Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Silicon quantum electronics

Floris A. Zwanenburg, Andrew S. Dzurak, Andrea Morello, Michelle Y. Simmons, Lloyd C. L. Hollenberg, Gerhard Klimeck, Sven Rogge, Susan N. Coppersmith, and Mark A. Eriksson
Rev. Mod. Phys. 85, 961 – Published 10 July 2013
PDFHTMLExport Citation

Abstract

This review describes recent groundbreaking results in Si, Si/SiGe, and dopant-based quantum dots, and it highlights the remarkable advances in Si-based quantum physics that have occurred in the past few years. This progress has been possible thanks to materials development of Si quantum devices, and the physical understanding of quantum effects in silicon. Recent critical steps include the isolation of single electrons, the observation of spin blockade, and single-shot readout of individual electron spins in both dopants and gated quantum dots in Si. Each of these results has come with physics that was not anticipated from previous work in other material systems. These advances underline the significant progress toward the realization of spin quantum bits in a material with a long spin coherence time, crucial for quantum computation and spintronics.

  • Figure
  • Figure
  • Figure
  • Figure
  • Figure
  • Figure
  • Figure
55 More
  • Received 20 June 2012

DOI:https://doi.org/10.1103/RevModPhys.85.961

Published by the American Physical Society

Authors & Affiliations

Floris A. Zwanenburg*

  • NanoElectronics Group, MESA+ Institute for Nanotechnology, University of Twente, PO Box 217, 7500 AE Enschede, The Netherlands and Centre of Excellence for Quantum Computation and Communication Technology, The University of New South Wales, NSW 2052 Sydney, Australia

Andrew S. Dzurak, Andrea Morello, and Michelle Y. Simmons

  • Centre of Excellence for Quantum Computation and Communication Technology, The University of New South Wales, NSW 2052 Sydney, Australia

Lloyd C. L. Hollenberg

  • Centre of Excellence for Quantum Computation and Communication Technology, University of Melbourne, VIC 3010 Melbourne, Australia

Gerhard Klimeck

  • School of Electrical and Computer Engineering, Birck Nanotechnology Center, Network for Computational Nanotechnology, Purdue University, West Lafayette, Indiana 47907, USA

Sven Rogge

  • Centre of Excellence for Quantum Computation and Communication Technology, The University of New South Wales, NSW 2052 Sydney, Australia and Kavli Institute of Nanoscience, Delft University of Technology, 2628 CJ Delft, The Netherlands

Susan N. Coppersmith and Mark A. Eriksson

  • University of Wisconsin–Madison, Madison, Wisconsin 53706, USA

  • *f.a.zwanenburg@utwente.nl

Article Text (Subscription Required)

Click to Expand

Supplemental Material (Subscription Required)

Click to Expand

References (Subscription Required)

Click to Expand
Issue

Vol. 85, Iss. 3 — July - September 2013

Reuse & Permissions
Access Options
Author publication services for translation and copyediting assistance advertisement

