+ All Categories
Home > Documents > Cluster-like resistive switching of SrTiO3: Nb surface layers

Cluster-like resistive switching of SrTiO3: Nb surface layers

Date post: 28-Nov-2023
Category:
Upload: independent
View: 0 times
Download: 0 times
Share this document with a friend
27
Cluster-like resistive switching of SrTiO 3 :Nb surface layers C Rodenb ¨ ucher 1,2,7 , W Speier 2 , G Bihlmayer 1,2,3 , U Breuer 4 , R Waser 1,2,5 and K Szot 1,2,6 1 Forschungszentrum J¨ ulich GmbH, Peter Gr ¨ unberg Institut (PGI-7), D-52425 J¨ ulich, Germany 2 Forschungszentrum J¨ ulich GmbH, JARA—Fundamentals of Future Information Technologies, D-52425 J¨ ulich, Germany 3 Forschungszentrum J¨ ulich, Institute of Advanced Simulation (IAS-1), D-52425 J¨ ulich, Germany 4 Forschungszentrum J¨ ulich GmbH, Zentralinstitut f ¨ ur Engineering, Elektronik und Analytik (ZEA-3), D-52425 J ¨ ulich, Germany 5 RWTH Aachen, Institut f¨ ur Werkstoffe der Elektrotechnik 2, D-52056 Aachen, Germany 6 University of Silesia, August Chelkowski Institute of Physics, 40-007 Katowice, Poland E-mail: [email protected] New Journal of Physics 15 (2013) 103017 (14pp) Received 28 June 2013 Published 17 October 2013 Online at http://www.njp.org/ doi:10.1088/1367-2630/15/10/103017 Abstract. The understanding of the resistive switching mechanisms in perovskites is of particular importance for the development of novel non-volatile memories. Nanoscale investigations recently revealed that in the model material SrTiO 3 a filamentary type of switching is present. In this paper, we show that upon donor doping with Nb the switching type changes fundamentally. We report on the observation of conducting clusters that can be switched independently between a high resistance and a low resistance state when applying a voltage. Furthermore, we show that the resistive switching takes place in a semiconducting surface layer on top of the metallic bulk of SrTiO 3 :Nb single crystals, which can change its properties easily under external gradients. 7 Author to whom any correspondence should be addressed. Content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI. New Journal of Physics 15 (2013) 103017 1367-2630/13/103017+14$33.00 © IOP Publishing Ltd and Deutsche Physikalische Gesellschaft
Transcript

Cluster-like resistive switching of SrTiO3:Nbsurface layers

C Rodenbucher1,2,7, W Speier2, G Bihlmayer1,2,3, U Breuer4,R Waser1,2,5 and K Szot1,2,6

1 Forschungszentrum Julich GmbH, Peter Grunberg Institut (PGI-7),D-52425 Julich, Germany2 Forschungszentrum Julich GmbH, JARA—Fundamentals of FutureInformation Technologies, D-52425 Julich, Germany3 Forschungszentrum Julich, Institute of Advanced Simulation (IAS-1),D-52425 Julich, Germany4 Forschungszentrum Julich GmbH, Zentralinstitut fur Engineering,Elektronik und Analytik (ZEA-3), D-52425 Julich, Germany5 RWTH Aachen, Institut fur Werkstoffe der Elektrotechnik 2,D-52056 Aachen, Germany6 University of Silesia, August Chełkowski Institute of Physics,40-007 Katowice, PolandE-mail: [email protected]

New Journal of Physics 15 (2013) 103017 (14pp)Received 28 June 2013Published 17 October 2013Online at http://www.njp.org/doi:10.1088/1367-2630/15/10/103017

Abstract. The understanding of the resistive switching mechanisms inperovskites is of particular importance for the development of novel non-volatilememories. Nanoscale investigations recently revealed that in the model materialSrTiO3 a filamentary type of switching is present. In this paper, we showthat upon donor doping with Nb the switching type changes fundamentally.We report on the observation of conducting clusters that can be switchedindependently between a high resistance and a low resistance state whenapplying a voltage. Furthermore, we show that the resistive switching takes placein a semiconducting surface layer on top of the metallic bulk of SrTiO3:Nbsingle crystals, which can change its properties easily under external gradients.

