Sketched oxide single-electron transistor

2 downloads 45 Views 3MB Size Report
Apr 17, 2011 - Sketched oxide single-electron transistor. Guanglei Cheng1, Pablo F. Siles2,3, Feng Bi1, Cheng Cen1, Daniela F. Bogorin1, Chung Wung ...
LETTERS PUBLISHED ONLINE: 17 APRIL 2011 | DOI: 10.1038/NNANO.2011.56

Sketched oxide single-electron transistor Guanglei Cheng1, Pablo F. Siles2,3, Feng Bi1, Cheng Cen1, Daniela F. Bogorin1, Chung Wung Bark4, Chad M. Folkman4, Jae-Wan Park4, Chang-Beom Eom4, Gilberto Medeiros-Ribeiro5 and Jeremy Levy1 * Devices that confine and process single electrons represent an important scaling limit of electronics1,2. Such devices have been realized in a variety of materials and exhibit remarkable electronic, optical and spintronic properties3–5. Here, we use an atomic force microscope tip to reversibly ‘sketch’ single-electron transistors by controlling a metal–insulator transition at the interface of two oxides6–8. In these devices, single electrons tunnel resonantly between source and drain electrodes through a conducting oxide island with a diameter of ∼1.5 nm. We demonstrate control over the number of electrons on the island using bottom- and side-gate electrodes, and observe hysteresis in electron occupation that is attributed to ferroelectricity within the oxide heterostructure. These single-electron devices may find use as ultradense non-volatile memories, nanoscale hybrid piezoelectric and charge sensors, as well as building blocks in quantum information processing and simulation platforms. The discovery of a high-mobility quasi-two-dimensional electron gas (q-2DEG) at the interface of two insulating oxides, TiO2-terminated SrTiO3 and LaAlO3 (ref. 6), has stimulated interest in the development of oxide-based electronics. The transition between insulating and conducting states in this system is an atomically sharp function of the number of LaAlO3 unit cells7. At or below a thickness of three unit cells of LaAlO3 , the interface is insulating, but for four or more unit cells the interface is conducting. The conductance of films grown at a critical thickness (three-unit-cell LaAlO3/SrTiO3) can be locally and reversibly controlled using a conductive atomic force microscope (c-AFM) probe technique8. Positive voltages applied to the c-AFM tip locally switch the three-unit-cell LaAlO3/SrTiO3 interface to a conducting state, and negative voltages locally restore the insulating state. The writing mechanism is believed to be governed by a ‘water cycle’9 in which the top LaAlO3 surface is locally charged through hydrogen passivation, resulting in high-resolution modulation doping of the LaAlO3/SrTiO3 interface. Using this technique, it is possible to create nanowires as small as 2 nm wide (ref. 10), islands as small as 1 nm in diameter (ref. 10), tunnel barriers10, rectifying junctions11, ‘SketchFET’ transistors10 and photoconductive switches12 with comparably small dimensions. Cryogenic operation of these devices13 raises the possibility that single-electron devices may also be created. Here, we report the creation and electronic characterization of a ferroelectric sketch-based single-electron transistor (SketchSET) at the three-unit-cell LaAlO3/SrTiO3 interface. These devices were created using the same c-AFM lithography technique presented in ref. 8. SketchSET devices can be created in a variety of ways, one of which is illustrated in Fig. 1a,b. Two crossed nanowires are written at the LaAlO3/SrTiO3 interface to a conducting state. The c-AFM tip is then positioned at the intersection, and an erase pulse is applied (duration, tb; tip voltage, Vtip ¼ Verase , 0), followed