Authorization Required


×

Images

  • Figure 1
    Figure 1
    Combining material and electrostatic confinement to create single-electron transistors. First column: Schematic of dopants, 0D, 1D, and 2D structures. Second column: In the corresponding confinement potentials in x, y, and z directions electron states are occupied up to the Fermi energy EF (dashed gray lines). Occupied and unoccupied electron states are indicated as straight and dashed lines, respectively. Third column: Schematic of the silicon nanostructure integrated into a transport device with source, drain, and gate electrodes. Fourth column: The potential landscape of the single-electron transistor is made up of a potential well which is tunnel coupled to source and drain reservoir and electrostatically coupled to gates which can move the ladder of electrochemical potentials, as described in Sec. 2b.Reuse & Permissions
  • Figure 2
    Figure 2
    Schematic diagrams of the electrochemical potential of a single-electron transistor. (a) There is no available level in the bias window between μS and μD, the electrochemical potentials of the source and the drain, so the electron number is fixed at N due to Coulomb blockade. (b) The μN level aligns with source and drain electrochemical potentials, and the number of electrons alternates between N and N1, resulting in a single-electron tunneling current.Reuse & Permissions
  • Figure 3
    Figure 3
    Zero-bias and finite-bias spectroscopy. (a) Zero-bias conductance G of transport vs gate voltage VG at both TTK (solid line) and TTK (dashed line). In the first regime, the full width at half maximum (FWHM) of the Coulomb peaks corresponds to the level broadening hΓ. In the Kondo regime (TTK), Coulomb blockade is overcome by coherent second-order tunneling processes (see text). (b) Stability diagram showing Coulomb diamonds in differential conductance dI/dVSD vs eVSD and eVG at T=0K. The edges of the diamond-shaped regions correspond to the onset of the current. Diagonal lines of increased conductance emanating from the diamonds indicate transport through excited states. The indicated internal energy scales EC, ΔE, hΓ, and TK define the boundaries between different transport regimes. Cotunneling lines can appear when the applied bias exceeds ΔE (see text). Adapted from 222.Reuse & Permissions
  • Figure 4
    Figure 4
    The five separate transport regimes in a three-terminal quantum device. (a) Schematic depiction of the regimes in which transport through a localized state takes place as a function of the external energy scales kBT and VSD. The transitions between regimes take place on the order of the internal energy scales EC, ΔE, hΓ, and TK. (b) Potential landscape of the three-terminal geometry, where the quantum states and the electrochemical potential of the leads are shown together with kBT, VSD and EC, ΔE.Reuse & Permissions
  • Figure 5
    Figure 5
    Silicon crystal in real and reciprocal space. (a) 3D plot of the unit cell of the bulk silicon crystal in real space, showing the diamond or face-centered-cubic lattice, which has cubic symmetry. (b) Silicon crystal in reciprocal space. Brillouin zone of the silicon crystal lattice. It is the Wigner-Seitz cell of the body-centered-cubic lattice. Γ is the center of the polyhedron. From 74.Reuse & Permissions
  • Figure 6
    Figure 6
    Band structure of bulk silicon. (a) The conduction band has six degenerate minima or valleys at 0.85k0. Results supplied by G. P. Srivastava, University of Exeter. From 74. (b) Zoom-in on the bottom of the conduction band and the top of the valence band (schematic, not exact). The band gap in bulk Si is 1.12 eV at room temperature, increasing to 1.17 eV at 4 K (135). The heavy and light hole bands are degenerate for k=0. The split-off band is separated from the other subbands by the spin-orbit splitting ΔSO of 44 meV.Reuse & Permissions
  • Figure 7
    Figure 7
    Valley splitting of dopants and of quantum dots in silicon quantum wells. (a) For a quantum well, in which a thin silicon layer is sandwiched between two layers of SixGe1x, with x typically 0.250.3, the sixfold valley degeneracy of bulk silicon is broken by the large in-plane tensile strain in the quantum well so that two Γ levels are about 200 meV below the four Δ levels (355). The remaining twofold degeneracy is broken by the confinement in the quantum well and by electric fields, with the resulting valley splitting typically 0.11meV. (b) For phosphorus dopants, strong central-cell corrections near the dopant break the sixfold valley degeneracy of bulk silicon so that the lowest-energy valley state is nondegenerate (except for spin degeneracy), lowered by an energy 11.7 meV. The degeneracies of higher-energy levels are broken by lattice strain and by electric fields.Reuse & Permissions
  • Figure 8
    Figure 8
