m) for wire bonding. 86 horizontal and vertical perspectives, to give the reader a visual feel for the major fabrication steps required. 3.4 Device Testing Assuming the ?-doped wafer is suitable for use in device construction (i.e. conducts current at low temperatures, etc.), at this stage the sample could be mounted on the dilution refrigerator and cooled to base temperature for measurement. However, it is prudent to perform some final diagnostic tests to ensure that the device will behave as desired and be straightforward to measure. The first such test is a simple ?continuity? check at room temperature to verify that all probes on the sample are in fact electrically connected with one another. This is quickly and easily done with a hand-held DVM. The second check is to ?dip? the sample in a liquid helium dewar and then re-measure the contact resistances. If there are any open circuits at 4.2 K but not at room temperature, then it is necessary to evaluate what may have gone wrong in the processing to cause this. Depending on the intrinsic resistivity of the sample, the voltage probes should display a low resistance between each other, and the resistance measured between adjacent contacts around the sample should be very nearly equal. The third test is to establish whether or not any conduction through the 87 substrate is taking place at 4.2 K. This is accomplished by the use of an ancillary contact, connected to the ?-doped device only through the substrate. The resistance between this ?floating? contact and any probe on the sample should be infinite, or at least much higher than the resistance between any two probes on the device. Otherwise, the current in the sample is not confined to the lithographically defined device region, and resistance measurements become meaningless. Finally, the sheet resistance (resistivity) can be checked with a four-wire measurement to get a rough idea of what the signals will be like at base temperature of the dilution refrigerator. This merely serves as a benchmark for the more accurate magnetoresistance measurements to be performed, which will be discussed in the next chapter. 88 Chapter 4 Measurement In this chapter I will discuss how four-wire conductance measurements on ?-doped devices were performed. The necessary background is developed by discussing the conductivity tensor for two-dimensional systems, the instrumentation used in making the measurements, the particular techniques employed for obtaining data, and the systematic errors involved in these experiments. 4.1 Background A proper 4-wire measurement [68] of a resistance R s is depicted in Figure 4.1 (a). The device is biased with a known current I using one pair of wires, and the voltage V induced by the current is measured with a second pair of wires. R s is given by Ohm?s law, i.e. by forming the ratio R s =V/I. 89 Typically the current source is realized by placing a large, known resistance R L in front of a known voltage source V 0 . The much smaller resistance of R s is then placed in series with this arrangement, as shown in Figure 4.1 (b). Since R L dominates the circuit impedance (R L greatermuch R s ), the current I is e?ectively given by I = V 0 /R L . Voltage measurements across R s are generally made using the combination of a di?erential preamplifier and a digital voltage meter (DVM). This high input impedance arrangement limits ?shunting? of current from the device, which would otherwise corrupt the measurement. Obviously, the more accurately V and I are known, the better the measurement. As straightforward as this 4-wire technique is, it is also incredibly powerful. Since the resistance of a device depends on numerous parameters of physical interest, such measurements can be used to extract useful information about a particular electronic system. For example, the conductance measurements on ?-doped devices discussed in this thesis depend on ?external? variables such as the sample dimensions, temperature T, magnetic field H, current density J, and applied electric field E,aswellas the frequency of these fields, ? i . Quantities ?internal? to the ?-layer, such as the carrier density n, mean free path lscript e , dephasing length lscript ? , e?ective mass m ? , and, of course, the ?-layer thickness T , can be deduced through the application of appropriate theoretical models to properly carried out resistance measurements. It should also be noted that resistivity (?) refers 90 Figure 4.1: 4-wire measurement schematic. In (a), a current bias, I, is passed through the device, and a voltage V is measured. (b) shows a practical realization of this, which incorporates a voltage source V 0 in series with a large resistance R L to serve as the current source. The resistance of the sample, R s , is deduced by measuring the voltage V s across the sample with a di?erential preamplifier and DVM, then forming the ratio R s =V s /I. 91 to the intrinsic electrical characteristics of the material system under study. Given ? and the geometry of a given device, one should be able to predict what resistances will be measured under a given set of conditions. I shall discuss this in more detail later. An elaborate theory of conductance measurements and their symmetry properties has been developed over time. We cannot go into this in any detail here, but merely cite a few results pertinent to this thesis. Since this work concerns magnetoconductance measurements of c-Si, Onsager?s result [69] that the local resistivity tensor of a cubic crystal obeys the symmetry relation ? ij (H)=? ji (-H) suggests that an analogous (though not identical) expression should hold here. Landauer?s many contributions to the modern understanding of conduction processes include quantifying how local variations of E in the conductor due to scatterers influence transport [70], modeling conduction using the basic quantum mechanical notions of reflection (R) and transmission (T) coe?cients, and generalizing these notions to real, multichannel conductors [71]. van der Pauw?s result [72], showing that the resistivity of an arbitrarily shaped conductor could be found through a conformal mapping, further established the power of 4-wire measurements, and will be used extensively in Chapter 5. Finally, B?uttiker?s reciprocity theorem [73] regarding 4-wire measurements in the presence of a magnetic field H leads to the global symmetry relation R ij,kl (H)= R kl,ij (-H). In this expression, the first pair of indices refers to the current 92 leads, the second to the voltage leads. Thus a perfect conduction symmetry is observed in a device under magnetic field reversal, if, in addition, the current and voltage leads are also exchanged with one another. I shall discuss this further in Chapter 5 as well. 4.2 The Two-Dimensional Conductivity Tensor Since the electron system associated with a ?-layer is inherently two-dimensional, in general a tensorial relationship will exist between the electric field E (related to the measured voltage V by V = - integraltext E?dl)andthe current density J in such a device. It is immediately apparent that the units of J in two dimensions are (amperes/meter) instead of the usual (amperes/meter 2 ), as it is for three-dimensional conductors. I shall elaborate on this shortly. In this work, R = V/I is a function of both magnetic field H and temperature T: R(H,T)= V (H,T) I(H,T) . (4.1) Such an expression makes the use of the term magnetoresistance obvious, and is really the ?scalar? version of the aforementioned tensorial relation. In order for this expression to be meaningful, a linear response must be observed between E and J, and therefore a more precise definition would be 93 R = dV/dI. This linear dependence was observed in all of the samples discussed here. Depending on the causal point of view one wishes to take, one can consider either the resistivity ? or the conductivity ?. In the first case the application of E ?causes? J to flow in the device, and Ohm?s law takes the form J = ?E. (4.2) One can also take the ?inverse? approach, and write E = ?J. (4.3) Accordingly, ?? = 1,where1 is the 2x2 identity matrix. For a Cartesian two-dimensional system, equation (4.3) is expressed in matrix notation as: ? ? ? ? E x E y ? ? ? ? = ? ? ? ? ? xx (H,T) ? xy (H,T) ? yx (H,T) ? yy (H,T) ? ? ? ? ? ? ? ? J x J y ? ? ? ? . (4.4) This can be further simplified for the devices investigated here, since not all of the components ? ij were found to be independent. The diagonal elements, ? xx and ? yy equal one another: ? xx = ? yy ? ?. This is easily understood due to the fact that the samples were fabricated from (100) Si, and all <100> directions in Si are equivalent [20]. Furthermore, the o?-diagonal components were found to be antisymmetric: ? xy =-? yx .Thisisalsoeasy to interpret, by considering how the Lorentz force (F =qv ? H)actsona charge q moving with velocity v. If we assume q > 0 and motion along the 94 +?x direction, the Lorentz force deflects q towards -?y, in e?ect experiencing an e?ective electric field in that direction. On the other hand, for motion along +?y, the deflection (and corresponding e?ective electric field) is in the +?x direction. Thus I arrive at the final form of the resistivity tensor for ?-doped silicon: ? ? ? ? E x E y ? ? ? ? = ? ? ? ? ? xx (H,T) ? xy (H,T) ?? xy (H,T) ? xx (H,T) ? ? ? ? ? ? ? ? J x J y ? ? ? ? . (4.5) Thus there are only two independent components of the resistivity (conductivity) tensor for ?-layers in Si. 4.2.1 Hall Bar Devices To see how these notions apply to actual device geometries, consider Figure 4.2, which shows a schematic of a typical Hall bar device constructed for use in ?-layer transport studies. The ?current channel? has a width W, and the voltage contacts are spaced a distance L apart. There are essentially only two independent resistance measurements one can perform on such a device. One is the ?longitudinal? resistance, when the induced voltage V WL (for ?weak localization voltage?) is measured between two adjacent contacts on the same side of the device (i.e. parallel to the current bias). The other is the ?transverse? resistance where the induced voltage V H (?Hall voltage?) is measured between contacts directly across the sample from one another (perpendicular to the current bias). Other resistances on the sample will be 95 Figure 4.2: Hall bar measurement schematic. A current bias, I, is passed through the device, and two voltages are measured, as discussed in the text. The weak local- ization voltage, V WL ,ismeasuredparallel to I, the Hall voltage, V H , perpendicular to I. When the voltage contact spacing, L, is equal to the sample?s width, W, the measured longitudinal resistance is the device?s intrinsic resistivity in units of ?/a50. linear combinations of these ?basis? measurements. It should be clear from the figure that V WL scales like V WL ? (L/W)I, whereas V H is independent of L and W (see Appendix A for a derivation). Thus the significant resistances of the device given a rectangular geometry are R WL = (W/L)(V WL /I), and R H =V H /I. If the current channel width W equals the voltage contact spacing L, W/L = 1, and the measured longitudinal resistance is by definition the two-dimensional resistivity: R WL 96 = ?. This two-dimensional ? is often referred to as the sheet resistivity, symbolized by R a50 , and has dimensions of ?/a50 (?ohms per square?), as opposed to ??cm for three dimensional samples. The set of equalities R WL = ? =R a50 will be used interchangeably throughout the remainder of this dissertation. Given this two-dimensional definition of resistivity, we now see that in order to be dimensionally consistent, the current density J must have units of (amperes/meter), as mentioned earlier. 4.2.2 van der Pauw Devices Not all of the samples fabricated in the course of this work used the rectangular geometry of Figure 4.2. Such devices, with arbitrary geometries, are collectively referred to as van der Pauw devices. A general van der Pauw sample is shown in Figure 4.3, and is simply a device with 4 contacts (labeled A, B, C, D) placed on its perimeter. Given a device with such an arbitrary geometry, it is clear that if a current is passed between any two electrodes, the current density will be non-uniform throughout the sample. This is very di?erent from a Hall bar, and given such non-uniformity, one might at first assume that being able to determine the samples conductivity tensor would be a di?cult task. Fortunately, this is not the case, and a conformal mapping exists which translates the resistances measured on these arbitrarily shaped samples to that of ? xx and ? xy as measured on the standard Hall bar device of Figure 4.2. This result, mentioned earlier, was established by L.J. van der 97 Figure 4.3: Arbitrarily shaped van der Pauw device. The electrodes A, B, C, D must be placed on the devices perimeter to apply van der Pauw?s theorem. Pauw in 1958 [72], and has proved extremely useful in this work. To extract ? from such a sample such as the one shown in Figure 4.3 one must measure two appropriate resistances, labeled R AB,CD and R AD,BC ,then numerically solve the following equation for ?: exp( ??R AB,CD ? )+exp( ??R AD,BC ? )=1. (4.6) The notation here is such that R AB,CD refers to a resistance measurement where current is sourced and drained in leads A and B, while voltage is measured between contacts C and D (similarly for R AD,BC ). The key point is that if one is interested in measuring ? xx , the current contacts in both R AB,CD and R AD,BC must be adjacent to one another. To perform a Hall measurement, the current contacts lie ?across? from one another, and the usual relationship between magnetic field H and carrier density n is obtained, i.e. ? xy =R AC,BD = IH/ne [74]. 98 One set of experiments I will discuss in Chapter 5 used square van der Pauw samples, with the contacts placed at the vertices of the square. The appropriate measurement configurations are shown in Figure 4.4. Obviously, such a perfectly symmetrical device should yield very nearly the same resistance for all ? xx measurements, and the same slope for ? xy measurements. Letting R 12,34 =R 13,24 = R and substituting into equation (4.6) yields ? = ?R/ln(2) similarequal 4.532R. Now that the basic notions of 2DEG transport measurements have been established, I shall discuss the systematic details of a typical magnetoconductance investigation. 4.3 Measurement Ultimately, the results of this thesis can be reduced to establishing the interrelationships between four physical quantities: an electric field, E,a magnetic field, H, a current density, J, and a temperature, T. In abstract terms these quantities define the appropriate ?phase-space? for modeling electronic transport in ?-layers, and this work, in a loose sense, establishes the appropriate ?metric? on this (E,H,J,T) space. In what follows I shall frequently refer to this idea. First, I will discuss the various instruments and apparatus which were used in my experiments. 99 (a) (b) Figure 4.4: Square van der Pauw device and appropriate configurations for mea- surement of the weak localization (a) and Hall (b) signals. In a perfectly symmetric device, all the measured weak localization signals will have the same magnitude, and the resistivity is related to these measured resistances (R = V/I) by the relation ? = ?R/ln(2) similarequal 4.532R. 100 4.3.1 Instrumentation Resistance measurements were performed in an Oxford Kelvinox 25 dilution refrigerator which cooled the devices to the low temperatures necessary to observe weak localization. After extensive repairs and modifications (discussed in Appendix C), the lowest achievable temperature of this instrument was T similarequal 130 mK. The temperatures of the 1 Kelvin pot, mixing chamber and still were determined via 4-wire measurements on ruthenium oxide (RuO 2 ) thermometers (produced by Lakeshore cryotronics). These measurements were performed continuously by a Neocera LTC-21 temperature controller, using the calibration curve (resistance vs. temperature) supplied by Lakeshore. Although no definitive measure of the instrument?s cooling power at base temperature was made, it seems reasonable to assume it was at least as large as the 50 ?W specified by Oxford for operation at T = 50 mK. The devices were mounted on a tilting stage in the refrigerator whose orientation was controlled mechanically via a hermetically sealed feedthrough. By rotating this stage, the angle ? between J and H could be placed at an arbitrary angle between 0 ? and 90 ? , as depicted in Figure 4.5. Theprecisevalueof? was determined using the Hall e?ect and the relation for the Hall voltage, V H =IHsin(?)/ne, derived in Appendix A. Electrical connections to the samples were made with coaxial cables via a ?breakout box? at room temperature equipped with 20 BNC connectors. 101 Figure 4.5: Tilting stage diagram. The current density J (proportional to the bias current I) lies in the plane of the sample, and the magnetic field H is fixed along the z-axis. The stage orientation is controlled mechanically via a hermetically sealed feedthrough, and the angle ? is determined through a measurement of the Hall resistance: R H =Hsin(?)/ne. Each of these 20 measurement lines was filtered in two stages, as depicted in Figure 4.6. At room temperature, a ??? section filter with a cuto? frequency of 1 MHz was used. Following this, RC filters with a cuto? frequency of 10 kHz were implemented at the 1 Kelvin pot. Contained in the same dewar as the dilution refrigerator was an Oxford cryomagnetic system. The maximum field this magnet was capable of reaching when immersed in liquid helium (T = 4.2 K) was H =12T.The homogeneity of the field (as specified by Oxford) was 1 part in 1000 over a 1 cm 3 spherical volume, at the center of the magnet bore. The power supply controlling the applied field H was equipped with an analog output voltage 102 Figure 4.6: Filtering schematic. Two stages of electronic filtering were incorporated to isolate the devices from high frequency noise. At room temperature (T similarequal 300 K) alow-pass??? section filter was used, with a cuto? frequency of 1 MHz. At T similarequal 1 Kelvin, a low-pass RC filter was used, with a cuto? frequency of 10 KHz. Twenty such lines comprise the total number of leads available for device measurement at base temperature, T similarequal 130 mK. 103 connection, whose output signal was proportional to H.AsH was varied (usually ?swept? at a constant rate ? H), this output voltage was recorded with a DVM to determine the field applied to the sample. In order to determine a suitable (low-noise) test frequency and investigate whether any anomalous signals were present in the experiment, a Stanford SR785 dynamic signal analyzer was used to measure the spectral density of voltage noise in the measurement configuration. Depending on the voltage signal levels induced in the devices due to current bias and magnetic field configuration (recall that ?