by a brief positive pulse (duration, td; Vtip ¼ Vwrite . 0). This procedure creates an ultrasmall island that behaves as a quantum dot at the intersection. The quantum dot is surrounded by an insulating barrier and is separated from the four nanowires by a narrow tunnel barrier. The centre island, produced by a 10 ms write pulse, is estimated to have a diameter d of 1.5 nm, based on a calibration of the writing process performed on the same sample (see Supplementary Information). One can roughly estimate the number N of electrons able to reside within the quantum dot, based on typical two-dimensional carrier densities for nanoscale writing (n ≈ 5 × 1013 cm22) at the LaAlO3/SrTiO3 interface: N ¼ pd 2n/4 ≈ 1 electron. This estimate agrees well with the observed behaviour, described in detail below. Electrical transport experiments were performed on six different devices, including one control structure on which no island was written. Here, we describe the results obtained for three devices (A, B and C), which were created using the parameters described in Supplementary Table S1 and Fig. S2. The experiments focus on the effect of the side gates (Vg1 and Vg2) and back gate (Vgb) on the source–drain differential conductance (Gsd) and capacitance (Csd). Figure 1c shows the differential conductance curve for device A immediately after cooldown, with the side gates and Vgb all grounded, and the drain Vd (where current is measured) held at virtual ground. For sufficiently small source– drain voltage (Vsd), the differential conductance is strongly suppressed, and increases rapidly above a well-defined threshold. Although the structure is nominally symmetric, the threshold for positive and negative Vsd is generally different owing to hardto-control variations at the scale of 1 nm. Figure 1c shows two clearly resolved Coulomb peaks for Vsd , 0, and another three for Vsd . 0 before being obscured by the large conducting background. Subsequent voltage cycles resulted in fewer peaks (Fig. 2a). Structures that do not have islands at the intersection (such as SketchFET devices13) do not exhibit any Coulomb peaks. Here, we focus on transport for SketchSET devices in the lowconductance regime. Figure 2a,b shows the differential conductance Gsd and capacitance Csd of device A as a function of temperature (16–40 K) and source–drain voltage Vsd (–0.3 V to 0.3 V), with all three side gates and the back gate grounded. The conductance exhibits distinct Coulomb peaks, which are associated with resonant tunnelling into the quantum dot (Supplementary Fig. S3). Coinciding with these conductance peaks are abrupt changes in capacitance. Generally, increases in quantum dot occupancy DN ¼ 1 are associated with roughly constant capacitance jumps DC (see also Fig. 4). In the blue shadowed regime in Fig. 2b, the conductance is negligible and the capacitance, Csd , DC, is insensitive to Vsd. For these reasons, we infer that N ¼ 0 electrons are contained in the quantum dot in this regime.

1

Department of Physics and Astronomy, University of Pittsburgh, Pittsburgh, Pennsylvania 15260, USA, 2 Laborato´rio Nacional de Luz Sı´ncrotron, Caixa Postal 6192, 13083-970 Campinas SP, Brazil, 3 Instituto de Fı´sica ‘Gleb Wataghin’, Universidade Estadual de Campinas-UNICAMP, Campinas SP, Brazil, 4 Department of Materials Science and Engineering, University of Wisconsin-Madison, Madison, Wisconsin 53706, USA, 5 Hewlett Packard Laboratories, 1501 Page Mill Road, Palo Alto, California 94304, USA. *e-mail: [email protected] NATURE NANOTECHNOLOGY | VOL 6 | JUNE 2011 | www.nature.com/naturenanotechnology

343

LETTERS

NATURE NANOTECHNOLOGY

a

c

b

3

G2

Source

G1

Gsd (nS)

Gate 2

D Gate 1

2

1

5 nm

0 1

1

2

0 −1

10

−0.3

0.0 Vsd (V)

Drain

lO 3 LaA

Back

20

Gsd (nS)

30

S

DOI: 10.1038/NNANO.2011.56

0.3 Data Fit

0

iO 3 SrT

−0.3

0.0 Vsd (V)

0.3

Figure 1 | SketchSET schematic and transport characteristics. a, c-AFM sketching of a single-electron transistor device. b, Energy illustration of the SketchSET device. The vertical scale has arbitrary units, but reflects the relative electron energy barrier height. Quantum-dot tunnel barriers are created by applying a negative voltage pulse of duration tb. The quantum dot is formed by applying a positive voltage pulse of duration td. c, Differential conductance Gsd from source to drain at T ¼ 16 K. Conductance is suppressed at low biases and increases rapidly above a threshold voltage of 0.2 V. The red curve is an exponential fit to the data. The peaks in the inset are Coulomb peaks after subtracting an exponential background. Numbers with arrows indicate electron occupation.