7 Author to whom any correspondence should be addressed.

Content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence.Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal

citation and DOI.

New Journal of Physics 15 (2013) 1030171367-2630/13/103017+14$33.00 © IOP Publishing Ltd and Deutsche Physikalische Gesellschaft

2

Based on various measurements, we postulate that ionic movements leading tothe creation of secondary phases as nano-filaments between the clusters have tobe taken into account in modelling the resistive switching.

S Online supplementary data available from stacks.iop.org/NJP/15/103017/mmedia

Contents

1. Introduction 22. Methods 33. Results 4

3.1. Resistive switching on the macroscale . . . . . . . . . . . . . . . . . . . . . . 43.2. Resistive switching on the nanoscale . . . . . . . . . . . . . . . . . . . . . . . 63.3. Surface layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

4. Discussion and conclusions 11Acknowledgments 13References 13

1. Introduction

In recent years SrTiO3, a prototype perovskite material, has attracted much attention becauseit is possible to change its resistance from insulator to metal and even to superconductor [1]by applying electric fields thus making such materials very promising for future applicationsin redox-based resistive switching memory cells (ReRAM) [2]. Resistive switching has beenreported in SrTiO3 thin films [3] as well as in single crystals [4]. The switching betweeninsulator and metal is related to the transition of a titanium ‘d’-electron from d0 to d2 and d3.This ‘d’-electron can be generated either

• by a chemical gradient (reduction in low pO2),

• by an electrical gradient (electroreduction),

• by a convolution of these gradients,

• by donor doping.

Here, pentavalent Nb is used as a donor for SrTiO3 substituting the tetravalent Ti.Although after this doping the crystal should be metallic at low carrier concentrations[5, 6], surprisingly, resistive switching can be observed in Nb-doped SrTiO3 thin films[7] and even in single crystals [8–26], although the nature of the switching process is stillunder discussion (for details see supplement). Furthermore, a highly resistive surface layer,which arises as soon as the crystal comes into contact with oxygen, was detected on SrTiO3:Nbinfluencing the properties of the whole sample [27, 28]. In general, it is well known that suchsurface layers exist in perovskites and they have been investigated intensively for BaTiO3 [29,30]. In our previous study on SrTiO3:Nb thin films [7], we showed that the resistive switchingon the nanoscale is not of a filamentary type, as in undoped SrTiO3, but is related to thepresence of regularly arranged switching blocks representing a cluster-like switching behaviour.In the present paper, we intend to focus on the fundamental physical mechanisms behind

New Journal of Physics 15 (2013) 103017 (http://www.njp.org/)

3

this behaviour by investigating the properties of Nb-doped single crystals on the macro-and the nanoscale. In order to obtain an insight into the origin of the resistive switchingphenomenon in single crystals, a careful investigation of the local structure and chemicalcomposition of surface layer and bulk has to be made since in SrTiO3 clustering of defectsas found in doped and undoped thin films [7, 31, 32] or a formation of secondary phases caneasily evolve. According to the redox equation, which describes the defect chemistry of theNb doping, the donor compensation can change under oxidizing conditions from electroniccompensation to compensation by Sr vacancies leading to the creation of SrO, which canform Ruddlesden–Popper phases or islands on the surface [33–36]. Simultaneously, the samplechanges from a metal to a highly insulating material

2SrSr + 2Nb•

Ti + 6OO + 2e′ + 12O2(g) V′′

Sr + SrSr + 2Nb•

Ti + 6OO + SrO(s.p.). (1)