(H bardbl )isalwayslessthan?(H ? ) in a given device), the signals were often preamplified and further filtered using battery-powered Stanford SR560 di?erential preamplifiers, with their bandwidth limited to ? 300 Hz. Figure 4.7 shows a typical noise spectrum, obtained by connecting two weak localization voltage contacts to the A and B inputs of the preamp, and feeding the ouput signal into the spectrum analyzer. No appreciable di?erence was observed between this spectrum and one taken with the A and B inputs shorted together. Consequently the noise shown in Figure 4.7 is actually the input referred voltage noise of the preamp itself. The voltages V WL and V H , as well as the bias current I, were measured using Perkin-Elmer digital signal processing (DSP) 7280 lock-in amplifiers. The principal advantage of using lock-in detection for conductance measurements lies in the ability to reduce the noise bandwidth of the measurement, as well rejecting out of band noise, by increasing the 104 Figure 4.7: Noise spectrum from a ?-doped sample. The spectrum shows the voltage noise V N ( in units of V/ ? Hz) as a function of frequency present in the measurement configuration. A typical test frequency used was f = 2?eHz? 17.08 Hz, which was su?ciently far from the low-frequency (1/f) portion of the spectrum to allow measurements with good signal-to-noise ratio (SNR). 105 measurement time constant, ?, and thus increasing the signal-to-noise ratio (SNR) in the experiment. Also equipped with an analog output voltage which was proportional to the signal level at the input of the lock-in, the data from these instruments were also recorded using a DVM. The coaxial cables connecting the device (BNC) contacts to the measurement instruments were configured using the standard ?twisted pair? configuration to minimize noise pickup due to time-varying magnetic fluxes, especially at 60 Hz. The DVMs used were Agilent 34401A 6-1/2 digit multimeters. To record a typical magnetoconductance spectrum, automated data acquisition was performed using a personal computer (PC) with LabView (version 5.1) software. The particular ?VI? (?virtual instrument?) used for recording data was rather simple in that it merely queried the DVMs and temperature controller for their readings at a user-specified frequency. I typically used an acquisition rate of 1 data point every 5 seconds. At a minimum, each data point consisted of four numbers, corresponding to the (E,H,J, T) parameters discussed earlier. The PC was connected by a GPIB cable to an IOtech 488/F bus extender, which converted the GPIB commands (voltage pulse sequences) into photonic signals (light pulses). These photonic commands were then transmitted via fiber optic cable to a second IOtech extender, which was in turn connected to the measurement DVMs via GPIB cables. A schematic of this situation is shown in Figure 4.8. 106 Figure 4.8: Data acquisition schematic. The lock-in amplifier measures the voltage across a resistance, and the analog output of the lock-in is measured by a DVM. GPIB commands from the PC, which controlled the data acquisition, were converted to light pulse signals with the IOtech extender, transmitted via fiber optic cable to another extender, and converted back to GPIB commands to read the DVMs. This arrangement eliminated all electrical connections between the computer and measurement instruments. 107 4.3.2 Grounding Given the electrical noise generated by the PC, the fiber optic cable was used to break all electrical connections with this instrument, eliminating high-frequency noise, cross-talk, ground loops, etc. In addition, several other steps were taken to further isolate the measurement system from unwanted noise sources. Each pumping line connected to the dilution refrigerator was clamped with non-metallic (insulating) centering rings and clamps, which eliminated electrical connections with the pumps associated with the 1 Kelvin pot and the 3 He: 4 He mixture. All instruments were powered through a single ?mains? 120 V AC power line, which was first fed through a Topaz ultra-isolation transformer. The ?earth ground? connection of this transformer was removed, and it?s case ?hard wired? to the earth ground of the magnet power supply. This served as the single point ground for the entire experiment, including the cryostat dewar. To prevent ground loop currents from flowing through the BNC shells of the lock-in input BNCs, the coupling inputs were operated in ?float? mode, which connected the shells of the BNCs to the chassis ground of the lock-ins via 1 k? resistors, establishing a fairly high impedance path to ground. This high impedance would help to minimize any residual issues with ground loops not dealt with by the previous methods. While measurements were being recorded, every e?ort was made to power down all unnecessary instruments (level meters, magnet thermometers, 108 high-powered lasers in the neighboring lab, etc.) in order to operate in as quiet an environment as possible. Although these e?orts were largely successful, I am certain that the system could be further optimized. I now turn to the experimental procedures followed in acquiring magnetoconductance data. 4.3.3 Experimental Procedure The following steps comprise a fairly typical sequence of measurements I performed on ?-doped devices in my work. After the sample was mounted on the tilt stage in the dilution refrigerator and cooled to base temperature, a noise spectrum was recorded to determine a suitable measurement frequency. A typical example of such a measurement is shown in Figure 4.7. As mentioned previously, the noise shown in Figure 4.7 is actually the input referred noise of the amplifier itself (we shall see later that the noise produced by the samples was generally less than this). For whimsical reasons, the test frequency was often chosen to be 2?eHz(similarequal 17.08 Hz). This frequency is su?ciently low that the measurements in this thesis can in essence be considered as DC resistance measurements. After this, the bias current was varied (from approximately 10 nA to 1 ?A) to determine whether or not the V WL vs. I curve was linear (i.e. whether or not the device was ohmic). If the device displayed ohmic behavior, a linear fit to the data was performed, and from the slope of this 109 fit, the H = 0 resistivity of the device was determined. An example of this is shown in figure 4.9. If the curve was nonlinear, the sample generally was not investigated further, due to the di?culty of interpreting such data using the theory of weak localization. Assuming a linear response was observed, the bias current was chosen to be as large as possible in order to maximize the SNR, but still low enough such that no heating was apparent (as observed on the mixing chamber thermometer). Oftentimes I = 1 ?A was found to be a suitable bias current, with heating usually becoming evident in the neighborhood of I similarequal 5 ?A. Next, a large magnetic field (|H|similarequal5 T) was applied to the sample in order to establish the H = H ? configuration. Since the Hall voltage V H in a field H is given by V H =IHsin(?)/ne, it is clear from this expression that V H attains it?s maximum value when ? =90 ? , i.e. when I and H are perpendicular. By gradually rotating the tilt stage and tracking the dependence of V H on the stage position, H = H ? was easily established. At this point meaningful magnetoconductance data for H = H ? could be acquired. The usual way this was done was to first set H = -1 T, sweep the field down to H = 0 while recording data, then set H = +1 T and sweep to 0 again. Thus positive field data and negative field data were recorded as separate files. The reason for ?partitioning? the field in this way was due to the unfortunate fact that the magnet power supply was really a ?unipolar? current source. In other words, H could not be continuously swept through 110 Figure 4.9: Data showing ohmic behavior in an Si:B ?-doped sample. The slope of the line through these data points gives the resistivity of the device, in this case R a50 similarequal 2.5 k?/a50. 111 Figure 4.10: Typical magnetoconductance weak localization data (points) with the- oretical fits (lines) for Si:P ?-layers. The upper curve corresponds to H ? measure- ments, the lower to H bardbl . 112 zero. If one attempted to do so, when changing field polarities an audible ?clunk? could be heard as the power supply mechanically switched the polarity of the magnet leads. This wreaked havoc with the measurements, producing unseemly spikes in the data. This probelm was eliminated by taking the two sets of data separately, and the full magnetoconductance spectra shown throughout this thesis were constructed by ?sewing together? the data for H ? 0, and H ? 0. The sweep rate of the magnetic field ( ? H) was varied to find a rate slow enough to render any heating e?ects (due Faraday induced eddy currents) negligible, but fast enough to allow a quick turnaround in data acquisition. In addition, depending on the time constant (?)ofthelock-inusedinagiven measurement, the sweep rate needed to be su?ciently slow to allow the lock-in to ?track? the voltage signals. Good data was obtained using ? H= 0.025 Tesla/minute, ? = 5 seconds, and was found to be compatible with results for measurements taken with smaller values of ? H. The noise levels in the system were su?ciently low that values of ? greater than 5 seconds were seldom necessary. Typical data for ?(H ? )and?(H bardbl )with|H|?[-1,+1] Tesla are shown in Figure 4.10. After the H ? data had been recorded, the tilt stage was gradually turned until the H = H bardbl configuration was obtained. Again the Hall e?ect (V H = IHsin(?)/ne) made this determination possible. Now, instead of maximizing V H ,theslope of the R H vs. H was used to find the angle ? between I and H. 113 Explicitly, we have R H = Hsin(?) ne . (4.7) Taking the derivative of this expression with respect to H we obtain dR H dH = sin(?) ne (4.8) which is (by definition) the slope of the R H (H)curve.ForH = H ? ,sin(?)= 1, and I denote the slope in this configuration by s 0 (i.e. s 0 = 1/ne). Obviously, we learn the value of the carrier density, n, from this measurement. Upon rotating to a non-perpendicular orientation, the slope of the Hall curve will be reduced because sin(?) is now less than 1. Let this new valueoftheslopebes 1 . It is clear that the two slopes are related to one another by the equality s 1 =s 0 sin(?). Dividing through by s 0 and taking the inverse sine of the ratio s 1 /s 0 , we obtain the desired expression for the stage orientation ?: ? = sin ?1 parenleftbigg s 1 s 0 parenrightbigg . (4.9) This analysis should make it clear that when the sample is in the H = H bardbl configuration, the slope of R H (H)iszero. Figure 4.11 shows several Hall traces taken as the stage was turned to H = H bardbl , along with the appropriate ? extracted from (4.9) for each trace. It is worth noting here that no evidence for the stage ?drifting? over time was observed, so that once it was rotated to a particular position, it stayed there. ?(H bardbl ) data was taken in precisely the same fashion as ?(H ? ), as previously described. A complete data set for both configurations was shown 114 Figure 4.11: Hall data showing the angle ? between J and H, as determined by equation (4.9). For ? = 0, a horizontal line is produced. 115 in Figure 4.10. Once the ?(H ? )and?(H bardbl ) data sets had been recorded, some final temperature-dependent measurements were performed to investigate the influence of the electron-electron interaction in the samples. By applying the maximum amount of power to both the mixing chamber (20 mW) and the still (50 mW), the mixture was removed from the dilution refrigerator and the unit ?warmed up? to slightly above T = 4.2 K. At this point, the powers were reduced to their typical base temperature operating levels (0 for the mixing chamber, ? 1.5 mW for the still), and a fairly low value of ? set on the lock-ins (? ? 1-2 seconds). Data acquisition was begun with T ? 4.2 K, and the refrigerator was cooled in the usual fashion to base temperature. Once base temperature was reached, data taking was stopped, and the process was repeated if additional measurements were desired. An example of such temperature-dependent resistance data is shown in Figure 4.12. The (logarithmic) resistance increase as the temperature is lowered is another characteristic of weak localization, as it is for the electron-electron interaction. Taking such data in the presence of a magnetic field (also shown in the figure) enables one to distinguish between these two phenomena, as I discuss in Chapter 5. One final detail in regard to the resistance measurements discussed in this dissertation will now be made. Although the configurations shown in Figure 4.1 were used to perform measurements, at times the limitations of 116 Figure 4.12: Temperature-dependent resistance data, taken in both zero and non- zero magnetic fields. Analysis of such curves quantifies the magnitude of the electron-electron interaction in the device. 117 Figure 4.13: Resistance bridge schematic. A room temperature potentiometer, R p was adjusted to balance the magnetoresistance of the sample R s at H =0. The application of an H negationslash= 0 throws the bridge o? balance, inducing a voltage ?V in the voltage lock-in. A second lock-in measures the current I through the bridge. the instruments became apparent. In particular, since a DSP lock-in amplifier was used, the finite resolution of the digital-to-analog converters (DACs) on these instruments would occasionally reveal the need for greater measurement sensitivity. (An example of this in a current measurement is shown in Figure 4.16.) The finite resolution of the DACs was especially apparent in samples with low values of R a50 in parallel fields, when the induced change in the resistance was exceedingly small. To overcome this deficiency, a custom-made resistance bridge was designed and constructed to improve the measurements. A schematic of this bridge is shown in Figure 4.13. The current source for 118 the bridge was supplied by driving an AC voltage from the ?current lock-in? through a 500 k? resistor, which was then split along the two branches of the bridge before being recombined and ?drained? (measured) at the lock-in input. At the top of each branch was a 1 % precision 1 M? resistor. Along one branch (in the cryostat) was the sample resistance R s , and along the other was the balancing element of the bridge, a variable resistance in the form of a room temperature potentiometer, R p . The idea behind this arrangement was that the 1 M? resistors at the top of the bridge would balance the current through each branch to almost exactly equal I/2. The voltages across R s and R p (V s and V p , respectively) were preamplified and measured with the ?voltage lock-in?, which was operated in differential mode. The choice of lock-in inputs (A and B) was chosen such that the voltage di?erence between the two signals, ?V, was given by ?V = V p -V s = (I/2)(R p -R s ). At H =0,R p was adjusted until ?V = 0 was obtained, i.e., until the bridge was balanced. Turning on a magnetic field of course causes R s to decrease (due to the negative magnetoresistance associated with weak localization), throwing the bridge out of balance. Since R p is independent of H,forH negationslash= 0, the signal di?erence always satisfied ?V ? 0. Although only changes in the magnetoresistance are relevant for fitting weak localization data, for reference the total magnetoresistance of R s in a field H is given by R s (H)=R p -2(?V/I). After a particular experimental run had been concluded, the mixture was 119 again extracted from the refrigerator, and the dilution unit was allowed to warm up to room temperature. The next step in the experimental investigation involves data reduction and error analysis. Before presenting these results, however, a brief discussion of the errors in the measurements themselves is warranted. 4.4 Measurement Errors This section is motivated by the question, ?What could systematically go wrong with these measurements??. It is important to distinguish this concern from other types of errors a?ecting these experiments, including random errors, such someone (Kenton) bumping the cables during measurements, device-specific anomalies (some of which will be discussed here), or theoretical errors in interpreting the data. Two broad ways of discussing this question are in terms of measurement accuracy (i.e. how close to the ?true? value was the recorded data?), and precision (how reproducible was a given result?). Since each recorded data point corresponds to the specification of a point in the (E,H,J, T) phase space, errors associated with quantifying these physical parameters shall be considered separately. 