    Sketch of the two lowest-energy eigenstates in an infinite square well of the two-band model presented in the Supplemental Material 474. The envelopes of the two eigenfunctions are very similar to each other and to the sine behavior obtained in the absence of valley degeneracy; the effects of the valley degeneracy give rise to fast oscillations within this envelope. For a square well, one eigenfunction is symmetric and the other is antisymmetric; the symmetries are different because the fast oscillations have different phases as measured from the quantum well boundaries. This sensitive dependence of valley splitting on the atomic-scale physics near the well boundary is the source of the sensitive dependence of the valley splitting on disorder at the quantum well interfaces.Reuse & Permissions
  • Figure 9
    Figure 9
    Valley-orbit mixing. (a), (b) If the valley splitting EV and orbital level spacing ΔE have very different values, the orbital and valley quantum numbers are well defined and there will be no mixing of orbital and valleylike behavior. (c) When EVΔE the valleys and orbits can hybridize in single-particle levels separated by the valley-orbit splitting EVO.Reuse & Permissions
  • Figure 10
    Figure 10
    Valley-orbit coupling from interface steps. Top: Gray-scale visualization of wave function oscillations in the presence of a perfectly smooth interface, oriented perpendicular to z^. Middle: The relationship between the phase of the wave function oscillations and the interface is different on the two sides of an interface step. When the steps are close together, the phase does not adjust to the individual steps, and the valley splitting is suppressed. Bottom: When steps are far enough apart, the oscillations line up with the interface location on both sides of the steps, which causes the phase of the oscillations to depend on the transverse coordinate. This coupling between the behavior of the wave function in the z direction and in the xy plane, which arises even when the well is atomically thin, is known as valley-orbit coupling.Reuse & Permissions
  • Figure 11
    Figure 11
    A silicon-based nuclear-spin quantum computer. (a) Schematic of Kane’s proposal for a scalable quantum computer in silicon using a linear array of P31 donors in a silicon host. J gates and A gates control, respectively, the exchange interaction J and the wave function, as shown in (b). From 180.Reuse & Permissions
  • Figure 12
    Figure 12
    Relative Stark shift of the contact hyperfine interaction for different donor depths (z) calculated for a uniform field in the z direction. (a) Using the tight-binding approach (258); (b) direct diagonalization in momentum space (442). Agreement in overall trends is reasonable, and for the z=10.86nm case both methods predict ionization at 6MV/m.Reuse & Permissions
  • Figure 13
    Figure 13
    Low-field Stark shift of the hyperfine interaction for momentum space diagonalization (BMB) and tight-binding (TB) methods. (a) Electric field response of hyperfine coupling at various donor depths (BMB and TB). (b) Quadratic (left-hand axis) and linear (right-hand axis) Stark coefficients as a function of donor depth (TB). (c) Shift of the ground state electron distribution (dipole moment) as a function of the electric field (TB). (d) The electric field gradient of the dipole moments as a function of donor depth (TB). From 328.Reuse & Permissions
  • Figure 14
    Figure 14
    J oscillations in the exchange coupling. Calculated exchange coupling between two phosphorus donors in Si (solid lines) and Ge (dashed lines) along high-symmetry directions for the diamond structure. Values appropriate for impurities at substitutional sites are given by the circles (Si) and diamonds (Ge). Off-lattice displacements by 10% of the nearest-neighbor distance lead to the perturbed values indicated by the squares (Si) and crosses (Ge). From 208.Reuse & Permissions
  • Figure 15
    Figure 15
    Smoothing out the exchange oscillations—the exchange coupling J as a function of donor separation along [110]. Top curve: Calculation using the effective-mass wave function. Middle curve: Calculation of J based on wave functions obtained using direct momentum diagonalization over a large basis of Bloch states (BMB) with no core correction of the impurity potential (η=0). Bottom curve: BMB calculation of J with a core correction (η=5.8) that reproduces the donor ground state and valley splitting. Note that the points refer to substitutional sites in the silicon matrix. Although the donor separations are relatively small in this case, the spatial variation of the exchange interaction appears to be significantly damped compared to the effective-mass treatment. All J values are calculated in the Heitler-London approximation. From 442.Reuse & Permissions
  • Figure 16
    Figure 16
    Gate control of the two-donor system. Averaged charge distribution along the interdonor axis for various strengths of the J-gate potential (μ) for the (a) singlet and (b) triplet states (fixed donor separation at 10aB). From 97.Reuse & Permissions
  • Figure 17
    Figure 17
    Band structure of the 1/4 monolayer phosphorus δ-doped layer. (a) The calculation by 321: the solid lines show the band structure without exchange correlation and short-range effects, while the dotted lines show the band structure obtained in the full model. (b) The DFT calculation in a supercell with 200 cladding layers by 56. The plane projected bulk band structure of Si is represented by the gray continuum. The Fermi level is indicated by a horizontal dashed line. From 321, and 56.Reuse & Permissions