1

b

40 K

40 15

0

35 K 35 K

c

0.6 0.3

Forward Reverse

d

T = 30 K

ΔVsd (V)

40 K

Gsd (nS)

2

a

0.0

25 K

20 25 K

20 K

20 K 10

16 K −0.3

0.0 Vsd (V)

0.3

−0.3

30 K

5

0

0.06 ΔVsd = 0.002 T−0.01

0.00

Csd (pF)

Gsd (nS)

30 30 K

10

ΔVsd = 0.01 T−0.2

0.12

16 K −0.3

0.0 Vsd (V)

0.0 Vsd (V) P

e

N = 1/2

Forward sweep −|e|VS

16

20

24 28 T (K)

f

N=1

Reverse sweep

Δϕ

32

−P

−|e|VS

Source

0.3

0.3

Drain Dot

Source

Drain Dot

Figure 2 | Temperature-dependent differential conductance and capacitance of device A. a, Differential conductance Gsd measured at temperatures ranging from 16 K (bottom black curve) to 40 K (top pink curve) in 1 K steps. Curves are shifted by 0.6 nS for clarity. b, Source–drain capacitance Csd measured over the range in a. Curves are shifted by 1.6 pF for clarity. A sharp change in Csd corresponds to a single electron tunnelling event. The shadowed blue, red and yellow regions indicate electron occupations of 0, 1 and 2, respectively. Green regions indicate hysteretic regions where electron occupation in the quantum dot changes by DN ¼+1. c, Gsd measured at T ¼ 30 K for forward (red) and reverse (blue) source–drain bias sweep directions. The Coulomb peaks are shifted by ferroelectric polarization in the SrTiO3. d, Coulomb peak width versus temperature. A kink is observed at TC1 ¼ 25 K, coincident with a ferroelectric phase transition in the SrTiO3. e, Schematic band diagram showing resonant tunnelling at Vsd ¼ 0 V in the forward sweep direction. The red arrow indicates the direction of ferroelectric polarization P. f, For the reverse sweep, the polarization is reversed (2P) and the electrochemical potential of the dot is lowered by Df so that the system is in the Coulomb blockade regime.

The Coulomb peaks and associated capacitance jump locations exhibit hysteresis with respect to the source–drain voltage sweep direction. The conductance peak position, peak width and hysteresis magnitude vary significantly with temperature (Supplementary Fig. S3). With increasing temperature, the peak position first shifts to more negative Vsd , then at TC1 ¼ 25 K it begins to increase with temperature. The width of the peak increases approximately linearly with temperature; above T ¼ TC1 , the slope increases by a factor of five (Fig. 2d). The observed hysteresis in the Coulomb peak position is attributed to ferroelectric switching in the SrTiO3 barrier. The lattice constants of LaAlO3 and SrTiO3 are 3.789 Å and 3.905 Å respectively (3% mismatch), and the three-unit-cell LaAlO3 is coherently strained biaxially to match the SrTiO3 lattice constant. The profound effect of strain on thin SrTiO3 layers is well known14–17. Field-emission experiments on LaAlO3/SrTiO3-based SketchFET devices13 show evidence of a diverging dielectric permittivity 344

associated with structural phase transitions in the near-interface SrTiO3 region at TC1 ¼ 25 K and TC2 ¼ 65 K. The hysteretic behaviour observed as a function of local in-plane applied electric fields is highly non-monotonic with respect to temperature, and exhibits anomalies at a known structural transition TC1. This hysteresis is qualitatively distinct from the hysteretic changes in polarization that have been reported for vertically gated LaAlO3/SrTiO3 heterostructures18. Hysteresis in both conduction and capacitance as a function of the back-gate bias is observed for device A (Supplementary Fig. S8); however, as noted in ref. 18, it is difficult to distinguish ferroelectric hysteresis from trap charging or other polarization effects in this geometry. Ferroelectric polarization has a profound influence on the resonant tunnelling characteristics of the SketchSET, and is capable of switching the conductance of the source–drain channel between ‘on’ and ‘off’ states (Fig. 2c). At T ¼ 30 K, when Vsd is swept in the forward direction, resonant tunnelling is observed at

NATURE NANOTECHNOLOGY | VOL 6 | JUNE 2011 | www.nature.com/naturenanotechnology

NATURE NANOTECHNOLOGY T = 25 K, Vg2 = 0 V

100 75 50

2

23 20

1

15

0

10 5

−1 25

0

−2

0

Gsd (nS)