Here, the Kroger–Vink notation is used and the equation is expressed for two units of the ABO3

perovskite crystal. (g) and (s.p.) denote the gas phase and a secondary solid phase, respectively.Due to the frozen-in Schottky equilibrium, the oxygen vacancy concentration in donor-dopedSrTiO3 is extremely low. If the chemical composition deviates from the stoichiometric case,which may occur for inhomogeneities of the surface layer, different reactions will take place.In Sr-deficient SrTiO3, reduction leads to the creation of Ti-rich phases [36], which have beenfound particularly in doped SrTiO3 [37–41]. Further reduction of TiO2 can then lead to theformation of substoichiometric titanium oxides such as Magneli phases [42]:

TiO2 →12O2 + TinO2n−1. (2)

In this paper, we investigate the resistive switching related to the transport behaviour on themacroscale by electrical four-point measurements and on the nanoscale by local conductivityatomic force microscopy (LC-AFM) under ultra-high vacuum (UHV) conditions. Based oncareful investigations of the surface layer under chemical and electrical gradients, we suggest ananoscale model of the resistive switching.

2. Methods

The present measurements were conducted on commercially available epi-polished SrTiO3:Nb(100) single crystals with a doping concentration of 1.4 at.% supplied by Crystec (Berlin) andMateck (Julich). X-ray diffraction measurements were performed using an STOE instrument onpowders of the single crystals with a doping concentration of 0.2, 1.4 and 10.1 at.%, and thelattice constant was determined using Rietveld refinement. Additionally, a Si standard (NIST640c) was added to take zero shift into account. Electrical measurements on the macroscalewere performed using a home-made probe station on samples contacted with Pt paste underambient conditions. For the four-point measurements in Valdes geometry, Pt electrodes weresputtered on the epi-polished side of the sample. Under UHV conditions, a voltage was appliedto two electrodes and the potential between all four electrodes was measured by electrometers.To carry out temperature-dependent four-point measurements up to a temperature of 1000 ◦C,Pt electrodes were pasted in the same geometry as the sputtered ones. Etching of the samplewas performed using hydrofluoric acid to obtain etch pits at the exits of the dislocations onthe surface. LC-AFM measurements were made using a JEOL JSPM setup with a Pt/Ir-coatedsilicon tip in contact mode under vacuum conditions and at different temperatures. During thescan, a bias voltage of 5–10 mV was applied to the tip and the measured current was used

New Journal of Physics 15 (2013) 103017 (http://www.njp.org/)

4

-6 -4 -2 0 2 4 6

10-6

10-5

10-4

10-3

10-2

| I |

(A)

U (V)

ref

900 °C

1050 °C

1200 °C

STO:Nb

10-7

Figure 1. Resistive switching on the macroscale. I–V curves obtained at samplesoxidized at different temperatures. (The dotted lines indicate the respective noiselevel.)

to calculate the resistance of the sample after subtraction of the noise level using Gwyddionsoftware [43]. X-ray photoelectron spectroscopy (XPS) spectra were obtained by a Perkin-Elmer instrument using monochromatized Al-Kα rays under UHV conditions. The backgroundof the spectra was subtracted by the Shirley method and the peaks were simulatedby Gauss–Lorentz functions. Atom probe measurements were conducted on a CAMECALEAP4000 X HR instrument using UV-laser pulsing after preparing a small needle ofSrTiO3:Nb (1.4 at.%) by a focused ion beam. Finite element simulations were conducted usingan ANSYS program simulating clusters with a diameter of 40 nm. The assumed values forthe resistivity were ρc = 1 × 10−6 � m for the clusters and ρm = 1 × 102 � m for the matrixaccording to the simulations for homogeneous SrTiO3 thin films [44]. The resistivity ofthe bridge was adjusted to ρb = 1.3 × 10−2 � m corresponding to our measurements wherea switching voltage of 4 V results in a current of 1 µA. For the thermal conductivity k,temperature-dependent values were chosen on the basis of laser flash measurements.At room temperature the thermal conductivity is k = 9 W m−1 K−1 and at higher temperatures itdecreases according to a 1/T law.