  • Figure 18
    Figure 18
    Calculated electronic spectrum of a single-atom transistor. Top left: Calculated energies of the D0 and D ground states (GS) as a function of the applied gate voltage VG. The difference in the energy of these two ground states gives a charging energy of EC46.5meV, which is in excellent agreement with the measurement in this device. Potential profiles between source and drain electrodes calculated for VG=0.45V (top middle) and 0.72 V (bottom left). The calculated orbital probability density of the ground state for the D0 potential (top right) is more localized around the donor than for the D potential (bottom right), which is screened by the bound electron. From 116.Reuse & Permissions
  • Figure 19
    Figure 19
    Self-assembled nanocrystals. (a) STM image of a Ge/Si(001) cluster with a height of 2.8 nm. Scan area is 40×40nm. From 272. (b) Band diagram for a Si/Ge/Si heterostructure, showing the accumulation of holes owing to the valence band offset between Ge and Si. (c) Schematic of a quantum-dot device obtained by contacting a single SiGe nanocrystal to aluminum source or drain electrodes. The heavily doped substrate is used as a backgate for the measurements in (d) where ISD is plotted as a function of VG and VSD. (c), (d) From 183.Reuse & Permissions
  • Figure 20
    Figure 20
    Bottom-up grown nanowires. (a) TEM image of a Si nanowire; crystalline material (the Si core) appears darker than amorphous material (SiOx sheath) in this imaging mode. Scale bar, 10 nm. (b) High-resolution TEM image of the crystalline Si core and amorphous SiOx sheath. The (111) planes (black arrows) are oriented perpendicular to the growth direction (white arrow). (a), (b) Adapted from 275. (c) Stability diagram of a pSi nanowire quantum dot. From 469. (d) SEM image of a nanowire quantum dot with NiSi Schottky contacts. From 472.Reuse & Permissions
  • Figure 21
    Figure 21
    Layer design and corresponding band diagram of a Si/SiGe modulation-doped heterostructure used to form top-gated quantum dots. From 26.Reuse & Permissions
  • Figure 22
    Figure 22
    (a) Scanning electron micrograph of the Schottky gates used to form a gated quantum dot in Si/SiGe. (b) Coulomb diamonds: Conductance of the dot as a function of the voltage VG applied to gates G1 and G2 and of the drain-source voltage VDS. From 25.Reuse & Permissions
  • Figure 23
    Figure 23
    Si-MOS quantum dot with large-area top gate. (a) Cross-sectional schematic, showing two oxide and two gate layers, formed on a silicon substrate. The lower SiO2 layer is thermally grown, while the upper oxide layer is formed using plasma deposition. The large-area upper gate induces a 2DEG at the Si/SiO2 interface, while the lower gates locally deplete the 2DEG to form a quantum dot. (b) Top-view schematic, showing lower depletion gates (black) and induced electron layer (gray). (c) Normalized spacings δ between Coulomb peaks in dot conductance as a function of upper gate voltage. Inset: Raw Coulomb oscillations in dot conductance as a function of upper gate voltage. From 373.Reuse & Permissions
  • Figure 24
    Figure 24
    Si-MOS quantum dot with compact multilayer gate stack. (a) Scanning electron microscope image of device. (b) Cross-sectional schematic, showing three oxide layers and three Al gate layers, formed on a silicon substrate. The SiO2 layer is thermally grown in a high-temperature process, while the thin Al2O3 layers between the gates are formed by low-temperature oxidation of the aluminum. (c) Stability map obtained by plotting differential conductance through the device as a function of source-drain bias VSD and plunger (P) gate voltage VP. The first diamond opens up completely, indicating that the dot has been fully depleted of electrons. (d) Coulomb oscillations as a function of plunger gate voltage VP for the first 23 electrons in the dot. From 239.Reuse & Permissions
  • Figure 25
    Figure 25
    Gated quantum dot formed from a Si/SiGe heterostructure with a global accumulation gate. (a) Cross-sectional view of the heterostructure and the two layers of gates. (b) Top-view SEM image of the gates with a numerical simulation of the electron density superimposed. From 261.Reuse & Permissions
  • Figure 26
    Figure 26
    Multigated quantum dot in etched silicon nanowire. (a) Schematic top view and cross-sectional view of the device. Three lower “wrap-around” gates (LGS, LGC, LGD) are used to form tunnel barriers in an etched silicon nanowire. (b) Top-view scanning electron microscope image of the device before the upper gate is deposited. (c) Equivalent circuit of the device. (d) Coulomb blockade oscillations in device conductance as a function of central gate voltage VLGC when the two outer gates (LGS, LGD) are biased to set each tunnel barrier to G=1μS. Inset: Coulomb oscillations for a range of values of barrier conductance from 20 nS to 8μS. From 119.Reuse & Permissions
  • Figure 27
    Figure 27
    Single-gated quantum dot in etched silicon nanowire. (a) SEM images and (b) cross-sectional schematics taken perpendicular to the nanowire (upper) and along the nanowire (lower). (c) Stability map (Coulomb diamonds) obtained by plotting differential current through the device as a function of source-drain bias Vd and wrap-gate voltage Vg. From 362, and 152.Reuse & Permissions