Gsd (nS)

b

2.5 V 1.0 V 0V −1.0 V −2.5 V

125

Vg1 (V)

a 150

LETTERS

DOI: 10.1038/NNANO.2011.56

−5 −2

−1

0 Vsd (V)

1

2

−2

−1

0 Vsd (V)

1

2

Figure 3 | Device B side gating at T 5 25 K. a, Differential conductance Gsd for various side-gate voltages Vg1 ¼ 22.5 V, 21.0 V, 0 V, 1.0 V and 2.5 V. The other side gate is grounded: Vg2 ¼ 0 V. b, Two-dimensional plot of Gsd versus Vg1 and Vsd. A single Coulomb oscillation is observed only when Vg1 and Vsd have opposite signs. The red arrows in a and b mark resonant tunnelling.

a

b

Vbg = 0 V

0.0

1.0

−0.1

0.8

−0.2

0.5

−0.3

0.2

Vbg (V)

2

Gsd (nS)

Gsd (nS)

3

1

0

−0.4 V −0.2

0.0

0.2

−0.4 −0.2

0.4

−0.1 −0.1

0.0

Vsd (V)

c

d

2 20 1

2

1

0.0

2

1

0.1 Vsd (V)

0.2

0

0.3

1

2

−0.1

0

7.5 Vbg (V)

15

Csd (pF)

Csd (pF)

10

Vbg = 0 V

−0.2

10

5 −0.3

5

−0.4 V −0.2

0.0

0.2

0.4

−0.4 −0.2

Vsd (V)

2.5 −0.1

0.0

0.1 Vsd (V)

0.2

0.3

Figure 4 | Differential conductance and capacitance dependence on back-gate voltage at T 5 16 K, device C. a, Differential conductance (Gsd) curves at back-gate voltages Vbg from 20.4 V to 0 V in 0.02 V steps. Curves are offset by 0.2 nS for clarity. The second electron emerges at Vbg ¼ 20.1 V. b, Intensity plot of Gsd. c, Capacitance curves at the same gating conditions as in a. Curves are offset by 0.3 pF for clarity. The discrete jumps coincide with changes in electron occupancy in the quantum dot. d, Intensity plot of Csd. Numbers in red in c and d indicate electron occupation N in the quantum dot.

Vsd ¼ 0, with Gsd ¼ 0.3 nS. After sweeping Vsd to 0.3 V and returning to Vsd ¼ 0, the conductance has vanished. The effect of ferroelectric polarization on the SketchSET characteristics can be understood qualitatively by expanding the constantinteraction (CI) picture2, in which there are well-defined energy levels in the quantum dot, to include the effect of ferroelectric polarization. Single-electron tunnelling occurs when an allowed energy state within the quantum dot becomes resonant with the electrochemical potential of either the source lead (ms) or drain lead (md). Such a resonant condition is indicated in Fig. 2e. Ferroelectric tunnel barriers can shift the chemical potential within the quantum dot (by an amount Df ) so that resonance only occurs

for one polarization direction (Fig. 2f ). The SketchSET can be regarded as a local sensitive detector of polarization in the SrTiO3; however, it is only sensitive when the SketchSET undergoes resonant tunnelling. Within the CI framework, the Coulomb peak spacing DVsd can be used to estimate the addition energy Ea by DVsd ¼ Ea/ea, where the lever arm a ¼ Cs/CQD is the ratio between the sourceto-quantum dot capacitance Cs and total quantum dot capacitance CQD. The width of the Coulomb peak is expected to broaden linearly with temperature19: dVsd/dT ¼ 3.5kB/ea, where kB is the Boltzmann constant. From Fig. 2d, we can estimate a ≈ 0.15 for T , TC1 and a ≈ 0.03 for T . TC1 , respectively. The addition energy below