3. Results

3.1. Resistive switching on the macroscale

The resistive switching behaviour was investigated by performing electrical measurementson the macroscale. Three unpolished samples (5 × 5 × 0.5 mm) cut from the same crystalwere investigated after annealing in air for 5 h at different temperatures. Additionally, oneas-received reference sample was measured. The samples were contacted using Pt paste onthe top and the bottom side, and I–V curves were measured as shown in figure 1. It can beseen that the conductivity decreases systematically upon increasing the annealing temperature.Furthermore, all samples besides that annealed at 1200 ◦C exhibited bipolar resistive switching,which is surprising since the geometry of samples and electrodes was nominally symmetrical.

New Journal of Physics 15 (2013) 103017 (http://www.njp.org/)

5

1A2 B

10 Ω

10 mΩ

VV V

Metallic bulk

Switchable surface

d

U (

V)

a

-0.4

0.0

0.4

0.8

-0.4

0.0

0.4

0.8

-10 -5 0 5 10

-0.4

0.0

0.4

0.8

(A-B)

(1-2)(A-1) (2-B)

-10 -5 0 5 10I (mA)

-2

0

2

A B1 2

-10 -5 0 5 10

U (

V)

I (mA)I (mA)-10 -5 0 5 10

I (mA)

(mV)

ref 900 1050 1200

0

20

40

d (µ

m)

ref 900 1050 1200

10-10

10-9

10-8

10-7

C (

F)

TA (°C) TA (°C)

20 mA 50 mA100 mA (A-B)

metallicsemi-conducting

0 200 400 600 800 1000

2

4

6

8

10

12

R (

Ω)

T (°C)

b

0 400 8000

2

4

6

T (°C)0 400 800

T (°C)0 400 800

0

2

4

6

T (°C)

(1-2)(A-1) (2-B)

R (

Ω)

c

Figure 2. Electrical measurement of the surface layer. (a) V–I curve obtainedunder UHV conditions between outer electrodes showing switching behaviour.Below: voltage between the inner and outer electrodes. (b) Simulation ofthe resistance of the sample by a network of resistors representing surfacelayer (green) and bulk (red). (c) Resistance as a function of temperature afterelectroreduction with different currents. (d) Capacitance and calculated thicknessof the surface layer measured on samples oxidized at different temperatures.

This indicates that an asymmetry was introduced in the system related to the preparationprocedure or to electroforming effects. In order to investigate the effect of electrodegradationon the surface layer, which could be the reason for the appearance of bipolar switching, weconducted four-point measurements in Valdes geometry (cf figure 2). The outer electrodeswere connected to a current source and the potential between the outer and inner electrodeswas measured. The V–I curves recorded between electrodes A and B displayed occasionalbipolar switching behaviour as shown in figure 2(a). Regarding the individual contributionsbetween the inner electrodes, we can see that switching only occurred between electrodesA and 1. Although the electrodes were symmetrical, electrode A had the role of the anodeat the beginning of the voltage sweep and the asymmetry could have been created by ionicmovements, as will be discussed below. The potential between the inner electrodes 1 and 2,representing the properties of the bulk, showed ohmic behaviour with low resistance confirming

New Journal of Physics 15 (2013) 103017 (http://www.njp.org/)