  • Figure 28
    Figure 28
    Noninvasive charge sensing of a Si/SiGe quantum dot using a quantum point contact (QPC) sensor. (a) SEM device image. (b) (Top) Derivative of the QPC current dIQPC/dVG as a function of gate voltage VG. The peaks correspond to changes in the number of electrons in the dot. (Bottom) Current Idot through the quantum dot as a function of VG. (c) QPC sensor output in the few-electron limit. No further transitions occur for VG<1.68V, indicating an empty quantum dot. From 374.Reuse & Permissions
  • Figure 29
    Figure 29
    Noninvasive charge sensing of a Si-MOS quantum dot using a single-electron transistor (SET) sensor. (a) SEM device image, showing a Si-MOS SET sensor (upper device) that is capacitively coupled to a Si-MOS quantum dot (lower device). (b) Transport current ID through the quantum dot shows Coulomb peaks as a function of dot plunger gate voltage VPD. The changing potential on the dot is detected by monitoring the uncompensated current IS through the SET sensor, which shows charge transfer events superimposed on a rising background, due to the coupling of the SET to VPD. This background can be largely removed by adding a linear correction (fixed compensation) to the SET gate voltage VPS, and then further enhanced by plotting the derivative dIS/dVPD. From 462.Reuse & Permissions
  • Figure 30
    Figure 30
    Gate design enabling few-electron occupation. The gate design in (a) is a natural way to form a quantum dot tunnel coupled to two reservoirs, as shown by the arrows. As the dot becomes smaller, however, it is very difficult to maintain a high tunnel rate to both reservoirs. The gate design in (b), based on Fig. 1 of 58, enables a small dot to be coupled to both reservoirs.Reuse & Permissions
  • Figure 31
    Figure 31
    Schematic diagram of a few-electron quantum dot formed from a Si/SiGe heterostructure with a double quantum well and an accumulation gate contacted by an air bridge. Inset: SEM micrograph of the gate region of a corresponding device. From 34.Reuse & Permissions
  • Figure 32
    Figure 32
    Transport data showing the last hole in a Si nanowire-based quantum dot. Inset: SEM image of the device showing the NiSi contacts and a Cr/Au side gate device. From 473.Reuse & Permissions
  • Figure 33
    Figure 33
    Excited state magnetospectroscopy in Si quantum dots. (a) Zeeman splitting at the 0-1 and 1-2 transition in a few-hole Si nanowire quantum dot; (b) the corresponding magnetic field dependence of the Zeeman energy. From 473. (c) Anisotropic g factors in SiGe nanocrystals; (d) the corresponding excited-state magnetospectroscopy. From 183.Reuse & Permissions
  • Figure 34
    Figure 34
    Ground state magnetospectroscopy. Three examples of even-odd hole spin filling. (a) A many-hole Si nanowire quantum dot (469); (b) a few-hole Si nanowire quantum dot (473); (c) a many-hole Ge/Si nanowire quantum dot (336).Reuse & Permissions
  • Figure 35
    Figure 35
    Spin filling in valleys in a planar MOS Si quantum dot. (a) Magnetospectroscopy of the first two electrons entering the quantum dot. The circle 2a marks a kink in the second Coulomb peak at 0.86T. The arrows in the boxes (VO1 for valley orbit 1 and VO2 for valley orbit 2) represent the spin filling of electrons in the quantum dot. (b) For B<0.86T, the first two electrons fill with opposite spins in the same valley-orbit level (top panel). The Zeeman energy at the kink is equal to the valley-orbit splitting (0.10 meV). From 239.Reuse & Permissions
  • Figure 36
    Figure 36
    Schematic stability diagrams for a double dot system. Maps are shown for (a) small, (b) intermediate, and (c) large interdot couplings. The equilibrium charge on each dot in each domain is denoted by (N1, N2). (e) Region within the dotted square of (b), corresponding to the unit cell of the double dot stability diagram at finite-bias voltage. The solid lines separate the charge domains. Classically, the regions of the stability diagram where current flows are given by the gray triangles. From 420.Reuse & Permissions
  • Figure 37
    Figure 37
    Evolution from a single dot to a double quantum dot in a gated silicon nanowire device. (a) Equivalent circuit. (b)–(e) Contour plots of the drain current as a function of the outer barrier gate voltages VLGS and VLGD. The central barrier gate voltages used were (b) VLGC=0.75, (c) 1.13, (d) 1.18, and (e) 1.284V. From 119.Reuse & Permissions
  • Figure 38
    Figure 38
    Gate tunable double quantum dots. (a) SEM image of a Ge/Si nanowire-based hole quantum dot. The Ge/Si nanowire at top (white in image) is gated by metal gates to form a double dot. (b), (c) Charge stability maps of the conductance as a function of plunger gate voltages. (d) SEM image of an electron quantum dot defined by electrostatic top gates in a Si/SiGe heterostructure. (e) Charge-sensing measurement showing the difference in the charge detection signal from the dot farthest from the QPC (4 small steps in IQPC) and the dot closest to the QPC (single large step) as a function of gate voltage. (e) Two-dimensional plot of the charge-sensing current showing the sequential addition of electrons to the left and right dots. (a)–(c) From 160. (d)–(f) From 375.Reuse & Permissions