NATURE NANOTECHNOLOGY | VOL 6 | JUNE 2011 | www.nature.com/naturenanotechnology

345

LETTERS

NATURE NANOTECHNOLOGY

25 K cannot be estimated, because N ¼ 2 states are not observed. As the Coulomb peak shifts to positive values of Vsd , a second electron state emerges at T . 31 K with spacing DVsd ≈ 0.3 V (Fig. 2b), which allows for an addition energy to be estimated as Ea ¼ 9 meV and thus CQD ≥ 18 aF. An upper limit of CQD ≤ 40 aF can be inferred from the thermal broadening effect, which occurs at T ≈ 45 K. The energy shift caused by the ferroelectric remnant polarization can be estimated in a similar manner: DEFE ≈ 2–6 meV. There are two principal mechanisms by which the ferroelectric polarization can produce an energy shift (Supplementary Fig. S7). If the polarization is spatially uniform, then it can only couple to the dipole moment of the quantum dot. If the polarization is spatially non-uniform, then it can couple directly to the charge. The first scenario is outlined in Fig. 2e,f, but broken symmetries could lead to a situation in which the polarization breaks into non-symmetric domains. Such a case would also produce a direct ferroelectric coupling to the charge in the quantum dot. Resonant tunnelling through the quantum dot can be modified through the application of either side gates or a back gate. The strength of the side-gate coupling can be adjusted by varying the barrier width (controlled by the duration of the erase pulse tb) or by writing side gates a fixed distance from the quantum dot. Two examples are shown for side gates created close to the island (device B) and separated by a 50 nm gap (device A). For device A (Supplementary Fig. S4), a clearly defined resonance is shifted by DVsd ¼ 40 mV as the side gate Vg is changed from 20.2 V to þ0.2 V, with no evidence of gate leakage. The low coupling factor, DVs/DVg ¼ 0.1 in this case, reflects the geometric separation of the side gate. For device B, the gates are separated by approximately the same distance as the source and drain leads; hence, significant gate leakage is observed. Sharp resonances are only observed when Vg1 and Vsd have opposite signs (Fig. 3). This ‘differential mode’ is more effective in aligning the lowest quantum dot chemical potential level than the ‘common mode’ configurations (Supplementary Fig. S5). SketchSET structures respond to modest (|Vbg| , 1 V) voltages applied to the bottom of the SrTiO3 substrate. The sensitivity of the SketchSET to back gating, as measured by changes in the effective carrier density within the quantum dot, is more than two orders of magnitude larger than what has been reported for back-gate control of the metal–insulator at room temperature7 and superconductivity at low temperature20. The principal reason for the relatively high sensitivity is that the high curvature of the metallic nanostructures focuses the electric flux lines, thereby greatly increasing the electric field effect at the device. Figure 4 shows the differential conductance and capacitance as a function of back-gate bias Vbg and source–drain bias Vsd for device C. For the lowest value Vbg ¼ 20.4 V, a single resonant peak is observed as Vsd is swept from 0 V in either the positive or negative directions. The resonant peaks mark a transition from N ¼ 0 electrons in the quantum dot to N ¼ 1 electrons. The increase in N is associated with a discrete jump in the measured capacitance (Fig. 4c,d), increasing from C0 ≈ 2.5 pF for Vsd ¼ 0 by an amount DC2 ¼ 3.5 pF (DCþ ¼ 2.5 pF) for scans with increasing |Vsd|. Although the structure is nominally symmetric with respect to source and drain, there are clear asymmetries, which result in different thresholds for positive and negative Vsd. As Vbg is increased, the resonances shift towards Vsd ¼ 0 V (Fig. 4d), and a second resonant peak is observed on the Vsd , 0 side. The higher stability of the N ¼ 2 state, as measured by the strong suppression of the conductance near Vsd ¼ 20.18 V, and the greater width of the N ¼ 2 plateau in the capacitance, is believed to be a manifestation of the well-known ‘shell filling effect’ that has been observed in twodimensional quantum-dot systems21,22 and III–V self-assembled quantum dots23. Higher N values are not observed because the 346