120 4.4.1 E To begin with E, it is perhaps more convenient to discuss the corresponding voltage, defined by V = - integraltext E?dl, since this is what was actually measured. For the Hall bar device in Figure 4.2, V WL = |E bardbl |L, and V H = |E ? |W, where E bardbl and E ? are the components of E parallel and perpendicular to the current bias. L and W were lithographically defined according to the relation L/W = 1, and measurements of these quantities with tools such scanning electron microscopes (SEM) and atomic force microscopes (AFM) revealed that the feature sizes were generally within 5% of their designed dimensions. Consequently, geometric issues associated with a given sample are not a source of error in a voltage measurement. In addition, even if L/W varied substantially from 1 between di?erent devices, for measurements on a particular device this ratio would be a constant. Thus by determining the actual value of L/W with an SEM or AFM, the data could accordingly be rescaled. To further discuss voltage measurement errors, I shall consider the recorded signal V to be composed of two parts: the signal of interest, V S , and an error signal (?noise?) ?V. Mathematically, V = V S +?V,where?V can be negative or positive. The SNR is defined as SNR = vextendsingle vextendsingle vextendsingle vextendsingle V S ?V vextendsingle vextendsingle vextendsingle vextendsingle 2 (4.10) whence it is clear that the higher the SNR, the ?better? the measurement. Since we are concerned with resistance measurements, V S can be controlled 121 by simply adjusting the device?s current bias. ?V is not as easily manipulated, and several noise sources contribute to its magnitude. These include the Johnson (Nyquist) noise of the device itself, amplifier noise (including digitization errors), o?set signals due to ground loops or a poor common mode rejection ratio (CMRR), instrumentation drifts induced by gradual changes in the room temperature or humidity, etc. The remainder of this section will demonstrate that the SNR in these experiments was so large as to be essentially irrelevant to the conclusions which were drawn from this work. As an experimentalist, this is a nice situation to be in. Consider first the voltage error ?V T due to the thermal noise of the device itself. The largest values of R a50 measured in the course of this work were R a50 similarequal 3k?/a50. The spectral density of the noise produced by such a resistance is given by the well-known Nyquist expression ? 4k B TR a50 ,and assuming V T was measured with ? = 1 second at T = 4.2 K, we have ?V T similarequal 1 nV. This would of course be lower at the base temperature of the refrigerator, and for lower resistance samples. Next, I consider Figure 4.7 to estimate the voltage error, ?V A , due to the preamplifier. From Figure 4.7, I estimate the spectral density of the amplifier noise to be similarequal 3nV/ ? Hz for the typical test frequencies I used of 10 - 20 Hz. Again using ? = 1 second for the lock-in time constant, I find ?V A =3nV. The final voltage error I consider is that due to the finite CMRR of the amplifier, ?V C .Toestimate?V C , I use the specified value of CMRR = 100 122 dB for the 7280 lock-in amplifier. Assuming a 1 ?A current bias through a total common mode resistance of 2k? (including the RC filter resistance of Figure 4.6), the voltage signal applied to both inputs of the lock-in is V 0 =2 mV. Since the common mode error signal is defined [75] by CMRR = 20log 10 (V 0 /?V C ) = 100, we obtain V C =20nV. The total voltage error in these measurements, ?V, is given by summing the separate contributions in quadrature, since they are indendent of one another. That is, ?V = radicalbig (?V T ) 2 +(?V A ) 2 +(?V C ) 2 (4.11) which yields ?V similarequal 21 nV. Thus we see that the finite value of the amplifer?s CMRR dominates the volatge errors in these measurements. Nevertheless, this noise level is extremely small when compared with actual signals. Even for the lowest resistance device measured in this work, R a50 similarequal 250 ?/a50,anda modest bias current of I = 100 nA, I obtain a SNR = (25 ?V/21 nV) 2 similarequal 1.4x10 6 ! Thus, the limitations of these experiments lie in interpreting the results, and not with the voltage measurements themselves. Although this analysis is rather convincing, a variety of other cross-checks were performed to ensure the quality of these measurements. One check included examining the dependence of the measured resistance on the test frequency, f. By varying f from DC (f = 0) to f = 1 kHz, it became obvious that current leakage to ground was a real issue at higher frequencies, presumably due to the reduced impedance of the capacitance in the coaxial 123 cables and RC filters, which varies like 1/fC. This implies that low frequency measurements are easier to interpret, since there is only one signal channel to consider at DC, while there are two for f negationslash= 0 (the ?in-phase? and ?quadrature? signal components). The usual way of establishing the appropriate signal phase was to ?autophase? the lock-in at H =0tobein phase with the current through the device. Nonetheless, even at the low measurement frequencies used in these experiments, the quadrature signal showed non-zero response to the varying magnetic field. No e?ort was made to analyze this component of the signal. Another step which was taken to enhance the quality of these measurements was to judiciously allocate the gain of the DSP lock-in to eliminate digitization errors. The 7280 had two stages of gain: AC gain (which amplified the signal before DSP), and demodulator gain (which amplified the signal using DSP). The largest possible value of AC gain was used to minimize such digitization errors, and it is worth mentioning that when using the bridge of Figure 4.13, no evidence of digitization was observed. With regard to the accuracy of the lock-in?s gain, Perkin-Elmer claims the 7280 gain is accurate to 0.6%. Finally, in order to investigate whether temperature or humidity drifts in the laboratory were an issue in these experiemnts, a long (17 hour) trace was taken of a sample with H = 0, shown in Figure 4.14. This shows that, although some drifts were apparent, the variation of the resistance over this 124 Figure 4.14: Measurement drift over time. Although the measured resistance does vary over time, and shows a slightly increasing trend, the variation is quite low. The maximum resistance over the 17 hour measurement period is R max = 55.868 ?, the minimum R min = 55.859 ?, and are therefore 99.98% consistent. Typical weak localization sweeps occurred on much shorter timescales (? 1 hour) than this trace, and produced much larger signal variations, therefore such drifts are inconsequential to data analysis. 125 interval was extremely small (approximately 2 parts in 10 4 ), and, when compared with the changes induced by application of a magnetic field and the typical sweep times (? 1 hour), are negligible. 4.4.2 H In reality, a direct measurement of H was not performed. The reason for this is that the power supply would source a given current, and I assumed that the field-to-current ratio was accurate. For this system, the conversion factor was 0.1258 (Tesla/ampere). However, because of hysteresis e?ects inherent to all superconducting magnets [76], the field experienced by the sample was in general di?erent from the intended value. To illustrate this point, consider the R(H) data shown in Figure 4.15. The abrupt discontinuity of R near H similarequal 275 gauss can be traced to the presence of the superconductor indium (In) near the sample, which was used as a solder on this device. One of the basic properties of type I superconductors such as In is the existence of a critical magnetic field H c ,belowwhichthemetalis superconducting, above which the metal is ?normal? (non-superconducting). For In, H c = 293 gauss [2], very close to the discontinuity in Figure 4.15. When H < H c , type I superconductors completely expel magnetic flux from their interior (the Meissner effect). On the other hand, weak localization measurements are extremely sensitive to the vectorial nature of the applied H. 126 Figure 4.15: Impact of the Meissner e?ect on the magentoresistance due the use of In solder. The transition from the superconducting to normal state in In occurs at H c = 293 gauss. This distorts the local magnetic field applied to the device, and consequently the measured resistance. 127 Given these two considerations, the explanation of the data in Figure 4.15 is that instead of the device experiencing a gradual change in flux bias, when sweeping through H c the flux bias changes discontinuously. In e?ect, the magnetoresistance of the device acts as an independent magnetometer of the field H. Clearly whether H < H c or H > H c has a profound impact on the local H applied to the sample, and consequently the measured resistance. Assuming that the In near the sample really did have H c = 293 gauss, then the onset of the transition region in Figure 4.15 gives an estimate of the actual H applied to the sample. Examining this figure shows that the discontinutiy occurs at roughly 275 gauss, and is therefore reasonably close to the expected value of 293 gauss. This di?erence is quite small when compared with the typical sweep ranges (+/- 1 Tesla). Incidentally, this analysis suggests that a convenient and highly accurate method of calibrating both H and T in a given low temperature apparatus is through 4-wire measurements on type I superconductors near their critical temperatures and magnetic fields. The homogeneity of the field is another concern in my measurements. Because the magnetoresistance is highly anisotropic, stray perpendicular fields would lead to a larger response than for H strictly parallel to the 2DEG plane. Given that ??B =0always, magnetic field lines are never truly straight: they always have a slight curvature. As mentioned earlier, Oxford claims that the homogeneity at the center of the magnet bore was 128 good to 1 in 1000 over a 1 cm 3 spherical volume. Therefore one way of estimating the influence of this e?ect is to ?scale down? this figure to a typical device size of 50 ?mx50?m. Doing this yields a field homogeneity of (10 ?3 ) 2/3 = 0.01 per cm 2 , and since the device area was (50?m) 2 = 2.5x10 ?5 cm 2 , this suggests that the field was homogeneous to better than 3 parts in 10 7 across each square of the device! This analysis is probably ?too good to be true?, and thus I take the following approach. The superconducting solenoid which was used in these experiments had a length L of approximately 35 cm, and a bore diameter of 52 mm, hence a radius a of 2.6 cm. Given the field/current ratio of 0.1258 T/A of this solenoid, we can find the approximate number of turns per unit length, N, from the equation [78] N = H ? 0 I = 0.1258( T A ) ? 0 . (4.12) Since ? 0 =4?x10 ?7 N/A 2 , this yields N similarequal 10 5 turns/meter. The issue of the field inhomogeneity near the center of a solenoid is treated in chapter 5 of [46], where the following result is obtained for the radial component H r of H, an on-axis distance z and radial distance r from the geometric center of the solenoid: H r = 24? 0 NIa 2 rz L 4 . (4.13) Inserting the values above, and a current corresponding to 1 T (7.949 A), I find H r similarequal 2 gauss. Since the radial field is proportional to the magnet current I, H r will generally be even less than this. When compared with the 129 axial field (1 T), we see that field inhomogeneities were not a concern in these measurements. 4.4.3 J The current density J was determined by the lithographically defined width of the samples, W, and measurement of the current bias, I, with the 7280 lock-in operated as a current amplifier. As mentioned earlier, the dimensions of the sample are relatively unimportant, since their precise values can be independently determined with SEMs or AFMs. Therefore I was primarily concerned with determining I. A measurement of I for |H|?[0,+1] T is shown in Figure 4.16. The resolution limitations of the DACs are strikingly apparent, and the discrete nature of the recorded data points is obviously unphysical. This measurement was performed in a manner analogous to that shown in Figure 4.1, with a 1 M? resistor placed in front of the 1 Volt AC source of the lock-in serving as the current source. The behavior displayed by this graph is easily understood by considering the fact that the entire sample, including the contact wires, etc., display weakly localized behavior. The entire sample has a resistance maximum at H =0,andasthefieldis increased, the entire sample becomes less resistive, therefore a larger current flows through the device. To overcome this, one might be tempted to construct a ?current bridge?, analogous to the resistance bridge of 130 Figure 4.16: Finite resolution of DAC in a bias current measurement. Despite the limitations of this measurement, the current over the full magnetic field range is still constant to 2 parts in 10 4 . Figure 4.13, but such an e?ort is unnecessary, since I is constant to roughly 2 parts in 10 4 . Thus we see that determination of the bias current was not a significant limitation of these measurements. 131 4.4.4 T The most uncertain quantity in these experiments was T, the temperature of the device. Several studies (see [77] and references therein) have demonstrated that, in general, the electronic temperature T e and the phonon (lattice) temperature T p are unequal. In the experiments discussed here, the sample and thermometer were physically separated by ? 2cm. Therefore assuming that T e was equal to the phonon (mixing chamber) temperature T p is inaccurate. However, this is insignificant for most of the considerations discussed in this thesis. The reason for this is that, for the most part, experiments were performed only at constant temperatures (T = 4.2 K and the base temperature of the refrigerator, T similarequal 130 mK). The only time the sample temperature was absolutely certain was at T = 4.2 K, when the inner vacuum chamber of the dilution refrigerator was filled with 3 He exchange gas and the cryostat was filled with liquid 4 He. T e was certainly colder at base temperature, and looking closely at the data in Figure 4.12, it is apparent that the resistance saturates in Si:P devices below T similarequal 200 mK. I also observed such saturation in Si:B devices, but in these samples the onset occurred at much higher temperatures: T similarequal 800 mK. The physical origin of this behavior will be discussed in Chapter 5, but Figure 4.12 suggests that despite T e being unknown at base temperature, the observed saturation makes it plausible that T e was constant during the measurements. Therefore, imprecise knowledge of T was not a significant 132 source of additional (propagated) error in these experiments: the samples were pretty cold regardless. 4.4.5 ? Although I have now discussed the errors associated with the physical variables (E,H,J, T), an additional parameter which is important to the conclusions drawn from these measurements is the angle ? between H and J. At this point it is reasonable to ask how sensitive the stage orientation technique was, or what angular resolution, ??, was achievable? The answer to this is straightforward. Since ? is determined through a measurement of V H ,avoltage signal,the voltage noise in the measurement system, V N , will ultimately limit the determination of ? = 0. A lock-in amplifier operated with a time constant ? will, on average, lead to an uncertainty ?V in a voltage measurement given by ?V=V N / ? ?. Intuitively, this makes sense, since one would expect that by averaging a given signal for a longer time a more precise measurement would be produced. Once again, I use V H =IHsin(?)/ne to derive an expression for ?? given ?V. Since in parallel fields we are concerned with very small angles between I and H (? lessmuch 1), we can expand sin(?) to lowest order in a Taylor series and use the approximation sin(?) similarequal ?. I equate this small angle ? with the resolution limit ??.WhenIandH are as parallel as possible, the ability to 133 distinguish V H from ?V is lost, or the equality ?V = IH?? ne (4.14) holds. Solving this equation for ??, I obtain the desired expression for the angular resolution limit: ?? = ne?V IH = neV N IH ? ? . (4.15) What is interesting here is that the numerator consists of quantities over which the experimentalist has no (or at best, little) control, whereas the denominator is under complete control. Thus, in order to improve (reduce) ??, one should work with a large bias current I in a large magnetic field H, and average the Hall signal for a long time ?.Using?V=5nV,n= 10 14 /cm 2 ,I=1?A, and H = 1 T, leads to ?? =8x10 ?4 radians (similarequal 0.05 ? ). Now that the measurement details of this research have been established, I turn to Chapter 5, and a discussion of my experimental results. 134 Chapter 5 Results The results of my low temperature magnetoconductance measurements on ?-doped Si are presented in this chapter. Although the primary focus will be upon Si:P ?-layers, a limited study of Si:B ?-layers is also presented. The theory of weak localization developed in Chapter 2 will serve as my primary tool for interpreting the data. In addition, I critically examine the various experimental and theoretical approaches incorporated throughout this work for their advantages and limitations. 5.1 Si:P ?-Layers The principal motivation for a detailed study of Si:P ?-layers lies in the potential these structures have in device applications. Some particular examples were mentioned in the Introduction (Chapter 1), and will not be 135 discussed further here. Instead, I shall concentrate on two particular aspects of Si:P ?-layer physics. One study examines the impact a high temperature anneal has on the electronic characteristics of the samples. The other investigates the dependence of the magnetoconductance signals on n, the sample carrier density. I also establish several limitations the present theory of weak localization has with respect to ?-doped Si:P systems. 5.1.1 Di?usivity Measurements The desire to establish weak localization as a tool for measuring the thickness of ?-layers can be traced back to a paper of Mensz and Wheeler [80]. By performing weak localization measurements on Si MOSFETs in perpendicular and parallel magnetic fields, these workers estimated that the rms value of the 2DEG thickness in their samples was similarequal 0.21 nm, i.e. less than a Si monolayer. Being able to achieve this type of resolution in ?