  • Figure 39
    Figure 39
    Bias spectroscopy of silicon double quantum dots. (a) Stability map with a source-drain bias VSD=1mV for a silicon nanowire double dot, depicted in Fig. 26, obtained by plotting source-drain current I as a function of two barrier gate voltages. The triple points have clearly evolved into bias triangles. (b) Bias triangles for two triple points at VSD=1mV, obtained in a Si-MOS double dot. (c) Line trace of ISD, taken along arrow in (b), showing resonances corresponding to excited states in the double dot. From 241, and 238.Reuse & Permissions
  • Figure 40
    Figure 40
    Single-electron occupancy in a Si/SiGe double quantum dot. (a) SEM of the device. (b) Charge stability map of the double dot, obtained by plotting the QPC charge sensor output as a function of the control gate voltages VL and VR. The charge configurations (n,m) are marked, showing depletion to the (0, 0) state. From 409.Reuse & Permissions
  • Figure 41
    Figure 41
    Pauli spin blockade in a silicon MOS double quantum dot. (a) SEM image; (b) cross-sectional schematic of the Si-MOS device. Gates L1 and L2 induce electron reservoirs at the Si/SiO2 interface, while barrier gates B1B3 define the double dot potential. Plunger gates P1 and P2 control the occupancy of each dot. (c), (d) Current ISD as a function of VP1 and VP2 for B=0T. (c) For VSD=+2.5mV, the ground state and excited states of a full bias triangle are observed. The current flows freely at the S(0,2)S(1,1) transition, as illustrated in the box marked by the dot. (b) The same configuration at VSD=2.5mV. Here the current between the singlet and triplet states is fully suppressed by spin blockade (box marked by star). (e) The measured singlet-triplet splitting ΔST, plotted as a function of magnetic field B. From 221.Reuse & Permissions
  • Figure 42
    Figure 42
    Spin blockade and lifetime-enhanced transport in a Si/SiGe double quantum dot. (a) Measured and (b) schematic charge stability map of current I through the double dot with a source-drain bias of VSD=+0.2mV. The dotted trapezoids in (a) and (b) mark the zero current regions due to spin blockade, as depicted in the schematics in (c). (d) Measured and (e) schematic charge stability map of current I with a source-drain bias of VSD=0.3mV. In this bias direction there is no blockade and current flows throughout the entire bias triangle; however, additional tails are observed due to lifetime-enhanced transport, as depicted schematically in (f) and described in the text. From 363.Reuse & Permissions
  • Figure 43
    Figure 43
    Conductance in micron-scale silicon MOSFETs. (a) Typical low-temperature conductance pattern of a 1980s generation MOSFET around the threshold regime. The strongly oscillating but chaotic pattern that appears at low temperature is associated with localized states in the channel region. (b) Schematic representation of the three major conduction mechanisms through the channel. From 105.Reuse & Permissions
  • Figure 44
    Figure 44
    The importance of discrete dopants in nanoscale MOSFETs. (a) The transition from continuously ionized dopant charge and smooth boundaries and interfaces to (b) a 4-nm MOSFET where there are less than 10 Si atoms along the channel. From 13.Reuse & Permissions
  • Figure 45
    Figure 45
    Transport through dopants ion implanted in a nano-FET. (a) Schematic of a nano-FET where roughly three donors have been implanted into the 50×30nm active area of the device. (b) The stability diagram showing the differential conductance as a function of the barrier gate and dc source-drain bias, highlighting the resonant tunneling peaks a1, b1, and c1 of the three donors. From 403.Reuse & Permissions
  • Figure 46
    Figure 46
    Three examples of device layouts that illustrate different transport regimes for the detection of a single dopant. (a) Capacitive coupling to the channel which leads to a modification of the channel current due to the charge state of a dopant. (b) Tunneling through a dopant in the access region in series with transport through the channel. (c) Direct tunneling through a dopant in the channel in the subthreshold regime. From (a) 297, (b) 151, and (c) 362.Reuse & Permissions
  • Figure 47
    Figure 47
    Three-dopant transport regimes in a transistor geometry. (a) An example of the dopant detection regime based on the capacitive coupling of the channel for an undoped (left) and doped (right) double-gate sample. The signature of a single acceptor charging event is evident in the doped sample. From 297. (b) An example of the second regime where the dopant is in the barrier of the access region in series with a quantum dot. The top line represents the room temperature FET characteristics and the line below the low-temperature Coulomb peaks From 151. (c) The third regime with direct transport through a dopant in the subthreshold limit. From 362.Reuse & Permissions
  • Figure 48
    Figure 48