6

that not the bulk of the sample but only the surface layer is responsible for the switching.To illustrate the difference between bulk and surface, we performed a simulation program withintegrated circuit emphasis (SPICE) simulation as depicted schematically in figure 2(b). Weassumed a network of resistors with two rows (10 � each) representing the highly resistivesurface layer while the resistors representing the bulk had much lower resistances of 10 m�.The simulation qualitatively reproduced the ratio of the measured voltages between the innerand outer electrodes and proved that the assumption of a surface layer with high resistivitywas justified. The difference between the bulk and the surface layer becomes even moreobvious with temperature-dependent measurements. While the as-received surface layer wassemiconducting with a high resistance (>108 �), the resistance dropped significantly afterperforming electroreduction by applying a constant current to the outer electrodes at 1000 ◦C.The resistance curves measured during cooling (figure 2(c)) showed a characteristic andreproducible temperature dependence. Starting at high temperature, at first the resistancedecreased with the decreasing temperature indicating metallic behaviour until a minimumwas reached around 250 ◦C. Upon further cooling, the resistance increased again indicatingsemiconducting behaviour. If more powerful electroreduction was performed by increasing thecurrent, the shapes of the curves at higher temperatures did not change at all, whereas theresistance of the semiconducting part was reduced. In contrast to the surface layer, the bulkhad metallic properties throughout the whole temperature range as expected for doped SrTiO3.In order to estimate the thickness of the surface layer, we measured the serial capacitance using asmall ac voltage (100 mV, 10 kHz) on the same samples as investigated in figure 1. We obtaineda very high capacitance, which decreases with the annealing temperature TA (figure 2(d)). Thehigh value of the capacitance indicates that not the whole volume of the sample but only thesurface layers contribute to the measured capacitance. This gives us the opportunity to estimatethe thickness of the surface layer. Therefore we assumed the relative permittivity to be ε = 300,which is a typical value of SrTiO3, and we calculated the thickness. While the thickness of thesurface layer of the as-received sample was only in the range of 30 nm, the thickness increasedto 40 µm after annealing under oxidizing conditions.

In conclusion, the electrical measurements provide evidence of the existence of a highlyresistive surface layer on top of the as-received crystal being responsible for the resistiveswitching.

3.2. Resistive switching on the nanoscale

Having seen that the resistive switching is related to the surface layer, we now proceedto investigate the spatial homogeneity of the resistivity of the surface layer. Therefore, weconducted a typical LC-AFM switching experiment on the as-received sample (1.4 at.% Nb)under HV conditions at 50 ◦C. First a region of 500 nm × 500 nm was scanned while a switchingvoltage of +4 V was permanently applied. After this, the same area was scanned again with areading voltage of 10 mV. The result is shown in figure 3(a). It can be seen that the wholescanned area has switched to the ON state with a significantly smaller resistance than theuntreated surface. Furthermore, it can be seen that the conductivity of the switched area isnot homogeneous but shows an array of conducting clusters with an average diameter ofaround 50 nm. The clusters are regularly aligned in rows oriented along the 〈100〉 crystalaxes, which is not an artefact of the scan procedure since this was excluded by switchingscans under different angles between tip movement and sample orientation. In order to furtherdemonstrate that the cluster-like switching behaviour is a property of the Nb-doped crystal itself,

New Journal of Physics 15 (2013) 103017 (http://www.njp.org/)

7

Single clusterArray of clustersa b

[010]

[100

]

0 250 nm0 750 nm

nA

400

0

0 750 nm 0 750 nm

de

c

1 G

2 G

1.5 M

1.0 M

as-received

switched at 200°C

100 k

050 100 150 200 250

T (°C)

switched at 50°C

R (

Ω)

-2 0 2

-10

0

10

I (µA

)

U (V)

150 °C

50 °C

175 °C

Figure 3. Resistive switching on the nanoscale. (a) LC-AFM images showing theswitching of an array of clusters by performing a continuous switching scan and(b) the switching of a single cluster by the application of the switching voltage(+4 V) in one point at 50 ◦C. (c) Local I–V characteristic during switching.(d) Back-switching of the array of clusters at higher temperatures. (e) Resistanceas a function of temperature calculated from LC-AFM scans for the as-receivedsurface and the ON states obtained by switching at 200 and 50 ◦C.

we attempted to switch one single point (marked by a cross in figure 3(b)) by applying theswitching voltage (+4 V) without moving the AFM tip. After this, not only the contact pointbut the whole cluster switched to the ON state proving that the nanoscale resistive switchingis not of a filamentary type, as in undoped SrTiO3, but reveals a cluster-like dimension. Thelocal I–V curve (figure 3(c)) measured at one point of the surface shows that the resistancecould be switched between two resistance states by electrical gradients comparable to themacroscopic measurements. The switching polarity corresponds to the so-called ‘eightwise’switching [45] and is in agreement with the switching scans conducted on Nb-doped thinfilms [7]. To check the stability of the ON state of the clusters, we increased the temperatureand obtained measurements of the switched array (figure 3(d)). Below 150 ◦C the conductivitydid not change at all, but at higher temperatures the clusters started to switch back. It canbe seen that the clusters always switched back as a whole and independently of each other.At 175 ◦C most of the clusters finally switched back after some time. The switching of theclusters by electrical gradients on the nanoscale is related to an insulator-to-metal transition,which is proved by the calculations of the resistance extracted from LC-AFM measurements atdifferent temperatures (figure 3(e)). The as-received surface had a very high resistance, which