-doped samples would far exceed the limitations of other established methods for measuring T , including capacitance-voltage (CV) profiling (chapter 9 of [1]), x-ray di?ractometry [81], transmission electron microscopy (TEM) [34], and SIMS (Secondary Ion Mass Spectrometry) analysis. Each of these techniques has its own advantages and disadvantages, and those of SIMS, the most common approach, were discussed in Chapter 3. I now turn to my experiments using weak localization for measuring T , reported in [79]. 136 Experimental Procedure For my measurements, four distinct 50 ?mx50?m van der Pauw devices were fabricated, as described in Chapter 3. The wafer used for these devices corresponds to the point labeled ?A? in Figure 2.10 (n similarequal 10 14 /cm 2 ), and a SIMS analysis from a sample of this wafer is shown in Figure 5.1 (this is the same plot as Figure 3.5). The reason for going through the somewhat elaborate fitting procedure associated with van der Pauw devices is simple. For measurement of a Hall bar, 6 wires are needed (source, drain, and four voltage contacts), whereas with a van der Pauw device it is the bare minimum of 4. In addition, one lead per device was necessary as an ancillary contact to check for conduction through the substrate (Chapter 3), making 7 contacts necessary for a Hall bar device. As discussed in Chapter 4, the dilution refrigerator used in these experiments had 20 total experimental leads. Thus using Hall bars, realistically only two devices could be measured in a given experimental run (cooldown). In contrast, van der Pauw devices allowed four di?erent samples to be measured in a single cooldown, in e?ect doubling the amount of data obtained. Since the work required to cycle a dilution refrigerator from room temperature to base temperature and back is time-consuming and non-trivial, the use of van der Pauw devices was an obvious choice. After defining the device ?mesa? and cleaning o? the resist mask as described in the device processing recipe of Chapter 3 (steps 10 through 15), 137 Figure 5.1: SIMS data from a sample of wafer A, with n similarequal 10 14 /cm 2 .Alofthe devices used in the di?usivity study were fabricated from this wafer. This SIMS profile corresponds to the cook 0 (unannealed) device, and is the same as shown in Figure 3.5. 138 three of the four samples were subjected to an 850 ? C anneal in a nitrogen atmosphere, performed in a Barnstead Thermaline 79300 tube furnace (not a rapid thermal annealer). The annealing times were for 5, 10 and 15 minutes, and henceforth these devices will be referred to by the time each was annealed as cook 0, cook 5, cook 10 and cook 15. Two considerations led to this choice of annealing times. First, prior experience had shown that if the devices were annealed for too long (40+ minutes), T in these devices would ?saturate? and become indistinguishable from one another. This leads to the second rationale for the chosen annealing times, which was to explore precisely how sensitive this weak localization approach is for measuring T . As will be shown, with an appropriate (and as yet undeveloped) theoretical footing, this technique should prove to be the most precise method of measuring ?-layer thicknesses yet demonstrated. Following the anneals, the fabrication process was completed in a parallel fashion for all of the devices (resuming at step 16, Chapter 3), the only di?erence being I used In solder instead of wire bonds (step 31) to connect wires to the Au contact pads. Thus, aside from statistical variations in the wafer itself, the only di?erence between the devices was annealing time. The samples were then mounted in the dilution refrigerator and measured as described in Chapter 4. 139 Results Before fitting data to the appropriate expressions, a standard symmetrization procedure was followed with weak localization data. This consisted of transforming the data according to: R WL (+/?H)= R WL (+H)+R WL (?H) 2 . (5.1) The e?ect of this was to remove any of the residual asymmetries which are present in all 4-wire magnetoresistance measurements [82] , and fit data only to the symmetric component of the conductivity tensor. This technique is both experimentally and theoretically well-founded, and has been discussed extensively elsewhere [68]. Figure 5.2 shows a plot of typical magnetoresistance data in both ?raw? (unsymmetrized) and symmetrized forms. Clearly, the data is highly symmetric to begin with, and this procedure has a negligible e?ect on the interpretation. Henceforth, all weak localization data presented in this thesis will be in symmetrized form. The reader should thus bear in mind that the plots I present are somewhat redundant, since by symmetrizing I am displaying how the data depends on the magnitude of H, not it?s sign. For completeness, I re-state the equations cited in Chapter 2 which I used for fitting data and calculating T . First, in perpendicular fields, the change in the magnetoconductance, ??(H ? ), is expected to take the form ??(H ? )=( e 2 2? 2 planckover2pi1 )[?( 1 2 + planckover2pi1 4eH ? lscript ? 2 )??( 1 2 + planckover2pi1 2eH ? lscript e 2 )+ln(2( lscript ? lscript e ) 2 )] (5.2) 140 Figure 5.2: Raw (unsymmetrized) weak localization data (solid points) and sym- metrized data (open squares). The measurements are highly symmetric before ap- plying equation (5.1), and hence this procedure has a negligible e?ect on data anal- ysis. 141 where?isthedigammafunction,lscript e the mean free path, and lscript ? the dephasing length. In parallel fields, ??(H bardbl )isgivenby ??(H bardbl )=( e 2 2? 2 planckover2pi1 )ln(1+?H 2 bardbl ) (5.3) requiring the single parameter ? for fitting. The model for T used in [79] assumes that the inequality lscript c lessmuch lscript e lessmuch lscript ? was valid in all of the devices measured, which corresponds to the ?homogeneous broadening? limit described in Chapter 2 and [62]: T = parenleftbigg 1 4? parenrightbigg 1/4 bracketleftBigg parenleftbigg planckover2pi1 elscript ? parenrightbigg 2 parenleftbigg ?lscript e lscript c parenrightbigg bracketrightBigg 1/2 . (5.4) A ?ball park? estimate of lscript c =1/ ? n was used in [79]. I return to this point later when I discuss the systematic errors associated with my work. The data and theoretical fits for my four devices are shown in Figures 5.3 through 5.6. The di?erence between the ?(H ? ) (upper curves) and ?(H bardbl ) (lower curves) configurations is easily understood using the phenomenological ?cross section? argument presented in Chapter 2. When H = H bardbl , the e?ective cross section of the 2DEG is significantly smaller than for H ? . Thus less flux penetrates the electron system, and less of a change in conductance is induced. Furthermore, we see that for |H| = 0.5 T, the signal ?contrast?, ??, decreases with annealing time (?? ? ?(H ? )-?(H bardbl )). Equivalently, the greater ??, the smaller is T of the 2DEG in the device, independent of the model used to interpret the data. I now turn to an 142 Figure 5.3: Cook 0 weak localization data. The best-fit parameters to equations (5.2) and (5.3) are lscript e =30nm,lscript ? = 101 nm, and ? = 0.49/T ?2 . The signal contrast at |H| = +/- 0.5 T is ?? similarequal 1.4. analysis of these results. Before doing so, however, I note that the fitting lines seem to systematically deviate from the data near H =0.Myguessis that this comes from a combination of e?ects I discussed in Chapter 3, including some amount of In shielding (these devices did have In contacts!), being unable to sweep the field continuously through H = 0, and field inhomogeneities across the DIP header e?ecting each device di?erently. In any case, my conclusions are independent of these discrepancies. 143 Figure 5.4: Cook 5 weak localization data. The best-fit parameters are lscript e =33nm, lscript ? = 117 nm, and ? = 5.31/T ?2 . The signal contrast ?? similarequal 1.1. 144 Figure 5.5: Cook 10 weak localization data. The best-fit parameters are lscript e =37 nm, lscript ? = 124 nm, and ? = 7.54/T ?2 . The signal contrast ?? similarequal 0.8. 145 Figure 5.6: Cook 15 weak localization data. The best-fit parameters are lscript e =38 nm, lscript ? = 127 nm, and ? = 11.25/T ?2 . The signal contrast ?? similarequal 0.7. 146 The best-fit parameters extracted from cook 0 through cook 15, along with their associated uncertainties, are shown in Table 5.1. Only the errors in lscript e and lscript ? (as determined by the corresponding fitting routines) are included, since these uncertainties dominate the corresponding error in T . ThemeansquareerrorinT was determined by ?propagating? these errors in the usual way [83], given by (?T ) 2 = parenleftBigg C 2 4lscript e lscript 2 ? parenrightBigg 2 (?lscript e ) 2 + parenleftBigg C 2 lscript e lscript 4 ? parenrightBigg 2 (?lscript ? ) 2 (5.5) where C =(1/4?) 1/4 (planckover2pi1/e) radicalbig ?/lscript c . Examination of Table 5.1 shows several trends, which I shall interpret by considering the e?ect the anneal had on the samples. I will primarily discuss lattice-type defects (e.g. vacancies and interstitials), since it is well-known that the quality of the Si crystal improves with annealing time, ?t, and therefore that these defects are reduced, or ?cooked out?, with annealing [20]. First, I note that ? monotonically decreases with ?t. Since this number is formed by the ratio of two directly measurable physical quantities (? ? V/I), it does not hinge on any particular model for interpretation. The fact that ? decreases with ?t is easily understood, since the scattering of electrons o? defects is the origin of resistivity in the first place, and annealing removes defects. Similar considerations will be invoked when discussing the mean free path, lscript e . No such monotonic trend is observed with n. However, if one arbitrarily (for the time being) ignores the cook 5 data point, n does become a 147 monotonically increasing function of ?t. The reason for pointing this out is that previous measurements, not presented in this thesis, demonstrated this e?ect very clearly, and it is in fact quite common; see [84], for example. Such behavior has to do with the notion of electrical activation, which essentially asserts that even if a P donor is placed in the crystal, it does not necessarily contribute an electron to the conduction band. This phenomenon may come about due to donors ?clustering? about one another during epitaxial growth, with some type of molecular bonding acting to confine the electrons, or perhaps another cause related to defects ?trapping? the electrons. Whatever the origin may be, these e?ects are small (< 10%), which supports the assumption that the wafer was quite homogeneous. Themeanfreepath,lscript e , steadily increases with ?t. Essentially the same argument as was used for ? applies here. As the density of defects is reduced, so are the number of scattering sites which disrupt an electronic trajectory as it propagates through the crystal. The average distance traveled by an electron between scattering events is lscript e , so again this result is intuitively appealing. Perhaps the most interesting result in Table 5.1 is the trend the of the dephasing length, lscript ? , to increase with ?t. Recall that lscript ? is the mean distance an electron can travel through a sample before its phase coherence is destroyed, such as occurs during an inelastic scattering event. The previous arguments related to the defect density require a ?renormalization?, 148 Table 5.1: Experimental results for ?-layer devices annealed (?cooked?) at 850 ? C forintervalsof0to15minutes.Theparameterslscript e and lscript ? were extracted from fits to equation (5.2), ? from fits to equation (5.3). The mean-square thickness of a given ?-layer was estimated using equation (5.4), and is proportional to the parallel field parameter ?. The dominant contribution to the error in T comes from lscript e . sample resistivity electron mean dephasing ? rms ? density free length thickness n path lscript ? T lscript e ?/a50 10 14 /cm 2 nm nm Tesla ?2 nm 30 101 14 cook 0 428.6 1.40 +/- +/- 0.49 +/- 13 3 3 33 117 42 cook 5 300.9 1.39 +/- +/- 5.31 +/- 14 11 10 37 124 51 cook 10 267.9 1.42 +/- +/- 7.54 +/- 13 3 9 38 127 61 cook 15 251.6 1.49 +/- +/- 11.25 +/- 14 9 12 149 so to speak, since the static lattice defects I have been discussing do not alter the electron?s energy in the course of a scattering event, only its momentum. Therefore understanding this result requires more elaborate considerations. As mentioned in Chapter 2, the electron-electron interaction is expected to be the dominant dephasing mechanism in Si:P ?-doped samples (data supporting this assertion will be presented shortly). Theoretical work by various authors [39, 40] has shown that the electron-electron interaction is enhanced by disorder. In brief, the explanation for this enhancement is that defects disrupt the translational symmetry of the lattice, resulting in a reduction of conserved momentum. This in turn increases the rate of inelastic scattering, and thus the increase of lscript ? with ?t can be interpreted as a manifestation of this disorder enhanced e?ect at work. This observation could lead to an interesting series of experiments exploring such behavior, possibly incorporating crystallographic x-ray data to quantify the defect density. The parallel field parameter, ?, also steadily increases with ?t. Given the relation between ? and the expression for T in equation (5.4), i.e. that T? ? ?, such behavior is anticipated. This leads us to consider the final column in Table 5.1, T itself. As expected, T is a monotonically increasing function of ?t. This result is consistent with thermal di?usion of the ?-layer induced by the anneals. To check this, I employ a simple model of the data to estimate the di?usion 150 coe?cient of Si:P at T = 850 ? C. These are standard results, see Shewmon [85], for example. Di?usion Model From an atomistic point of view, as an Si:P sample is heated, each particle in the crystal acquires an rms thermal energy of order k B T (Boltzmann?s constant k B = 1.38x10 ?23 J/T). If the temperature is raised su?ciently high, the atoms become mobile, and perform three-dimensional ?random walks? throughout the lattice. By considering the system?s entropy it is clear that if di?usion is allowed to proceed indefinitely, the steady-state solution will be one of complete homogeneity. That is, if there are N P phosphorous atoms in a sample of crystal with total volume V, then, after di?using for a su?ciently long time, locating a small volume with a density appreciably di?erent from N P /V ?almost never? happens (in the sense of Poincar?e cycles and ergodic theory). So the ?-layer gets wider after annealing, with probability 1. Assuming isotropic di?usion (which is the case for P in Si [20]), these three-dimensional random walks can be factored into the product of three, one-dimensional random walks. Therefore, I now show how one-dimensional random walks can be used to model the results for T (?t) in Table 5.1. I start by considering a single P atom located at the origin (z = 0), and assume that it undergoes a one-dimensional random walk for a time ?t. 151 Then the probability that after this time the atom lies in the infinitesimal interval (z, z + dz) is given by P(z, ?t)dz, where P(z, ?t) is the Gaussian expression [86] P(z,?t)= exp( ?z 2 4D?t ) ? 4?D?t . (5.6) Here D is the diffusion coefficient characterizing the random walk. Note that P(z, ?t) is a normalized probability density ( integraltext ? ?? P(z, ?t)dz = 1), and satisfies the one-dimensional di?usion equation [87] ?P(z,?t) ?t = D ? 2 P(z,?t) ?z 2 (5.7) both of which are essentially mathematical statements that the dopant atoms are conserved during the process of di?usion. The mean square displacement, , an atom undergoes from its original (z = 0) position is given by = integraldisplay ? ?? z 2 P(z,?t)dz (5.8) which yields = 2D?t. This relationship will be used subsequently to define the di?usion length, L D . If I assume a Gaussian-distributed ?-layer with initial square thickness T 2 0 , the e?ect of annealing is to simply add to the distribution (since these are two independent Gaussians [88]). Thus a new Gaussian ?-layer is produced, with square thickness T 2 related to T 0 ,Dand?tas T 2 = T 2 0 +2D?t. (5.9) Accordingto(5.9),ifT 2 is plotted versus ?t, a linear relation should be 152 observed, with slope 2D and intercept T 2 0 . Such a plot is shown in Figure 5.7, with the cook 0 data point omitted from the linear fit. The reason for the omission is that the initial rapid increase in T observed between cook 0 and cook 5 likely corresponds to an enhanced di?usion due to the defects, as mentioned earlier. As ?t is increased, the defects are cooked out, the curve ?settles down? from this initial e?ect, and an accurate estimate of D can be obtained. Fitting a line to the annealed data points, I obtain a slope of 186 nm 2 /min, which converts to D = 3.1x10 ?14 cm 2 /s in conventional di?usivity units. A legitimate question at this point is to ask how well this data agrees with other measurements of D in Si:P? This is somewhat di?cult to quantify, since di?usivity is heavily dependent on the techniques incorporated during sample preparation, and the interpretation of these weak localization results is of course model-dependent. A fairly recent treatment using SIMS analysis [89] claims D = 4.14x10 ?16 cm 2 /s, while values of D over two orders of magnitude higher than this are cited in [90]. Given this spread in values, it is clear that when quoting a measured value for D, more is required than simply specifying the temperature. Therefore I conclude that my work [79] is in accord with other published results, and that weak localization in parallel fields can be used for measurement of D. Before I can unambiguously assert that this technique is better than SIMS for measuring T in ?