    Direct tunneling through a dopant in a short-channel FET. (a) Illustration of a Monte Carlo simulation of the doping profile in a 20 nm channel where some dopants diffused into the channel region from the source and drain. (b) The dashed curve shows the current averaged over many devices where the black line indicates the threshold. Two devices show a drastically lower threshold linked to resonant transport at low temperature as indicated in (c) for the device with the lowest Vth. These data show the clear connection between the low threshold of these devices at room temperature and the resonant transport at low temperature, both mediated by a single dopant. From 309.Reuse & Permissions
  • Figure 49
    Figure 49
    Few-electron quantum dot. (a) An STM image of the central device region of a few-electron single-crystal quantum dot acquired during hydrogen lithography, showing a four terminal device with source (S), drain (D), and two in-plane gates (G1, G2). The bright regions correspond to areas where phosphorus donors will be incorporated. (b) A close-up showing the central quantum dot containing 6±3 donors. (c) Stability diagram showing the conductance dI/dVSD through the dot as a function of gate voltage VG and bias voltage VSD. (d) A close-up of the transition [white square in (c)] reveals a high density of conduction resonances with an average energy spacing of 100μeV. (e) The sixfold degeneracy of the conduction band minima of bulk silicon is lifted by confining the electrons vertically to two dimensions and is then split again by abrupt, lateral confinement. From 115.Reuse & Permissions
  • Figure 50
    Figure 50
    A single-atom transistor. (a) 3D perspective STM image of a hydrogenated silicon surface. Phosphorus will incorporate in the bright shaded regions selectively desorbed with an STM tip to form electrical leads to a single phosphorus atom patterned precisely in the center. (b) The source (S), drain (D), and two gate leads (G1, G2) to the central donor, which is incorporated into the dotted square region. (c) The electronic spectrum of the single-atom transistor, showing the drain current ISD as a function of source-drain bias VSD and gate voltage VG applied to both gates. (d) The differential conductance dISD/dVSD as a function of VSD and VG in the region of the D0 diamond shown in (c). (e) A comparison of the potential profile between the source and drain electrodes in this device (straight line) to an isolated bulk phosphorus donor (dashed line), where the D0 state resides 45.6 meV below Ecb. In contrast, the D0 state in the single-atom transistor resides closer to the top of the potential barrier. From 116.Reuse & Permissions
  • Figure 51
    Figure 51
    Excited state spectroscopy of single-gated donors. (a) Differential conductance of a dopant in a FET. Excited states are indicated by the dots and arrows. Inset in (a) shows current ISD as a function of gate voltage at Vb=40mV where each plateau indicates the addition of a quantum channel due to an orbital. (b) Simulations of the gated donors eigenstates: wave function density of the D0-ground state (|ΨGS|2) located 4.3 nm below the interface in three different electric field regimes: Coulomb confinement regime 0MVm1 (left), hybridized regime 20MVm1 (middle), and interfacial confinement regime 40MVm1 (right). The gray plane indicates the Si/SiO2 interface. From 225.Reuse & Permissions
  • Figure 52
    Figure 52
    Sequential transport through a double donor device with independent gate control. The left panel shows the two opposing gates similar to a conventional FinFET geometry but with a split gate. The channel received a background doping of 1018P/cm3 and this device demonstrates independent gate control of two dopants. The right panel shows a finite-bias stability diagram revealing bulk-like excited states of the dopant. From 335.Reuse & Permissions
  • Figure 53
    Figure 53
    A donor-based double quantum dot in silicon. (a) An overview STM image of the device showing the two quantum dots, tunnel coupled to the source and drain (S/D) leads and capacitively coupled to the gates G1(2). (b) Close-up of the two quantum dots 4nm in diameter. The double quantum dot angle α=60°±3° has been optimized for maximum electrostatic control while suppressing parallel leakage through the dots. (c) Modeled and (d) measured charge stability diagrams show excellent agreement, demonstrating independent electrostatic control of the individual dots. From 433.Reuse & Permissions
  • Figure 54
    Figure 54
    Charge sensing using a donor-based single-electron transistor coupled to a small donor dot. (a) Filled-state STM image of the overall device pattern, showing (in lighter contrast) the regions where the hydrogen resist monolayer has been desorbed to create the source (S) and drain (D) contacts of the single-electron transistor, and the two gates (G1, G2). (b) High-resolution image of the device pattern within the white box in (a), showing the SET island (D1) and the quantum dot (D2) (c) Charge stability plot showing the dependence of ISD on the gate voltages (VG1, VG2), for a constant VSD=50μV. The high current lines correspond to the Coulomb peaks of the SET. Inset: High-resolution map of a small section of (c) showing discontinuity of a current line, due to a particular charge transition of D2. The triangles in the main map indicate a total of seven such transitions of D2. From 254.Reuse & Permissions