DOI: 10.1038/NNANO.2011.56

structure becomes highly conducting outside the range shown, but a similar non-uniform spacing is apparent in Fig. 4c. A remarkable feature of the SketchSET is the high sensitivity of the capacitance to changes in electron occupation in the quantum dot. The change in capacitance (pF) observed during a single-electron charging event at the quantum dot is approximately three orders of magnitude too large to be accounted for solely by electrostatic effects. It is well known that SrTiO3 is a high-permittivity incipient ferroelectric with a dielectric constant that can exceed 1 ≈ 1 × 104 at low temperatures, which is easily perturbed by structural deformation, strain and electric fields24. The presence of a single bound electron at the LaAlO3/SrTiO3 interface is predicted to produce a large distortion of the SrTiO3 octahedra that extends far beyond the location of the charge8. This structural distortion increases the polarizability of the nearby SrTiO3 , thus increasing the parasitic capacitance Cp between source and drain (Supplementary Fig. S6). The inferred CQD is roughly four orders of magnitude smaller than the measured capacitance Cp. The unique properties of this ferroelectric SketchSET provide new opportunities for combining the ultrahigh electrostatic sensitivity of SET devices with ferroelectric-derived sensitivity at the nanoscale. Because all ferroelectric materials are also piezoelectric, a natural coupling between charge and nanomechanical motion is expected for the SketchSET. Furthermore, a variety of phenomena associated with the spin degree of freedom for single-electron devices25,26 is expected to hold for these devices. By integrating oxide heterostructures with silicon27, it may be possible to integrate ferroelectric SketchSET scanning probes that are capable of measuring charge and displacement simultaneously at the nanoscale28. The existence of a ferroelectrically programmable SET constitutes a new type of nanoscale memory architecture that could be useful for lowpower, ultrahigh-density storage, if the charging energy and ferroelectric polarization could be made to persist to room temperature14,15,27. The method for creating a single SketchSET is readily replicated as one- or two-dimensional arrays, which may find use in quantum dot-based quantum computation29 or as a versatile solid-state ‘Hubbard toolbox’30 capable of exploring new artificial quantum states of matter.

Methods The three-unit-cell LaAlO3/SrTiO3 samples for devices A, B and C were grown at 780 8C with an oxygen background pressure of 7.5 × 1025 mbar using pulsed laser deposition. The samples were then patterned with six gold electrodes contacting the interface by ion milling 25 nm and backfilling with 2 nm titanium and 23 nm gold. AFM lithography was carried out using a commercial atomic force microscope (Asylum Research MFP-3D) under conditions of controlled relative humidity (40%). The tip position was controlled by a custom-made Labview program that converts graphic information to voltage commands, which are sent to the AFM scanner. Before writing the SketchSET device, the sample surface was cleaned by slowly scanning the AFM tip with an applied voltage of 210 V to remove residue charges at the interface. During writing, conductance between the electrodes of interests was monitored simultaneously. Transport measurements were performed using a dualphase lock-in amplifier at 25.5 Hz. The lock-in phase was adjusted so that the x channel reflected the conductance and the y channel reflected the capacitance.

Received 9 February 2011; accepted 16 March 2011; published online 17 April 2011

References 1. Kastner, M. A. The single-electron transistor. Rev. Mod. Phys. 64, 849–858 (1992). 2. Hanson, R., Kouwenhoven, L. P., Petta, J. R., Tarucha, S. & Vandersypen, L. M. K. Spins in few-electron quantum dots. Rev. Mod. Phys. 79, 1217–1265 (2007). 3. Klein, D. L., Roth, R., Lim, A. K. L., Alivisatos, A. P. & McEuen, P. L. A singleelectron transistor made from a cadmium selenide nanocrystal. Nature 389, 699–701 (1997). 4. Kubatkin, S. et al. Single-electron transistor of a single organic molecule with access to several redox states. Nature 425, 698–701 (2003). 5. Stampfer, C. et al. Tunable graphene single electron transistor. Nano Lett. 8, 2378–2383 (2008).