-layers, however, a discussion of the systematic errors 153 Figure 5.7: Di?usion coe?cient data. The slope of the line yields D = 3.1x10 ?14 cm 2 /s. Note that the intercept of the line lies well above the cook 0 data point, which has been omitted from the fit. This is consistent with transient enhanced di?usion in the annealed devices, as discussed in the text. 154 associated with this approach is necessary. The two categories I will explore are errors associated with the experimental techniques themselves, and those with the theoretical approach I have taken to interpreting the data. Experimental Systematic Errors Several issues with regard to the experimental procedures described previously require a more in-depth analysis to establish that this approach is on a sound footing. Since the systematic errors associated with the resistance measurements were discussed in Chapter 4, I will primarily concentrate on issues pertinent to the samples themselves. These include variances intrinsic to the wafer, non-uniformities in the devices due to unintended processing di?erences, and artifacts associated with the annealing procedure. Although this list is far from exhaustive, it seems likely that these issues dominate the systematic errors involved in this di?usivity study. At this point it should also be mentioned that the techniques associated with magnetic field application cited in [80] are probably better than those used in this work. The authors of [80] were able to apply H ? and H bardbl independently and simultaneously, while I was only able to rotate the tilt stage to some fixed angle ? and then vary |H|. Although the two approaches should be equivalent, the flexibilty of controlling two fields independently would have been preferable. A thorough investigation into the uniformity of the wafer used for these 155 measurements was not carried out, but there are several reasons to discount this as a significant source of error. First, the pieces of the wafer used to construct the devices were taken from the same part of the wafer, near the center. As discussed in Chapter 3, the wafer is rotated during epitaxial growth in an e?ort to eliminate the type of non-uniformities I am considering here. Thus, one would expect that variance in the doping density would appear as a radial e?ect. Since the center of the wafer was essentially at rest during growth, this area should be the most uniform on the entire sample. Second, as shown in Table 5.1, even after the anneals, the density was found to vary by less than 10% from sample to sample. As discussed earlier, given the psuedo-trend of n increasing with ?t and the phenomenon of electrical activation, it seems quite likely that the variance in n was due to the anneals, not intrinsic anomalies in the wafer. Third, given that T increased from cook 0 through cook 15 precisely as expected, there is little reason for believing that this e?ect was from some bizarre coincidence in the wafer precisely corresponding to the annealing time of the device. In short, the data shown in Table 5.1, taken as a whole, essentially eliminates the possibilty that the observations presented here were due to non-uniformities in the wafer. Undoubtedly, some part of the observed results were due to unequal processing techniques among the devices. Small di?erences in device features would lead to a corresponding change in the associated 4-wire measurements. A trivial example is that the farther apart two voltage contacts are spaced, 156 the larger is the resistance between them. Using the same device for the entire series of experiements would eliminate such issues, but was not the method followed here, for two reasons. One was prior experience with this ?recycling? approach. To subject the same device to multiple anneals requires the electrical contacts to be removed and re-established each time. The way this was accomplished originally was by using In solder to make contact directly to the ?-layer, and removing the In by soaking the device in a room-temperature HCl bath. However, as discussed in Chapter 4, In, being a superconductor, expels magnetic flux from its interior when immersed in a field |H| < H c .ForIn,H c = 293 gauss. This unwanted superconducting behavior would manifest itself by shielding portions of the device from the appliedfieldfor|H| < H c , an example of which was shown in Figure 4.15. In light of this, one of the main experimental lessons of this work is that the use of superconductors should be assiduously avoided when performing weak localization measurements. The small amounts of In I used for wiring apparently had little to no e?ect. No such ?discontinuities? were observed when samples using Au-Sb contacts were measured. However, these contacts were ?annealed in? (step 24, Chapter 3), and therefore essentially permanent. Furthermore, at T = 850 ? C, it is a bad idea to anneal Au in Si, given gold?s large di?usivity in Si and the associated trap states in the middle of the Si energy gap [91], which wreak havoc with proper measurements. The second reason for measuring 157 separate devices was mentioned earlier: working with a dilution refrigerator takes time, and instead of doing 4 cooldowns to obtain the data in [79], one (long) cooldown using van der Pauw devices was su?cient. My own belief is that such device-dependent e?ects are irrelevant to the conclusions which I have drawn from these measurements. The reasons for this are twofold: the use of the symmetrization procedure and the van der Pauw theorem. As discussed in [68], all 4-wire measurements are a mixture of the symmetric and antisymmetric components of the conductivity tensor. By symmetrizing the data, one automatically obtains the symmetric component of the magnetoresistance, which is precisely what the weak localization correction to the conductivity, equation (5.2), is calculated for. In addition, no matter how convoluted a devices geometry, van der Pauw?s theorem allows one to calculate the associated resistivity, again the pertinent quantity for the weak localization correction. That is, the theorem relates the resistance measured in a given configuration to the sample?s intrinsic resistivity by a single number, appropriate to that particular device. Once the two appropriate, lead-switching van der Pauw measurements have been performed and ? determined via equation (4.6), any other measured resistance is simply renormalized to equal ?. Thus, for arbitrarily shaped samples the appropriate correction factor is easily calculated (as described in Chapter 4). Finally, since the data will ultimately be used with a least-squares fitting routine, this factor can be further optimized by making 158 it into a fitting parameter and finding the best fit. Since a single number multiplies all magnetoresistance data, the legitimacy of the procedure should be clear. An issue I have glossed over in the above argument is the role the voltage probes play in the measurements. Probes will always be invasive, to some degree influencing the data. My guess is that as long as the contacts are su?ciently ?point-like?, the above arguments should hold. Point-like essentially means that the spatial extent of the probes should be small compared to the perimeter of the device (the e?ect of a contact with finite extent was discussed by van der Pauw [72]). Although potentially interesting and useful, a study of the e?ect of probe size on weak localization data was not carried out in this work. In any case, inspection of the lithographically defined devices throughout the fabrication process indicated samples which were highly uniform. Thus I assert that device-dependent di?erences are insignificant to the results in Table 5.1 when compared with the e?ect of annealing. The systematics of the annealing procedure will be the final experimental errors I consider. Both the temperature and cooking times introduce errors into the results, and this is most likely the dominant error contribution to the estimate of the di?usivity. For Si:P, the well-established relationship between the di?usion coe?cient 159 and temperature takes the Arrhenius form [85, 89] D(T)=D 0 exp parenleftbigg ?? k B T parenrightbigg , (5.10) where D 0 is a constant (with di?usivity units of cm 2 /s) and ? is the activation energy (also constant). This exponential dependence on temperature is the primary culprit for the experimental systematic errors in my measurements. To see this, consider the plot in Figure 5.8, which depicts an idealized annealing process. After being loaded in the furnace, the sample temperature is ?ramped up? linearly to the desired maximum temperature T M (850 ? C here), ?dwells? at T M for the desired time ?t, and is then symmetrically ramped back down to room temperature. However, di?usion happens not only during the cooking step proper (the rectangular area below T M in Figure 5.8) but also as the temperature is ramped up and down (triangular areas). The obvious geometric picture is that the total (square) di?usion length induced by the annealing process is the sum of these three areas (really, a trapezoid), and therefore that the actual di?usion in the cooked devices is larger than quoted in [79], which neglected the contribution of the ramps. This dependence on the ramps will be represented by writing the total induced di?usion length as L 2 D = L 2 a50 +2L 2 triangle , (5.11) where L 2 a50 corresponds to the di?usion during the dwell time ?t, and L 2 triangle to the contribution of the ramps. Since the temperature profile is assumed to be symmetric, 2L 2 triangle must be included in L 2 D . 160 Figure 5.8: Model of annealing procedure. Although the di?usion was only intended to occur for the cooking time ?t at temperature T M , di?usion actually occurs throughout the ramping process while T ? T C . The magnitude of this error is calculated in the text. 161 To estimate the magnitude of this accounting error, I will assume that whenever the sample?s temperature is above a critical temperature T C (arbitrarily defined), D obeys equation (5.10), while below T C ,Dis identically 0. In order to calculate the relevant quantities of interest, Hadara?s result that D = (1.71x10 ?3 )exp(-2.81 eV/k B T) cm 2 /s will be used [89], although it should be stressed that sample preparation methods are often as important as the annealing times and temperatures themselves, and the use of this D is only meant to give an estimate of the error. In fact, this probably sets a lower bound on the error, given the disparity between the estimates of D reported in [89] and [79]. Once again being somewhat arbitrary, I ignore values of D for which the di?usivity is less than 1/100 of D at 850 ? C, which leads to T C similarequal 700 ? C. For a variable D, the di?usion length L 2 triangle must be weighted by an integral over the product of the di?usion coe?cient and the infinitesimal time dt [85]: L 2 triangle =2 integraldisplay ? 0 D(T)dt =2D 0 integraldisplay ? 0 exp parenleftbigg ?? k B T parenrightbigg dt. (5.12) ? is the time interval over which D varies. The exponential dependence on 1/T in this integral requires invoking the incomplete gamma function to obtain a closed form solution, and since this is expressed yet again in terms of an integral, provides little physical insight to the problem. Instead, I have calculated the integral numerically by partitioning the domain of integration into a finite number of subintervals, calculating the corresponding di?usion length for each of these elements, and summing the results to approximate 162 L 2 triangle . Carrying out this computation with a 180 element partition (one for each second of the ramp between 700 and 850 ? C), I obtain L 2 triangle =3.7nm 2 . Therefore, assuming a symmetric ramp profile, the total contribution of the ramps to L 2 a50 is 2L 2 triangle ,or7.4nm 2 . Since all of the devices studied used the same ramp profile but di?erent dwell times, the error introduced by L 2 triangle is larger for shorter anneals. This is clear by forming the ratio L 2 triangle /L 2 a50 for each device, for which one obtains 0.297 for cook 5, 0.149 for cook 10, and 0.099 for cook 15. Thus, the L 2 triangle error is clearly a non-trivial contribution, especially for cook 5, where the error is nearly 30 % of L 2 a50 . To combat this e?ect, fast ramp profiles should be used, which would make the ?annealing trapezoid? I have considered in this analysis more rectangular, and therefore closer to ideal. This also suggests that the resolution limits of the weak localization technique used here could be probed by subjecting the samples to very brief ?heat pulses? (for example, in a rapid thermal annealer), whose duration is systematically reduced until the ?-layer broadening can no longer be resolved. To consider the final aspects of the annealing process experimental errors, I take a much less quantitative approach than for the ramping e?ects. One issue is the fact that, in reality, the true temperature of the sample is not precisely known as it is cooked. No thermometer was near the sample to monitor the temperature, so e?ects such as ?overshooting? T M and fluctuating about T M cannot be ruled out. The anneals were performed in a 163 Barnstead Thermaline 79300 tube furnace, whose total length was approximately 1 meter. Obviously there are substantial temperature gradients in such a processing tool, with the center presumably being hotter than the ends. Although every e?ort was made to place the samples in the exact center of the tube, this procedure was approximate at best, and thus one sample may have actually been annealed at a temperature slightly higher than 850 ? C, another slightly below, etc. In addition, the ramp was not even symmetric, let alone trapezoidal in shape, since the ramp down from T M always took longer than the ramp up. To stress the point one final time, since the di?usivity depends exponentially on 1/T, these phenomena could have played a noticeable e?ect in the reported results, but a quantitative, systematic analysis of these points was not carried out. Despite the considerations discussed here, the experimental results produced by these measurements, shown in Table 5.1 and discussed in detail subsequently, are quite convincing. In particular, the weak localization data in Figures 5.3 through 5.6, showing the signal contrast ?? in these samples monotonically decreasing with ?t, provides strong evidence that experimental systematic errors were not the primary limitation facing the interpretation of this data. The next section of this chapter is devoted to exploring the theoretical interpretation of this data, which we shall see is the limiting factor of this weak localization analysis. 164 Theoretical Systematic Errors Although the weak localization data shown in Figures 5.3 through 5.6 was well fit by equations (5.2) and (5.3), alternatives to this are possible. For example, another common expression for ??(H ? ) which also produces good fits is [53] ??(H ? )=( e 2 2? 2 planckover2pi1 )[?( 1 2 + planckover2pi1 4eH ? lscript ? 2 )?ln( planckover2pi1 2eH ? lscript 2 e )] (5.13) which can be obtained from equation (5.2) by replacing ?(1/2 + planckover2pi1/2eHlscript 2 e ) with its asymptotic form, the natural logarithm. More generally, one might want to take into account all types of scattering (electron-electron, electron-phonon, and electron spin interactions) and fit data using the full expression for the weak localization correction, equation (2.36), mentioned in Chapter 2. The point is that although there will be certain features common to all two-dimensional electron systems (such as the magnetoresistance being anisotropic), there will also be important di?erences particular to each, and a theory specific to weak localization in Si:P ?-layers including band structure e?ects, density dependence, etc., has yet to be developed. Because of this, the results shown in Table 5.1 have theoretical systematic errors associated with them, stemming from the use of imperfect modeling, and it is my assertion that these modeling limitations dominate the errors in this work. I have identified several desiderata for a theoretical model of the weak localization correction ??(H) in Si:P ?-layers, which I shall write as ??(H,?) to emphasize the vectorial nature of the measurement (again, ? is the angle 165 between H and the current density J; see Figure 4.5). In addition to being as specific as possible to the system under study (i.e. accounting for subband e?ects, etc.), these criteria include that ??(H,?) should be a continuous function of ?, have well-defined parameter regimes over which the theory applies, predict crossover behaviors, be consistent with previous results, and incorporate lscript c throughout the analysis, not merely for parallel fields. I shall touch on these points only briefly. For purely physical reasons, if measurements are performed over a finite range of magnetic field (say |H|?[-1,+1] T), ??(H,?) should be a continuous function of ?.GivenafixedH = H 0 , the magnetoconductance is maximal for ? = ?/2, minimal for ? = 0, and should monotonically decrease as ? is reduced from ?/2 to 0. This behavior is shown in Figures 5.3 through 5.6. Thus far theoretical papers, especially [62], have addressed this point by decomposing H into components perpendicular (H ? = |H|sin(?)) and parallel (H bardbl = |H|cos(?)) to the 2DEG, and phenomenologically introducing an additional dephasing rate due to the parallel field: 1/? bardbl ? 1/H 2 bardbl . Although phenomenology is acceptable and ?on the right track?, a first-principles theory would be preferable. That well-defined parameter regimes and predictable crossover behaviors are lacking in the present work is demonstrated by the data shown in Figures 5.9 through 5.13 and the fitting results in Table 5.2. Figure 5.9 captures the essence of the point of view I advocate here. Although quite simple 166 Figure 5.9: Large field Hall data for the cook 0 device. The slope yields a carrier density of n = 1.4x10 14 /cm 2 . 167 (Appendix A), the Hall e?ect is an extremely accurate theory; the measured transverse magnetoresistance is highly linear over the entire accessible magnetic field range in these experiments. The same cannot be said of using equation (5.2) to fit the data shown in Figure 5.10. The longitudinal magnetoconductance shows an almost ?resonant? behavior near H = 0, but instead of ??