New Journal of Physics 15 (2013) 103017 (http://www.njp.org/)

8

y: 1.6 µmx: 1.6 µm

z: 1

20 n

m

HF Etching

Topography

250

125

0

nA

0

1

2

µm

Current

3D

125 nm

Figure 4. Investigations of clusters. Topography and current measured byLC-AFM in an etch-pit showing the presence of three-dimensional (3D) clustersin the bulk.

clearly showed a semiconducting behaviour. After switching at 200 ◦C, the resistivity of theclusters is in a lower resistance state. The temperature dependence of the resistance showsa minimum resistance at approximately 225 ◦C comparable to the temperature dependenceobtained by the macroscopic measurement in figure 2(c). In contrast, the third graph, obtainedafter switching at 50 ◦C from the measurement in figure 3(b), shows a metallic behaviour upto 150 ◦C. At higher temperatures, these clusters switched back to the OFF state probably dueto re-oxidation effects (cf supplementary figure S3 and supplementary video 1, available fromstacks.iop.org/NJP/15/103017/mmedia). The different behaviour of the resistivity depending onthe conditions during switching indicates that the surface layer is highly variable under theapplied gradients.

In the next section we will examine the nature of the conducting clusters. In order to checkwhether these clusters are only related to the surface layer or also exist in the bulk, we etched thecrystal for 10 min in buffered hydrofluoric acid (12.5% HF) at 90 ◦C. During this procedure, etchpits evolved from the surface providing a ‘tomographic view’ into the bulk of the sample. Asshown in figure 4(a), the conductivity measured in one etch pit shows a conductivity pattern withclusters, which indicates that the clusters are present everywhere in the crystal. The clusters arealigned in rows at the sides of the etch pit revealing a long-range order in the conductivity. Sucha long-range order in the related density of states was previously found in Nb-doped SrTiO3 asmeasured by scanning tunnelling microscopy on cleaved crystals [46].

Summarizing the investigations on the nanoscale by means of LC-AFM, we foundregularly arranged conducting clusters that could be switched independently between the ONand OFF state. In contrast to the filamentary switching in undoped SrTiO3, in Nb-doped SrTiO3

a different cluster-like switching mechanism is present.

New Journal of Physics 15 (2013) 103017 (http://www.njp.org/)

9

OxidizedSr/Ti = 2.05Ø 4 µm

As-receivedSr/Ti = 1.31Ø 0.4 µm

ReducedSr/Ti = 0.03Ø 0.4 µm

ScrapedSr/Ti = 1.07Ø 8 µm

470 455eV

470 455eV

470 455eV

470 455eV

4+3+2+

Ti2p

Figure 5. Investigation of the surface layer. Illustration of a cross section throughthe sample before and after removing the surface layer by scraping with thecorresponding AFM topography images and Sr/Ti ratios and Ti 2p valencesobtained by XPS after application of different gradients.