NATURE NANOTECHNOLOGY | VOL 6 | JUNE 2011 | www.nature.com/naturenanotechnology

NATURE NANOTECHNOLOGY

DOI: 10.1038/NNANO.2011.56

6. Ohtomo, A. & Hwang, H. Y. A high-mobility electron gas at the LaAlO3/SrTiO3 heterointerface. Nature 427, 423–426 (2004). 7. Thiel, S., Hammerl, G., Schmehl, A., Schneider, C. W. & Mannhart, J. Tunable quasi-two-dimensional electron gases in oxide heterostructures. Science 313, 1942–1945 (2006). 8. Cen, C. et al. Nanoscale control of an interfacial metal–insulator transition at room temperature. Nature Mater. 7, 298–302 (2008). 9. Bi, F. et al. ‘Water-cycle’ mechanism for writing and erasing nanostructures at the LaAlO3/SrTiO3 interface. Appl. Phys. Lett. 97, 173110 (2010). 10. Cen, C., Thiel, S., Mannhart, J. & Levy, J. Oxide nanoelectronics on demand. Science 323, 1026–1030 (2009). 11. Bogorin, D. F. et al. Nanoscale rectification at the LaAlO3/SrTiO3 interface. Appl. Phys. Lett. 97, 013102 (2010). 12. Irvin, P. et al. Rewritable nanoscale oxide photodetector. Nature Photon. 4, 849–852 (2010). 13. Cen, C., Bogorin, D. F. & Levy, J. Thermal activation and quantum field emission in a sketch-based oxide nano transistor. Nanotechnology 21, 475201 (2010). 14. Haeni, J. H. et al. Room-temperature ferroelectricity in strained SrTiO3. Nature 430, 758–761 (2004). 15. Warusawithana, M. P. et al. A ferroelectric oxide made directly on silicon. Science 324, 367–370 (2009). 16. Jang, H. W. et al. Ferroelectricity in strain-free SrTiO3 thin films. Phys. Rev. Lett. 104, 197601 (2010). 17. Zubko, P., Catalan, G., Buckley, A., Welche, P. R. L. & Scott, J. F. Strain-gradientinduced polarization in SrTiO3 single crystals. Phys. Rev. Lett. 99, 167601 (2007). 18. Singh-Bhalla, G. et al. Built-in and induced polarization across LaAlO3/SrTiO3 heterojunctions. Nature Phys. 7, 80–86 (2010). 19. Bockrath, M. et al. Single-electron transport in ropes of carbon nanotubes. Science 275, 1922–1925 (1997). 20. Caviglia, A. D. et al. Electric field control of the LaAlO3/SrTiO3 interface ground state. Nature 456, 624–627 (2008). 21. Tarucha, S., Austing, D. G., Honda, T., Van der Hage, R. J. & Kouwenhoven, L. P. Shell filling and spin effects in a few electron quantum dot. Phys. Rev. Lett. 77, 3613–3616 (1996). 22. Kouwenhoven, L. P. et al. Excitation spectra of circular, few-electron quantum dots. Science 278, 1788–1792 (1997).

LETTERS

23. Fricke, M., Lorke, A., Kotthaus, J. P., Medeiros-Ribeiro, G. & Petroff, P. M. Shell structure and electron–electron interaction in self-assembled InAs quantum dots. Europhys. Lett. 36, 197–202 (1996). 24. Mu¨ller, K. A. & Burkard, H. SrTiO3: an intrinsic quantum paraelectric below 4 K. Phys. Rev. B 19, 3593–3602 (1979). 25. Goldhaber-Gordon, D. et al. Kondo effect in a single-electron transistor. Nature 391, 156–159 (1998). 26. Morello, A. et al. Single-shot readout of an electron spin in silicon. Nature 467, 687–691 (2010). 27. Park, J. W. et al. Creation of a two-dimensional electron gas at an oxide interface on silicon. Nat. Commun. 1, 94 (2010). 28. Knobel, R. G. & Cleland, A. N. Nanometre-scale displacement sensing using a single electron transistor. Nature 424, 291–293 (2003). 29. Loss, D. & DiVincenzo, D. P. Quantum computation with quantum dots. Phys. Rev. A 57, 120–126 (1998). 30. Jaksch, D. & Zoller, P. The cold atom Hubbard toolbox. Ann. Phys. 315, 52–79 (2005).

Acknowledgements This work was supported by US National Science Foundation (DMR-0704022 and DMR-0906443), US Defense Advanced Research Projects Agency (W911NF-09-10258), US Army Research Office (W911NF-08-1-0317), The Fine Foundation, US Air Force Office of Scientific Research (FA9550-10-1-0524), a David and Lucile Packard Fellowship and the Fundac¸a˜o de Amparo a` Pesquisa do Estado de Sa˜o Paulo – FAPESP (contact project 05/04643-7).

Author contributions G.C. carried out the major experiments. P.F.S. and C.C. carried out preliminary experiments. F.B., G.C. and D.F.B. contributed to device fabrication. C.W.B., C.M.F., J.W.P. and C.B.E. contributed to sample growth. J.L., G.C., C.C. and G.M.R. discussed and analysed the results. All authors contributed to writing of the manuscript.

Additional information The authors declare no competing financial interests. Supplementary information accompanies this paper at www.nature.com/naturenanotechnology. Reprints and permission information is available online at http://www.nature.com/reprints/. Correspondence and requests for materials should be addressed to J.L.

NATURE NANOTECHNOLOGY | VOL 6 | JUNE 2011 | www.nature.com/naturenanotechnology

347