(H ? ) asymptotically approaching (2e 2 /?h)ln(2lscript ? /lscript e )as predicted by (5.2), the derivative d?/dH instead changes sign near H similarequal 2T, and progressively decreases as the field magnitude is increased. This e?ect is likely due to the classical magnetoresistance (CM) contribution to the conductivity, a well-known phenomenon [2] wherein the resistance of a doped semiconductor increases like |H| 2 for large |H|. Such considerations have been omitted from the majority of weak localization analyses, and being able to predict when the CM begins to dominate based on lscript e , lscript ? ,etc.wouldbe extremely valuable for interpreting data. With regard to a consistent theory of weak localization in Si:P ?-layers, Figures 5.11 and 5.13 immediately show several features that are necessary, namely characteristic magnetic field and temperature dependences. The graph in Figure 5.11 shows two measurements of the longitudinal resistance of a device as the temperature is reduced to the dilution refrigerator base temperature. The upper curve corresponds to zero applied magnetic field, the lower curve to a perpendicular field of 2 T. Except for an approximately constant o?set, these two curves are nearly identical. Such behavior is 168 Figure 5.10: Large field weak localization data. The fact that ?(H) is not a mono- tonically increasing function of |H| shows the limited validity of equation (5.2). The reduction in ?(H)for|H|?2 T is likely due to classical magnetoresistance. 169 expected in a system where dephasing is dominated by the electron-electron interaction [40, 34] (a fit to the predicted logarithmic dependence on T is shown). Thus a theory of localization in these systems should from the beginning consider this interaction as the primary mechanism responsible for the magnetoconductance. Turning to Figure 5.13, the measured change in resistance versus H ? is plotted for a sample at two fixed temperatures, T = 4.2 K and T = 130 mK (a complement of the data in Figure 5.11, where the temperature was varied and the field held constant). Since the width of this line shape gives a measure of the dephasing time (the sharper the curve, the longer the dephasing time), it is reasonable that the higher the temperature data displays a broader line shape (presumably in part due to phonon-induced dephasing). The canonical interpretation of these results is to argue that the dephasing time becomes infinite as T ? 0 according to a power law, i.e. ? ? ? 1/T p . However, if one looks closely at the data corresponding to the lowest temperatures of Figure 5.11, it is evident that the resistance saturates (ceases to change) for T ? 200 mK, implying a concurrent saturation of ? ? . To make this saturation more apparent, the H = 2 T data is plotted on a logarithmic temperature scale in Figure 5.12. Such saturation behavior has been observed in all experimental weak localization studies thus far reported, and a possible explanation in terms of so-called intrinsic decoherence was advanced in [92]. The theoretical fervor which this prompted is still 170 Figure 5.11: Temperature dependent data (constant field). The upper curve (solid points) was recorded with|H|= 0, the lower curve (open squares) with H = 2 T. The fact that the form of curve does not change with the large field applied demonstrates that the electron-electron interaction is responsible for the dephasing in Si:P ?-doped samples. The solid line is a fit to the equation R(T) = R 0 + Aln(BT), with values R 0 = 88.1 ?, A = -0.36 ?, and B = 0.86/K. The jumble of data points for |H| =0 near T = 1 K is due to the refrigerator ?hanging up? (ceasing to cool continuously) at this point as mixture was circulated. 171 Figure 5.12: Data of Figure 5.11 plotted on a logarithmic temperature scale, high- lighting the saturation of R(T) for T ? 200 mK. 172 unfolding at the time of this writing, and remains unresolved. The final theoretical point I consider is the appropriate incorporation of the correlation length, lscript c , throughout the magnetoconductance analysis. Although the influence of lscript c on ??(H bardbl ) has been considered at length in terms of estimating T , the e?ect of lscript c on ??(H ? ) has been neglected. Obviously, since weak localization is ultimately reducible to the Aharanov-Bohm e?ect, and this is itself a manifestation of a geometric phase, the discussion of Chapter 2 makes it clear that the inclusion of lscript c in a calculation of ??(H ? ) is necessary for a complete theory. Furthermore, calculations for systems with multiple subbands, intersubband scattering, etc., should be undertaken for a better understanding of magnetotransport in ?-layers. Without resorting to any particular model of the ?-layer thickness, a reasonable estimate of T in the cook 0 device can be obtained by considering the SIMS data of Figure 5.1 and the Hall data of Figure 5.9 (which yields n =1.4x10 14 /cm 2 ). Since the SIMS data gives a three-dimensional P density versus depth in the sample, while the Hall data gives a two-dimensional electron density, integration of the SIMS profile produces a two-dimensional density. By starting at the peak of the SIMS profile and progressively stepping away from this point (symmetrically in either direction), the number of two-dimensional electrons in each infinitesimal interval is summed until the resultant two-dimensional density equals the measured Hall density, 173 Figure 5.13: Temperature dependent weak localization data. The width of the line shape is narrower at 130 mK, due to the logarithmic dependence of the dephasing length on temperature: lscript ? ? ln(T). 174 n. The MATLAB routine which carries out this computation is shown in Appendix B. The result, which I denote by T S (for ?SIMS?), is T S =7.5nm. This shall serve as an important benchmark for the theoretical models of T I consider below. To obtain an order of magnitude estimate for how theoretical limitations influence the data analysis, consider the results displayed in Table 5.2. Shown are the best-fit results to equations (5.2) and (5.3) for data obtained from the cook 0 device as the fitting range of |H| is reduced from +/- 1 T to +/- 0.1 T. Also shown for each range are estimates for the ?-layer thickness in this device based on the ?uncorrelated? (T ) and ?correlated? (T C )models discussed in Chapter 2. Several features are evident. First, the best-fit values of lscript ? steadily decrease as the range of |H| is reduced, and the same trend occurs for lscript e , although the variation in this case is much greater. Second, the best-fit values of ? are by contrast rather uniform, suggesting that equation (5.3) is a su?ciently accurate model for parallel field data. Lastly, the T and T C estimates are actually quite consistent (T = 2.3 +/- 0.06 nm, T C = 15.8 +/- 0.6 nm), provided one only considers ranges of |H|?0.4 T. For smaller ranges of |H|, the parallel field magnetoconductance changes very little, and the fact that a negative value of ? was found for |H|?0.1 T is a consequence of the data being more noise than signal. Therefore, thickness estimates were excluded in this range. Obviously, the average thickness of the donors in the sample is 175 Table 5.2: Dependence of the parameters lscript ? , lscript e , ?, and consequentlyT , on magnetic field range used in fitting routine. H lscript ? lscript e ? T T c limits T nm nm T ?2 nm nm +/- 1.0 103.2 41.1 0.455 2.2 15.9 +/- 0.9 102.2 39.9 0.467 2.2 16.1 +/- 0.8 101.2 38.6 0.483 2.3 16.2 +/- 0.7 100.2 37.0 0.492 2.3 16.2 +/- 0.6 99.3 35.0 0.488 2.3 15.8 +/- 0.5 98.6 33.2 0.493 2.3 15.6 +/- 0.4 97.8 29.9 0.47 2.3 14.6 +/- 0.3 96.7 3.3 0.37 2.1 4.3 +/- 0.2 96.7 3.0 0.02 0.5 1.0 +/- 0.1 92.7 4.1 -0.8 - - 176 independent of the range of |H|, so the less the fitting parameters vary with the |H| interval, the better. Given the enormous variance between the best-fit values of lscript e , it is clear that equation (5.2) does not provide a reliable estimate of this parameter. Ignoring this for the time being and using the result T S = 7.5 nm, it seems that the ?uncorrelated? thickness model (T = planckover2pi1 ? ?/2elscript ? ) underestimates the ?-layer thickness by roughly a factor of 3.5, while the model corresponding to T C (equation (5.4)) overestimates the thickness by approximately a factor of 2. Even using the estimate for lscript e obtained from n and ? (Appendix A), T C = 11.7 nm, which is still nearly 60% larger than T S . What this analysis shows is that although the parallel field weak localization technique is in principle very powerful, and should ultimately prove to be more sensitive than SIMS, the theory necessary for interpretation of the data is still far from adequate for this purpose. Nevertheless, this work has provided a good starting point for further study. 5.1.2 Density Dependence This subsection of my Si:P ?-layer study is concerned with how the considerations I have discussed depend on the carrier density, n. Originally my hope was to perform the perpendicular and parallel field weak localization measurements on Si:P samples with progressively larger sheet resistivities, apply the analysis previously described, and establish the 177 critical density (n c ) for the metal-insulator transition in two-dimensions. However, I was not able to carry this study to its conclusion, due to my inability to obtain appropriate samples. SIMS data for one successfully measured sample is shown in Figure 5.14. This sample came from wafer ?B? in Figure 2.10. The corresponding Hall data is shown in Figure 5.15 (the linear fit yields n = 6.5x10 13 /cm 2 )andthe base temperature, zero-field resistivity was ? = 625 ?/a50. The corresponding data for perpendicular and parallel field weak localization measurements is shown in Figure 5.16. The fitting parameters established for this device via equations (5.2) and (5.3) are lscript ? =94nm,lscript e =28nm,and? = 1.12/T ?2 .As discussed earlier, a better estimate of lscript e = 20.4 nm can be obtained by combining n and ? as described in Appendix A. Using the uncorrelated estimate of the thickness, I obtain T = 4.3 nm in this sample, while integration of the SIMS plot yields T S = 20.8 nm. I o?er no explanation for this rather large discrepancy. SIMS data from wafer ?C? of Figure 2.10 is shown in Figure 5.17. Hall data for this sample is shown in Figure 5.18. The density in this device was n = 3.7x10 13 /cm 2 , the zero-field resistivity ? = 1063 ?/a50, and combining these parameters yields lscript e = 15.9 nm. The best-fit parameters for the corresponding weak localization data, shown in Figure 5.19, are lscript ? =81nm, ? = 1.6/T 2 . Putting everything together, T = 5.1 nm for this device, while integration of the SIMS profile yields T S = 11.6 nm. 178 Figure 5.14: SIMS data from a sample of wafer B (see Figure 2.10). 179 Figure 5.15: Hall data for a device fabricated from wafer B. The linear fit yields n =6.5x10 13 /cm 2 . 180 Figure 5.16: Weak localization data from the same device as the Hall data of Figure 5.15 (wafer B). The best-fit weak localization parameters for this device were lscript ? = 94 nm, ? = 1.12/T 2 . 181 Figure 5.17: SIMS data from a sample of wafer C (see Figure 2.10) 182 Figure 5.18: Hall data for a device fabricated from wafer C. The linear fit yields n =3.7x10 13 /cm 2 in this device. 183 Figure 5.19: Weak localization data from the same device as the Hall data of Figure 5.18 (wafer C). The best-fit weak localization parameters for this device were lscript ? = 81 nm, ? = 1.6/T ?2 . 184 One additional sample was measured, but only at T = 4.2 K. The reason for the abbreviated investigation of this device is that after measuring the density (n = 2.7x10 13 /cm 2 ) and resistivity (? = 2670 ?/a50), it was clear that this sample was still far from the desired parameter range of R a50 similarequal h/e 2 . Although a cursory perpendicular field measurement showed the expected negative magnetoresistance, instead of performing the standard weak localization measurements and analysis, the dilution refrigerator was cycled to room temperature, and additional samples were sought. As mentioned earlier, despite four di?erent attempts, I was unable to obtain adequate samples, and was forced to conclude this aspect of my investigation. The results of my work with respect to the two-dimensional MIT in Si:P are graphed in Figure 5.20. We see the expected behavior in this figure that as n is reduced, R a50 increases. Furthermore, weak localization was clearly observed in all of these samples, although a thorough investigation was only performed on the three highest density devices. Completing this graph with progressively lower density samples would be my first research goal if I were to continue studying Si:P ?-layers. 5.1.3 Final Remarks and Summary This section concludes my discussion of Si:P ?-layers. The material I present here is somewhat ?sprawling?, and follows no particular pattern. I include it merely to document my own thoughts and perspectives on the 185 Figure 5.20: Metal-insulator transition data. As expected, the samples become more resistive as the carrier (doping) density n is decreased. 186 work I have performed, and to clarify why my density dependence work was truncated. The reason for the di?usivity study was hinted at in Chapter 3, namely to determine how temperature influences T . Being able to precisely quantify both T and how T varies with temperature is significant since this information sets thresholds on the temperatures and times which can be incorporated in various fabrication scenarios. Given a particular ?-doped wafer, it seems reasonable to assume that T throughout the sample is more or less uniform. It is also clear that heating the wafer to a su?ciently high temperature should cause the dopants to di?use from their original positions in the crystal and increase T . Given these considerations it is obvious that if oneweretomeasureT of two devices from the same wafer, the device which was annealed at the same temperature but for a longer time should display a larger value of T , and indeed this is shown by my data, as reported in [79]. The density dependence aspect of this work was concerned with establishing how a weak localization measurement of T depends on n, as well as establishing n c , the crtical density for the metal-insulator transition in two-dimensions. Particularly interesting for study would have been devices with R a50 ? h/e 2 (h/e 2 similarequal 25.8 k?), since this condition is equivalent to k F lscript e ? 1 (appendix A), and the usual assumptions made when calculating ??(H) are invalid. Such an investigation would have allowed the crossover behavior of localization (from weak to strong) to be experimentally established, a 187 useful result for future theoretical work in this area. Four separate attempts were made to obtain appropriately doped wafers, but each time the measured resistivity was either too low or too high to be meaningful. Lightly doped samples were tested, but all showed nonlinear I-V characteristics, and even measuring the Hall coe?cient on these samples proved impossible (thus n could not be determined). The probable explanation for this behavior is that n in these samples was actually below n c , the observed conduction was due to transport through an impurity band, and was consequently non-ohmic. As discussed in Chapter 3, n is controlled by varying the wafer?s exposure time to the P source during growth (lower exposure times correspond to lower doping densities, and therefore higher resistivities). The non-ohmic samples were grown with exposure times of less than 40 seconds, while the low resistivity samples were exposed to the P source for 180 seconds or longer. In a seemingly foolproof e?ort to obtain highly resistive but ohmic wafers, I requested a series of samples with shutter times of 60, 80, 100, 120, 140, 160, and 180 seconds. Unfortunately, this attempt also failed. After roughly one week of trying to understand these results by performing a battery of processing tests, the e?ort proved hopeless and was abandoned. Perhaps the failure to obtain proper samples can be explained by considering the SIMS data shown in Figures 5.1, 5.14, and 5.17. Several features are common to all of these plots. First, all show that the 188 background P level is much higher after opening the shutter (similarequal 3x10 18 /cm 3 ), than before ?-layer growth (similarequal 5x10 17 /cm 3 ). This is obviously undesirable. Second, the ?bumps? in the spectra before the doping peak (i.e. deeper in the sample) likely correspond to the changes in substrate temperature during growth. These features are clearly much worse in samples B and C, which are decidedly ?un-?-like?. Third, although the large surface signal common to all of these spectra is supposedly due to Si:H (see Chapter 3), I found that the surfaces did conduct T = 4.2 K, which is also unwanted. My guess is that the degradation in sample quality I observed and my inability to obtain wafers with specified doping levels had to do with some repairs that Phil made to his MBE system. An interesting idea for future work would be to convolve a ?SIMS potential? with my calculation for band structure. Since a higher doping density corresponds to a deeper potential well, ?inverting? a SIMS spectrum would essentially define the potential energy environment experienced by the electrons in the sample. In other words, I expect that the eigen-energies would be given by an expression roughly of the form E = integraldisplay n(z)E(n(z))dz (5.14) where n(z) is specified by the SIMS data, and E(n(z)) is found from my calculation, given n. This is analogous to density functional theory, discussed in Chapter 2. To summarize my results, the data shown in Figures 5.3 through 5.6, and 189 the corresponding parameters in Table 5.1, demonstrate that the e?ective conduction thickness, T , of the 2DEG associated with a ?