3.3. Surface layer

Having seen that the surface layer of the crystal is responsible for the resistive switchingphenomena, we now want to focus on the fundamental physical properties of this layer. ByXPS measurements of the as-received surface we found an excess of Sr and no metallic statesat the Fermi edge, which is in agreement with the conclusions derived from the electricalmeasurements. However, by removing the surface in situ under UHV conditions by scrapingwith a diamond tip (figure 5), we obtained access to the bulk properties and we found aperfect stoichiometry with a Sr/Ti ratio close to 1 and we were able to measure indicationsof metallic states at the Fermi edge in the electronic structure, which is consistent with theresults on cleaved samples obtained by Haruyama et al [28]. We can thus conclude that thesamples were grown with a prefect stoichiometry but that a surface layer with propertiesdifferent than the bulk has evolved in air. Regarding the shape of the Ti 2p core line afterscraping, we can detect a contribution of the valence state +3, indicating that Nb doping is notmainly compensated by the creation of cation vacancies but by a change of the Ti valences,which can be expressed by the structural formula SrTi4+

1−xTi3+x Nb5+

x O3. The stoichiometry ofthe surface layer can be altered dramatically by different treatments (see also supplementaryfigure S8, available from stacks.iop.org/NJP/15/103017/mmedia). If the surface is reduced

New Journal of Physics 15 (2013) 103017 (http://www.njp.org/)

10

20

-10

0

40

20

10

-30

-20

x (n

m)

Sr-rich cluster O TiOTi Sr

z ' (nm)5 0 -5 -10 -15

1

10

insidecluster

Sr/

Ti

z (nm)Sr

aLDOS stoichiometric Sr-rich

-2 00

0.05

0.1

E - EF (eV)

ONb

Ti

Srb

Figure 6. Sr-rich surface layer (a) LDOS simulations of a stoichiometric andSr-rich surface. (b) Atom probe measurement of Sr-movement induced by FIB.

by heating for 24 h at 1100 ◦C under UHV conditions, only a small amount of Sr is left inthe surface layer. Simultaneously, the electronic structure changes considerably with Ti andNb mainly displaying valence +2, which is consistent with our previous findings [47]. Thisindicates that reduction can form Ti-rich phases at the surface. Effusion measurements duringreduction revealed that no Sr-related components are emitted from the sample indicatingthat Sr diffuses into the bulk of the crystal. Heating the sample under oxidizing conditionsinstead increases the amount of Sr and no additional valences are visible thus indicatingthe formation of Sr-rich micro crystals on the surface, as predicted by formula (1) and alsoproved by previous measurements [48, 49]. In order to understand the differences between theSr-rich as-received surface layer and the stoichiometric bulk, we performed ab initio calcu-lations of Nb-doped SrTiO3 in two different structural models (for details see supplementaryfigure S9, available from stacks.iop.org/NJP/15/103017/mmedia), which revealed that metallicstates at the Fermi edge are suppressed by an excess of Sr modelled by Ruddlesden–Popperlayers (figure 6(a)). This corresponds to the electrical measurements in figure 1 where weshowed that annealing under oxidizing conditions increases the thickness and the Sr contentof the surface layer and simultaneously suppresses the conductivity. In order to recon-struct the metallic properties of the surface layer, which would be the decisive step in the

New Journal of Physics 15 (2013) 103017 (http://www.njp.org/)

11

resistive switching process, a Sr movement has to take place. To illustrate the possibility ofSr movements under external gradients, we performed measurements using a 3D atom probe(figures 6(b) and S10, available from stacks.iop.org/NJP/15/103017/mmedia). A sharp needlewas cut out of the SrTiO3:Nb by a focused ion beam resulting in the formation of a Sr-richcluster at the end of the needle associated with the evolution of a Ti-rich region below. Thismeasurement confirms the possibility of Sr movements, which have also been reported inrelation to the formation of Ruddlesden–Popper phases under electrical [50] and thermal [51]gradients, as well as in relation to the segregation of SrO [48]. In general, we can concludethat the chemical composition of the surface layer is highly variable under chemical gradientsand in particular Sr-rich or Ti-rich phases can be formed, which we have to bear in mind whenestablishing a nanoscale model of resistive switching in the following.