  • Figure 55
    Figure 55
    (a) Spin-lattice relaxation rate T11 of P donors in bulk Si, at B0.3T and T=1.2K, as a function of the field orientation. The angular dependence allows the separation of “valley repopulation” and “single valley” contributions. From 448. (b) T11(B) for single P donors in two different devices. Both show a T11B5 contribution, but device A also exhibits a B-independent plateau, attributed to dipolar flip-flops with nearby donors. Also shown is T11(3.3T) in bulk Si:P. From 277. (c) T11(B) in a gate-defined Si/SiGe dot (●), compared to data for a InGaAs dot (▪,⧫). From 144. (d) T1(B) in a gate-defined Si-MOS dot, for the one-electron (▪) and two-electron (○) states. From 456.Reuse & Permissions
  • Figure 56
    Figure 56
    (a) Experimental echo decay and cluster-expansion theory for SinatP at different angles of the magnetic field with respect to the crystallographic [001] axis. Notice the echo envelope modulation arising from anisotropic hyperfine coupling between donor electron and Si29 nuclei. From 453. (b) Decoherence time T2 for Si:P as a function of Si29 concentration CN for different dopant concentrations CE. Symbols are experimental data points. From 450.Reuse & Permissions
  • Figure 57
    Figure 57
    (a) Sketch of the spin-dependent transition between a donor-bound electron and an interface trap, following the creation of free carriers through illumination. (b) Schematics of an EDMR device. P donors close to charge traps at the Si/SiO2 interface contribute a spin-dependent scattering mechanism for the electrons traveling between the Au contacts. A resonant microwave excitation alters the polarization of the donor-bound electrons, causing a measurable change of the overall device resistance (c) Electrically detected Rabi oscillations of P-donor electrons at different values of the driving power. From 388.Reuse & Permissions
  • Figure 58
    Figure 58
    (a) Schematics of a single-charge trap coupled to the channel of a Si transistor. (b) Single-electron spin resonance measurement, obtained by monitoring the average current through the transistor as a function of magnetic field, while applying a microwave excitation at 45 GHz. The excess current at the resonance frequency arises from the change in charge occupancy of the trap, made possible by the driven flipping of its electron spin. From 458.Reuse & Permissions
  • Figure 59
    Figure 59
    (a) Spin-to-charge conversion scheme for a single donor tunnel coupled to the island of an SET. The presence of quantized states inside the SET island can be ignored if the single-particle energy level spacing is smaller than the thermal broadening. From 276. (b) Single-shot readout of a donor electron spin. The individual traces show the evolution of the readout signal as a function of the donor electrochemical potential with respect to the Fermi level. From 277.Reuse & Permissions
  • Figure 60
    Figure 60
    Single-shot readout of singlet-triplet states in a Si/SiGe double quantum dot. (a), (b) QPC current traces IQPC while pulsing the detuning with a square wave. Singlet states are identified when IQPC returns to a high value as in (b). (c) Charge stability diagram and pulsing levels. (d)–(f) Time traces of IQPC at different magnetic fields as indicated. Increasing B extends the lifetime of the T11 (constant current) state. (g) Control sequence, pulsing outside the spin-blockade region. From 314.Reuse & Permissions
  • Figure 61
    Figure 61
    Coherent manipulation of singlet-triplet states in a Si/SiGe double quantum dot. (a) Charge stability diagram of the double dot system. Arrows describe the trajectory in gate space during the pulsing sequence shown in (b). The (0,2) singlet state is prepared at point F. Adiabatically moving to point S, where the exchange coupling is very weak, brings the system to the (1,1) singlet. Pulsing to point E turns on the exchange and causes the two-spin state to oscillate between the (1,1) singlet and triplet. M is the measurement point where the electrons recombine in the (0,2) state if in a singlet state. (c) Rabi oscillations of the singlet probability as a function of the exchange pulse duration (time spent at point E) and (0,2)-(1,1) detuning ϵ. (d) Bloch sphere representation of the trajectories of the two-spin states for different initial values of the hyperfine fields. From 261.Reuse & Permissions
  • Figure 62
    Figure 62
    Single-atom electron spin qubit based on an implanted P31 donor. (a) Optimized design of an on-chip planar transmission line capable of delivering coherent microwave pulses at frequencies up to 50 GHz. From 76 (b) Scanning electron micrograph of the spin qubit device. (c) Rabi oscillations of the electron spin state with 10 dBm driving power at 30 GHz. (d) Measurement of spin coherence with Hahn echo and XYXY dynamical decoupling. (b)–(d) From 310.Reuse & Permissions
×

Sign up to receive regular email alerts from Reviews of Modern Physics

Log In

Cancel
×

Search


Article Lookup

Paste a citation or DOI

Enter a citation
×