-doped layer can be measured using weak localization. Furthermore, all models for T are related to the perpendicular and parallel field weak localization parameters as T? ? ? lscript ? . (5.15) By determining T in several di?erent samples annealed at T = 850 ? C, an estimate of the di?usivity of Si:P has been established at this temperature [79]. Second, the ability to measure T was shown in samples with carrier densities ranging from approximately 10 13 /cm 2 to 10 14 /cm 2 ,andthisweak localization technique works best in very thin ?-layers, since the signal contrast implied by (5.15) is higher in this situation. However, because lscript ? decreases with doping density n, for su?ciently low values of n this technique will ultimately fail, presumably when R a50 similarequal h/e 2 . The influence of the Si:P Bohr radius on this limit has yet to be determined, although since the present theory of weak localization works best for samples with low values of R a50 , this implies a large overlap between the donor wave functions. Finally, temperature and magnetic field dependent resistance measurements suggest that the phase coherence length, lscript ? , is limited by the electron-electron interaction, and my annealing experiments provide evidence for an enhancement of this dephasing due to lattice defects. Although a great deal of experimental knowledge has been gained about Si:P ?-layers in 190 Table 5.3: Results for all measured Si:P delta-doped samples discussed in this work, excluding the ?cooked? devices of Table 5.1. The carrier density n was obtained from a linear fit to Hall data, and R a50 is the zero-field resistivity. lscript e was estimated using equation (A.16), while lscript ? and ? correspond to the best fit values of the data to equations (5.2) and (5.3), respectively. T was estimated using T =(planckover2pi1/2e) ? ?/lscript ? . Weak localization data was not obtained for the sample with n = 2.7x10 13 e ? /cm 2 . wafer n R a50 lscript e lscript ? ? T T S e ? /cm 2 ?/a50 nm nm Tesla ?2 nm nm A 1.4x10 14 429 20.3 101 0.49 2.3 7.5 B 6.5x10 13 625 20.4 94 1.12 3.7 20.8 C 3.7x10 13 1063 15.9 81 1.6 5.1 11.6 D 2.7x10 13 2670 7.4 - - - - 191 the course of these investigations, it is also clear that more theoretical work is needed to more accurately model the data, in particular with respect to the influence the correlation length lscript c has on the magnetoresistance. These observations are contained in the data presented in Table 5.3 for all of the Si:P samples measured here. 5.2 Si:B ?-layers Although the majority of the experimental work I performed was on Si:P ?-layers, the other shallow donor in Si, namely Boron (B), was investigated using the same techniques. Being a dopant from column III of the periodic table, B is an acceptor impurity in Si, and therefore the low temperature conduction characteristics displayed in Si:B samples are due to holes,not electrons. This makes these samples interesting to study in their own right, but having the two sets of data side-by-side is rather striking, demonstrating several di?erences that cannot be explained simply by changing the sign of the charge carriers. An accounting of the MBE growth and device fabrication methods, theoretical background, and results of these experiments now follows. The measurement techniques are exactly the same as for Si:P samples (see Chapter 4). 192 5.2.1 Sample Preparation Epitaxially grown Si:B ?-doped wafers have, to this point, been studied much more intensively than Si:P samples; see, for example [93]. Segregation in these samples is not an issue, as evidenced by the SIMS analysis shown in Figure 5.21 (performed on a piece of the Si:B wafer used in these measurements). Unlike Si:P, the backside of the doping region in this sample appears to be more extended than the front, which is possibly an artifact of the SIMS measurement itself. For this profile, O 2 primary ions were used to bombard the sample, with an incident energy of 3 keV at angle of 52 ? . A standard 50 ?mx50?m Hall bar was constructed for measurement as described in Chapter 3. In order to make ohmic contacts to the B ?-layer, aluminum was evaporated and annealed, instead of the Au-Sb alloy used for Si:P devices. The reason for this choice is purely one of suitable interface chemistry, as trying to use Au-Sb to contact the ?-layer produced insulating samples at low temperatures. This behavior is easily understood when one considers that the Schottky barrier energy is raised when an n-type metallic alloy is brought into contact with a p-type semiconductor [19]. Aside from this contact detail, this sample was lithographically identical to the Si:P devices of the previous section and shown in Figure 4.2. 193 Figure 5.21: SIMS data for an Si:B ?-doped sample. The density of data points was su?ciently small that a line has been drawn through them to guide the eye. 194 5.2.2 Theory Since we are considering hole transport in this section, several theoretical considerations come into play. One is the that the wave function of a hole has a p?wave character, instead of the s-wave type P donors exhibit. Thus the symmetries obeyed by these wave functions will be di?erent, and in turn so will the transport properties of devices fabricated from an Si:B ?-doped wafer. The primary manifestation of this asymmetry is a strong spin-orbit coupling, which leads to a positive magnetoresistance. The appropriate equation for modeling the magnetoconductance for H = H ? is given by [53] ??(H ? )=?( e 2 4? 2 planckover2pi1 )[?( 1 2 + planckover2pi1 4eH ? D? ? )+ln( 4eH ? D? ? planckover2pi1 )]. (5.16) Note that this produces the desired positive magnetoresistance (??(H ? ) < 0 ? H), and the prefactor is (1/2) of that shown in (5.2). As before, instead of working with ? ? Iconsiderlscript ? through the relation lscript ? = radicalbig D? ? . In parallel fields, I will use a slightly modified version of (5.3): ??(H bardbl )=?( e 2 2? 2 planckover2pi1 )ln(1+?H 2 bardbl ) (5.17) where ? is an additional fitting parameter which I have introduced ?by hand? (post hoc). In other words, I have no physical argument justifying the use of ?, except that it?s introduction allows the data to be fit better! 195 Figure 5.22: Hall data for the Si:B ?-doped device. The slope yields a hole density of n = 1.43x10 14 /cm 2 . 5.2.3 Results First, I note that the Hall data in Figure 5.22 yields n = 1.43x10 14 /cm 2 in this sample, a density comparable to the cooked Si:P devices of [79]. The zero-field sheet resistivity, however, is substantially higher: R a50 = 2450 ?/a50. Combining these quantities, I estimate the mean free path lscript e = 3.5 nm in this device. The density of data points in Figure 5.21 was too small to integrate and thus estimate the conduction thickness. 196 The weak localization data shown in Figure 5.23 yields the following results for the best-fit parameters: lscript ? = 44.3 nm, ? = 0.09, and ? = 5.35/T 2 . The fact that ? is significantly di?erent from 1 strongly suggests the need for a better parallel field fitting equation than (5.17). Putting this question aside for future work and plowing ahead with the canonical analysis, I obtain an uncorrelated estimate of T = 17.2 nm for this device. Carrying out the analogous temperature and field dependent measurements on this sample produced several interesting results. First, by measuring the longitudinal resistance as the sample was cooled from T = 4.2 K to base temperature, a very strong saturation of the resistance was observed, setting in near T similarequal 800 mK; see Figure 5.24. Although the physical reason for this behavior is unknown, similar results have been reported previously [94, 95]. Second, performing weak localization magnetic field sweeps at constant temperatures (Figure 5.25) showed that the width of the line shape decreases with T, as one would expect for a temperature-dependent dephasing time. I also note that the longitudinal magnetoconductance is well-fit by equation (5.16) all the way out to |H| = +/- 10 T. The best-fit value of lscript ? for this data is lscript ? = 47.7 nm, within 10% of the value obtained for sweeps between +/- 1 T. Thus another di?erence between Si:B and Si:P samples has been established: evidence of classical magnetoresistance e?ects is harder to discern in Si:B devices. To summarize my results for this limited Si:B ?-layer study, I first note 197 Figure 5.23: Weak localization data for the Si:B sample. Note the negative magnetoconductance, as compared to the positive magnetoconductance of Si:P samples. The best-fit parameters to equations (5.16) and (5.17) are lscript ? = 44.3 nm, ? = 0.09, and ? = 5.35/T 2 .Thefactthat? is substantially di?erent from unity suggests that (5.17) is an inadequate model for the parallel field magnetocon- ductance. 198 Figure 5.24: Si:B temperature dependent data in zero magnetic field. The saturation of this data occurs at a much higher temperature (T similarequal 800 mK) than in Si:P samples (T similarequal 200 mK). The origin of this behavior is unknown. 199 Figure 5.25: Si:B temperature-dependent data. As with Si:P samples, the lower the temperature, the narrower the line shape, due to the logarithmic dependence of lscript ? on temperature. 200 Figure 5.26: Large field sweep for Si:B sample. This data is to be contrasted with that of an Si:P sample (Figure 5.10), where non-monotonic behavior is observed. The line fit to this data yields lscript ? = 47.7 nm, close to that obtained when the field limits were only +/- 1 T (lscript ? = 44.3 nm). 201 that a positive magnetoresistance is observed for all fields and orientations, presumably due to the strong spin-orbit scattering e?ects associated with the p-wave B acceptor impurity. The device was substantially more resistive (? 6x) than its Si:P counterparts of comparable density. The asymmetry between the H ? and H bardbl data suggests that one should be able to measure T using weak localization in these samples, and from the best-fit parameters I estimate T = 17.2 nm. However, the anomalously low value of ? in equation (5.17) suggests that this area of the theory needs further development. Perhaps some interesting results relating to the spin-orbit e?ect in parallel fields are involved. Finally, I note that the R(T) curve saturates at a much higher temperature than in Si:P devices. 202 Chapter 6 Conclusion A brief look at possible directions of future research and a summary of the results I obtained in the course of this work comprise the final chapter of this dissertation. 6.1 Future Directions Several avenues for further work would be worthwhile pursuing. For one, determining the two-dimensional metal-insulator transition densities for both Si:P and Si:B would provide answers to a variety of interesting questions. Is this transition of the Anderson or Mott type? For what value of the carrier density does the canonical theory of weak localization, used throughout this dissertation, collapse? How do holes behave di?erently from electrons in this limit? Can aspects of the two-dimensional scaling theory of localization be 203 observed in ?-doped devices? Can x-ray di?raction studies be used to quantify lattice disorder and therefore probe the electron-electron interaction theory? Do Hall measurements yield additional information on the electron-electron dephasing time, ? e?e ?Howdoesthenoise of these samples change when in the presence of a magnetic field, including perpendicular and parallel to the plane of the ?-layer? The list is extensive, and no doubt will be eventually undertaken. Another aspect of these samples worth exploring is in terms of one ?dimensional transport physics. This includes phenomena such as universal conductance fluctuations, di?erent weak localization characteristics, etc [68]. Simply by reducing the width of the Hall bar to less than the phase coherence length (W lessmuch lscript ? ), the device becomes e?ectively one-dimensional, literally a nanowire. Such an approach would be fairly easy to address from a fabrication standpoint using electron beam lithography, and a wealth of experimental possibilities also exist in this context. In terms of actual device fabrication, constructing silicon single electron transistors (SETs), using either Schottky barriers or electrostatic gates, would be useful from a variety of perspectives. In immediate terms, having an alternative SET design would produce more knowledge on the charge o?set phenomenon, which presently prevents SETs from being used in a current standard [13]. From a long-term perspective, investigating these devices would clarify whether or not their use as controllable charge 204 detectors is feasible. Finally, detailed theoretical calculations of ?-doped silicon need to be realistically addressed. Most importantly for this dissertation, models of magnetoconductance phenomena incorporating the correlation length lscript c suitably defined for ?-layers need to be performed, in order that a competent experimentalist has the modeling necessary for a quantitative comparison with measurements. 6.2 Summary To summarize this dissertation, ?-doping in silicon is technologically important, and Chapter 1 motivated my experiments. Chapter 2 discussed the three theoretical aspects of ?-layers crucial for interpreting the measurements presented in this work, including band structure, weak localization, and Berry phase. In Chapter 3, I discussed the epitaxial growth methods and device fabrication techniques incorporated in these experiments, along with the research of other workers. The measurement conditions and data acquisition approaches were the topics of Chapter 4. The results of my magnetoconductance experiments were presented in Chapter 5, including a study of the di?usivity of Si:P at T = 850 ? C, testing the parallel field weak localization technique in Si:P devices with di?ering carrier densities, and an analogous set of measurements on Si:B ?-layers. Many di?erences were apparent in these samples, most notably that Si:P 205 ?-layers exhibit a negative magnetoresistance, while a positive magnetoresistance is observed in Si:B ?-doped devices, as expected. 6.3 Conclusion Several noteworthy results were achieved in the course of this work. First, by adapting a previously demonstrated experimental technique employed on Si MOSFETs, I have established that a comparison of weak localization signals in perpendicular and parallel magnetic fields enables a precise measurement of the e?ective conduction thickness, T ,ina?-doped Si device. Second, by measuring T in several devices subjected to T = 850 ? C anneals, an estimate of the di?usivity in these epitaxially grown Si:P ?-layers has been produced. Third, by investigating the temperature and magnetic field dependence of these measurements, I have established that the electron-electron interaction is the primary mechanism limiting quantum mechanical phase coherence in Si:P ?-layers, and shown evidence for the enhancement of this e?ect due to lattice disorder, inherent in MBE growth techniques. Fourth, these characteristic weak localization phenomena were observed in numerous samples (four were discussed here), with carrier densities ranging from approximately 10 13 to 10 14 cm ?2 , showing a reasonably broad range of applicability. Fifth, a brief study of Si:B ?-layers suggests that a complementary localization investigation in these structures would be informative and straightforward to carry out. Sixth and finally, 206 several theoretical issues with regard to the interpretation of these measurements have been suggested. In all, this work has been broad in scope, informative in its results, and, on the whole, enjoyable to carry out. These results establish weak localization as a powerful tool in the materials analysis of ultra-thin conducting samples. Although weak localization has long been appreciated from rather esoteric theoretical and experimental perspectives, there are indeed practical applications of this phenomenon, as is evident by the work I have presented here. 207 Appendix A Elementary Transport Theory In this appendix I will derive several useful results concerning 2DEG transport. Although these are standard relations and straightforward to derive, having them all in one appendix is beneficial from a completeness standpoint, as well as establishing equivalent ways of interpreting my measurements. The quantities I consider include the conductivity, ?;the carrier density, n; the mean free path lscript e ; the carrier mobility, ?;andthe density of states, ?(E). A.1 Classical Conductivity Recall that the classical conductivity of a 2DEG, ?, is related to its resistivity ? (or, equivalently, R a50 )by? =1/?. An expression for ? using Newton?s second law, combined with Ohm?s law, can be obtained as follows. 208 First, the application of an electric field E (in the form of a voltage bias) accelerates the electrons of a conducting sample in the direction opposite to E. The electrons (with e?ective mass m ? ), on average, accelerate for a time ? e before experiencing a scattering event in which their momentum is randomized. Therefore, they acquire an average drift velocity, v d ,before scattering. For the time-independent, steady state situation, the acceleration due to E and the momentum relaxation balance each other, i.e., we have the equality F = m ? a = eE? m ? v d ? e =0. (A.1) From this we obtain v d = parenleftBig e? e m ? parenrightBig E. (A.2) On the other hand, Ohm?s law relates the current density J to the applied field E via J = ?E. (A.3) If we assume that the sample has n electrons per unit area, J can also be expressed as J = nev d . (A.4) Substituting (A.2) into (A.4), and equating this with (A.3), we obtain the well-known Drude result for the classical conductivity ? = ne 2 ? e m ? . (A.5) I now discuss how n can be determined with Hall e?ect measurements. Before proceeding, however, I note that even with ? and n experimentally 209 determined (and the electron charge e obviously well-known), we still have incomplete information about the electrons in the sample, since the parameters ? e and m ? are unknown. Other measurements, beyond the scope of this thesis (such as Shubnikov de-Haas oscillations), must be performed to determine these quantities. A.2 Hall E?ect Since the results of this thesis are very sensitive to the precise angle ? between the magnetic field H and the current density J, I now derive the relationship between these quantities and the Hall resistance, R H . To proceed, I consider the square Hall bar geometry of Figure 4.2, and assume, for the moment, that the electric current is carried by particles with positive charge q (I will later take q = e, the electron charge). When q moves with a velocity v in a magnetic field H, it experiences a Lorentz force F which is orthogonal to both these vectors, given by F = q(v?H). (A.6) Since this force is proportional to the cross product of v and H,the magnitude of F depends on the angle ? between these vectors: |F| = q|v||H|sin(?). (A.7) In words, if v and H are parallel (? = 0), no force acts on q, whereas if they are perpendicular (? = ?/2), F is maximal. For perpendicular fields, q is 210 deflected towards one side of the device, determined by applying the ?right-hand rule? to (A.6). Next, I take v to be the drift velocity v d of the electric current in a sample with n charges per unit area: J = nqv d . (A.8) From (A.7), I define an e?ective electric field, E H , associated with the Hall e?ect, given by |E H | = |v d ||H|sin(?). (A.9) Solving (A.8) for |v d | and substituting this into (A.9) I obtain |E H | = |J||H|sin(?) nq . (A.10) This ?deflection? field pushes the electrons towards one side of the device, thereby inducing a charge gradient across the sample, and consequently a transverse Hall voltage,V H .SinceV H =- integraltext E H ?dl, and the line integral is simply the transverse distance W across the width of the sample, V H = |J||H|Wsin(?) nq . (A.11) However, as mentioned in Chapter 4, for a 2DEG, |J| = I/W, where I is the current bias through the device. Incorporating this relation into (A.11) and taking q = e, I arrive at our final expression for the Hall voltage induced in a 2DEG: V H = I |H|sin(?) ne . (A.12) 211 It is worth noting that V H is independent of the device width W, and that V H can naturally be associated with a Hall resistance,R H , by dividing through by the bias current I: R H = |H|sin(?) ne . (A.13) The sin(?) dependence was crucial in this research, since the Hall e?ect was used to determine the angle between the J and H. As implied by (A.13), plotting R H versus |H| in perpendicular fields (sin(?) = 1), produces a line whose slope determines the carrier density, n. The importance of n can be appreciated by noting the frequency with which it appears in the following derivations. A.3 Mean Free Path An extremely useful length scale which can be inferred from magnetoconductance measurements is the mean free path, lscript e . The obvious relationship which holds on a microscopic scale is lscript e = v F ? e (A.14) since the electrons are ?whizzing? around the lattice at the Fermi velocity v F and are scattered, on average, every ? e seconds. Substituting this for ? e in the Drude expression (A.5) for the resistivity yields 1 R a50 = ne 2 lscript e mv F = ne 2 lscript e p F . (A.15) 212 The advantage of doing this is that now lscript e canbeexpressedintermsoftwo measurable quantities: the resistivity R a50 and the density n (via Hall measurements). Substituting p F = planckover2pi1k F = planckover2pi1 ? 2?n into (A.15) and solving for lscript e , I find lscript e = parenleftbigg h R a50 e 2 ? 2?n parenrightbigg = parenleftbigg h e 2 parenrightbiggparenleftbigg 1 R a50 k F parenrightbigg . (A.16) Several features about this expression are noteworthy. For one, the prefactor of the resistance quantum, (h/e 2 ) highlights the quantum mechanical nature of electronic transport. Second, the fact that lscript e scales like (1/R a50 )is physically appealing, since a device with a low value of R a50 should have a long mean free path. Similarly, given that the donor atoms become ionized scattering centers for the electrons to bounce o? of, the fewer of these scatterers are present, the longer lscript e should be, as suggested by the (1/ ? n) dependence. Most relevant for this work, however, is the fact that lscript e ? k ?1 F .As mentioned in Chapter 2, calculations of the weak localization correction to the Drude conductivity, ??(H), are predicated on the so-called ?disorder parameter?, k F lscript e , being large compared to unity: k F lscript e greatermuch 1. What is immediately apparent from (A.16) is that this condition is equivalent to the resistivity being small compared to the resistance quantum: k F lscript e greatermuch 1 ? R a50 lessmuch h e 2 . (A.17) This inequality was satisfied for all of the samples investigated in the course of my research, with the smallest observed value of k F lscript e being approximately 213 8. One of the original goals of this work was to determine precisely for what values of k F lscript e did the ?weakly localized? regime break down, which, according to (A.17) should occur for R a50 similarequal (h/e 2 ) (approximately 25.8 k?). Unfortunately (as mentioned in Chapter 5), I was unable to obtain suitable samples for investigation, and consequently could not carry this program to fruition. A.4 Mobility An alternative measure of ?how well a sample conducts? is the electronic mobility, ?, defined as the proportionality constant between the drift velocity, v d , and the applied electric field E: v d = ?E. (A.18) Of course, we can also express the current density as in (A.3) and (A.4): J = ?E = nev d . (A.19) Substituting (A.18) into (A.19), I find that ne? = ? = 1 ? . (A.20) Therefore our final expression for ? becomes ? = 1 ne? . (A.21) 214 That nothing new has been introduced by considering ? can be seen by solving (A.21) for ? and equating the result with that of (A.5). The result is ? = e? e m ? . (A.22) We are still left with the di?culty of disentangling the elastic scattering time ? e with the e?ective mass m ? , as mentioned in section A.1. A.5 Density of States I present here two calculations for the density of states, ?(E), of a two-dimensional electron gas. The first calculation assumes an isotropic mass (circular dispersion relation), while the second assumes an anisotropic mass (elliptical dispersion). Both of these results are used in my band structure calculation described in Chapter 2. The physical significance of ?(E) is that it describes the number of electrons per unit area in a 2DEG as a function of energy. That is, if there are n 0 electrons in a given subband of a 2DEG with energies between E 1 and E 2 ,then?(E) is defined by n 0 = integraldisplay E 1 E 0 ?(E)dE. (A.23) Equivalently, ?(E) can also be defined by di?erentiating both sides of equation (A.23) with respect to E: ?(E)= dn dE . (A.24) 215 Before carrying out the calculations, I briefly digress to make an important point regarding low temperature electronic transport. Because electrons possess an intrinsic ?spin? angular momentum equal to planckover2pi1/2, they are classified as spin-1/2 fermionic particles. Consequently, they are subject to the Pauli exclusion principle, and obey Fermi-Dirac statistics. This is significant for my work since at low temperatures electronic conduction is entirely due to electrons within roughly k B ToftheFermi energy, E F . This is apparent by considering the low temperature behavior of the Fermi-Dirac distribution function, which gives the probability that an electronic state with energy E is occupied [87]: f(E)= 1 e ?(E?E F ) +1 (A.25) where ? = 1/k B T. For T ? 0, all states with E < E F are occupied, while all states with E > E F are unoccupied. Given the low temperatures associated with the measurements in my experiments (T similarequal 130 mK), the assumption that the transport physics is dominated by electrons very close to E F is clearly a good approximation. A.5.1 Isotropic Mass To derive ?(E) for a 2DEG with an isotropic mass m ? , I consider a system of N such electrons confined to a square of side L at T = 0. According to the Pauli principle, the lowest single-electron energy levels are 216 doubly occupied up to the highest energy in the system, E F . I now establish how E F depends on N and L. In k-space, the electrons will occupy all states out to some radius k F ,the Fermi wavevector, and states with |k| > k F will remain unoccupied. The area of this ?Fermi circle? is ?k 2 F . Due to the boundary conditions that the wavefunction must be zero on the perimeter of the square, the smallest allowed wavevector, k min will have a magnitude of ?/L, and all other allowed wavevectors in the system will be integral multiples of k min . For spin-zero particles, N is simply the ratio of the total volume occupied in k-space (?k 2 F ) divided by the volume occupied by the standing wave mode corresponding to k min ((2?/L) 2 ) [78]. For spin-1/2 electrons this ratio must be multiplied by 2 to take account of the spin degeneracy, and we have N = k 2 F L 2 2? . (A.26) Clearly the density of electrons n = N/L 2 , and thus I obtain the following relation (used previously) between k F and n: k F = ? 2?n. (A.27) Since I am assuming an isotropic dispersion relation (i.e. E = planckover2pi1 2 k 2 /2m ? ), the Fermi energy is given by E F = ?planckover2pi1 2 n m ? . (A.28) Therefore E F in a 2DEG is directly proportional to n. Furthermore, since E F is purely kinetic in nature, we can also identify E F =m ? v 2 F /2 = p 2 F /2m ? , 217 where v F is the Fermi velocity and p F the Fermi momentum of the conduction electrons (p F = planckover2pi1k F ). Given the discussion in Chapter 2 on band structure, it should be noted that this derivation neglects e?ects associated with interactions between the electrons, or band degeneracies due to the lattice symmetry, which reduce the value of k F from that of (A.27) [28]. ?(E) is now easily found by solving equation (A.28) for n and then di?erentiating with respect to E F .Theresult is ?(E)= m ? ?planckover2pi1 2 (A.29) A.5.2 Anisotropic Mass Inowcalculate?(E) for a 2DEG with an elliptical dispersion relation. In other words, I consider a system (such as Si) in which the energy of an electron depends not only on the magnitude of k, but also it?s direction. Representing this dependence in Cartesian coordinates E(k x ,k y )= planckover2pi1 2 k 2 x 2m x + planckover2pi1 2 k 2 y 2m y . (A.30) Generalizing my calculation for the isotropic mass, I assume that N electrons occupy a spatial region of dimensions L x by L y at T = 0. In k-space, the length of the Fermi wave vector will now be directionally dependent, i.e. k F =(k F x ,k F y ) (I use a superscript to keep the notation uncluttered). Thus the occupied area is given by an ellipse with semi-major and semi-minor axes 218 lengths of k F x and k F y . Again equating N with the ratio of occupied k-space to the minimum allowed based on L x and L y , I find N = k F x k F y L x L y 2? (A.31) or n = k F x k F y 2? (A.32) since the sample area A = L x L y . E F can be attained either for k x =0,k y =k F y or k x =k F x ,k y = 0 (i.e. k F x = radicalbig 2m x E F /planckover2pi1 2 ,k F y = radicalbig 2m y E F /planckover2pi1 2 ). Substituting k F x and k F y into (A.32), I obtain n = E F ? m x m y ?planckover2pi1 2 . (A.33) Finally, by di?erentiating (A.33) with respect to E F , I arrive at my final expression for ?(E) in an anisotropic 2DEG: ?(E)= ? m x m y ?planckover2pi1 2 . (A.34) Both of the above results for ?(E) shed light on why quantum Hall e?ect physics is of little concern for 2DEGs in ?-layers. The first quantum Hall plateau occurs when the ratio of the magnetic field to the carrier density equals h/e: |H|/n = h/e. Inserting our canonical density of 10 14 /cm 2 , I find that this requires a field of |H|similarequal4000 Tesla! It is highly improbable that such fields will ever be achieved on earth, and even samples with densities near the metal-insulator transition limit (n similarequal 5x10 12 /cm 2 )require|H|similarequal200 T; still impractical fields. Thus, it seems unlikely that the quantum Hall 219 e?ect will ever be observed in ?-doped Si samples, even if one considers higher index plateaus (since the spacing between these plateaus decreases as the Landau level index increases). 220 Appendix B Data Analysis Routines This appendix contains the MATLAB (version 6.5) code which was used in various aspects of this work. Included are algorithms for fitting data to the perpendicular field correction to the conductivity, ??(H ? ), determining the resistivity of a device using the van der Pauw theorem, and estimating the ?-layer thickness based on SIMS data and Hall measurements. Linear fits to Hall data and logarithmic fits to ??(H bardbl ) data were performed using pre-packaged routines in the Origin (version 6.1) library. B.1 ??(H ? ) The following three sets of code were used to fit perpendicular field weak localization data. The algorithm used by MATLAB was a nonlinear least-squares routine, which, according to MATLAB, incorporated a 221 ?subspace trust region method based on the interior-reflective Newton method? (whatever that means). There were three subroutines to the which comprised the complete fitting algorithm: WL1p3, deltaGWL1p3, and perpfitWL1p3. The reason MATLAB was used for this task is that the digamma function, ?(x), (?psi(x)? in MATLAB) is a built-in function in MATLAB, but not in Origin. Similar sets of code were written for fitting Si:B data, but have been omitted from this section. B.1.1 WL1p3 The first part of the fitting routine, WL1p3 (for ?weak localization 1, 3 parameters?) is simply the MATLAB version 6.5 code for equation (5.2), with fitting parameters p(1), for the o?set conductance (in units of (e 2 /?h)), p(2) for the dephasing length lscript ? (in nm), and p(3) for the mean free path lscript e (in nm)). H obviously stands for the magnetic field value. function Y = WL1p3(H,p) if H==0 Y=p(1); else 222 Y = p(1) + (psi(0.5 + (164.57)/(abs(H)*p(2) 2 )) - psi(0.5 + (329.14)/(abs(H)*p(3) 2 )) + log(2*(p(2)/p(3)) 2 )); end B.1.2 deltaGWL1p3 The second part of the fitting routine, deltaGWL1p3 (for ?di?erence (delta) between weak localization theory and measured data?), ?loaded? the measured data (gm) into the routine, calculated the theoretically expected conductance (gc) using WL1p3 based on the values of the fitting parameters p(i) defined above, and compared the di?erence between the two. function G = deltaGWL1p3(p) gm = load(?C:\ data, etc\ perpSiPdata\ filename.dat?); [M,N]=size(gm); k=1:M; gc=zeros(M,N); gc(k,1)=gm(k,1); 223 for i=1:M gc(i,2)=WL1p3(gc(i,1),p); end for i=1:M x(i)=gm(i,1); G(i)=gm(i,2)-gc(i,2); end B.1.3 perpfitWL1p3 The last part of the routine, perpfitWL1p3 (for ?fitting perpendicular field data to WL1p3?) iterated the fitting procedure until the sum summationtext i (gc(i)-gm(i)) 2 was minimized, which is essentially the same as minimizing ? 2 . The fit was started with the initial parameters p0 and was constrained to remain within the physically reasonable bounds defined by lb and ub. This was carried out using the MATLAB?s built-in non-linear least-squares fitting routine, ?lsqnonlin?. p0 = [100 100 10]; lb = [0,10,1]; 224 ub = [1e5,1e4,1e4]; [p,resnorm] = lsqnonlin(@deltaGp3,p0,lb,ub) B.1.4 deltaLp Unfortunately, the lsqnonlin routine in MATLAB 6.5 only produces the best-fit values for the parameters p(i), and not the uncertainties in these parameters. To overcome this deficiency, separate routines were written to perform this task, and here we show deltaLp (for ?uncertainty in lscript ? ?). The way the uncertainties were determined was to find the best-fit values of the parameters using perpfitWL1p3, gradually increment lscript ? away from its minimum while holding all other values fixed at their best-fit values, calculating a new value of ? 2 for each step, and iterating this procedure until ? 2 was doubled from its minimum value: ? 2 ? 2? 2 . According to [83], this procedure probably overestimates the errors in the fitting parameters, but when compared with the theoretical uncertainties facing the interpretation of these measurements, is relatively unimportant. The routine for finding the error in lscript e is virtually identical to that of lscript ? , and will not be shown here. function f = deltaLp(gm,p) [m,n]=size(gm); 225 for i=1:m H(i)=gm(i,1); gc(i)=WL1p3(H(i),p); end r0=chi2(gm,gc); r=r0; while (r <= 2*r0) p(3)=p(3) + 0.001; gc(i)=WL1p3(H(i),p); r=chi2(gm,gc); end r0 r f=p(3); B.2 van der Pauw Theorem This routine takes two properly measured resistances from a van der Pauw device, R1 and R2, and applies the van der Pauw theorem to 226 determine the resistivity ? (here denoted by R). R1 = input(?Enter the first resistance:?); R2 = input(?Enter the second resistance:?); R=(R1 + R2)/2; x=exp(-pi*R1/R) + exp(-pi*R2/R); if x < 1 while (1-x)>.0001 R= R + .001; x = exp(-pi*R1/R) + exp(-pi*R2/R); end else while (x-1)>.0001 R = R - .001; x = exp(-pi*R1/R) + exp(-pi*R2/R); end end 227 R B.3 SIMS T As discussed in Chapter 5, an estimate of the conduction thickness T can be obtained once given a SIMS profile and a Hall density, n. The way this is accomplished is by integrating the three-dimensional density of the SIMS plot until the resulting two-dimensional density equals that obtained by the Hall e?ect. The integration is started at the maximum value of the density in the SIMS plot, then symmetrically stepping away from the maximum and summing the contribution from each interval. The density of each interval is estimated by averaging the values of the endpoints, while the length of the interval is given by the di?erence in depths. The algorithm is shown below, where the parameter nH is the input Hall density, n2D is the ?running tally? of the integrated SIMS density, and T the corresponding thickness estimate. function F = SIMS(nS,nH) nS=load(?C:\data, etc\perpSiPdata\filename.dat?); nH=1.43e14; [m,n]=size(nS); 228 for i=5:m n(i-4,1)=1e-8*nS(i,1); n(i-4,2)=nS(i,2); end [k,l]=size(n); M=max(n); for i=1:k if (n(i,2)?=M) continue else dM=n(i,1); j=i; end end T=(n(j+1,1)-n(j-1,1))*1e7; 229 n2D=((n(j-1,2)+n(j,2))/2)*(n(j,1)-n(j-1,1)) + ((n(j,2)+n(j+1,2))/2)*(n(j+1,1)-n(j,1)); p=0; while (n2D