4. Discussion and conclusions

Summarizing the results presented, we first of all found that the resistive switching is relatedto conducting clusters. Hence, the switching mechanism in Nb-doped SrTiO3 differs fromthe switching of extended filaments in undoped SrTiO3. Furthermore, we demonstrated byLC-AFM measurements that resistive switching is possible without a permanent electrodeindicating that fundamental changes take place during switching in the surface layer. Sincethe conducting clusters are present in Nb-doped SrTiO3 and exhibit higher conductivity thanthe surrounding matrix, one possible explanation for the evolution of the clusters could bethat the Nb content inside these clusters is slightly higher than in the surrounding matrix.Although the XRD measurements (supplementary figure S1) suggest that on average most ofthe Nb is distributed homogeneously, small fluctuations on the nanoscale could be responsiblefor the formation of the clusters. As demonstrated by density functional theory calculation(supplementary figure S7), such a Nb clustering could be possible in principle. This is supportedby the fact that we were able to detect Nb segregation effects on the macroscale in the outershell of several crystals related to the crystal growth process (supplementary figure S6). Usingan atom probe we demonstrated that due to ion beam irradiation not only the ratio of Sr andTi but also the Nb content can be changed illustrating the possibility of Nb segregation on thenanoscale (supplementary figure S10). A further indication of the possibility of Nb clusteringis that such donor clustering has been shown for Ba-doped SrTiO3 by calculations [52] andexperiments [53], and the existence of Nb clusters was also proposed [31] for Nb-dopedSrTiO3 thin films. An alternative explanation for the evolution of the clusters could be thatan agglomeration of Sr vacancies leading to a small local distortion of the lattice takes place,which was proposed earlier for non-stoichiometric SrTiO3 as well as for Nb-doped thin films[7, 32].

In order to model the resistive switching, not only the presence of clusters but also theexistence of the semiconducting surface, which was found on top of the metallic bulk, has tobe taken into account. Especially the ability to form Ti-rich or Sr-rich phases under appliedgradients such as reduction and oxidation is highly relevant since the application of electricfields during the resistive switching leads to the presence of comparable gradients on thenanoscale. Since the bulk was found to be fully metallic, resistive switching can only take inthe surface layer. Based on the observation that the conducting clusters can always be switchedas a whole and independently of each other, we assume that not the clusters themselves butrather the contact regions between the clusters, which we call bridges, are switching. Hence, we

New Journal of Physics 15 (2013) 103017 (http://www.njp.org/)

12

Potential Temperature

0 4 25 125(°C)(V)

40

nm Nano-Filament

Surface-Cluster

Bulk-Cluster

Figure 7. Model of the electroformation and the resistive switching of Ti-richnano-filaments between conducting clusters.

propose the following two-step model of resistive switching. While the clusters and bridgesof the bulk are always in the metallic state and are not affected by the switching process,the topmost clusters, which build up the semiconducting Sr-rich surface layer, have to betransformed to metallic clusters with perfect stoichiometry related to Sr/SrO movement viaa first electroforming process by applying a positive voltage to the AFM tip. When the clustersare in the metallic state, the greatest drop in potential occurs in the region between the clustersdisplaying higher resistivity. We simulated this situation by a finite element method, as shownin figure 7, by calculating the potential and temperature during switching. It was assumed thattwo conducting clusters are connected by a bridge with higher resistivity. The simulation revealsthat in this case the entire potential drops in the region between the clusters, which in the modelwas only a few nanometres in size. This leads to the presence of high electric fields, althoughthe temperature only increases slightly indicating that thermal effects play a minor role duringthe switching process. Hence we assume that ionic movements take place caused by electricgradients. Since we have seen that Sr/SrO is highly mobile under applied gradients, we concludethat the high electric field in the region between the clusters leads to the formation of a Ti-richbridge, probably consisting of one or several TinO2n−1 phases. Once such a bridge, which canbe regarded as a nano-filament, is formed, the resistive switching can be explained by takinginto account oxygen movements switching the bridge from a highly conducting TinO2n−1 phaseto a poorly conducting phase and vice versa, depending on the polarity of the applied voltage.In comparison to undoped SrTiO3, in which filaments related to extended defects were foundto be responsible for the switching, Nb doping leads to the presence of clusters either directlyby Nb clustering or indirectly by the formation of clustering vacancies as described above. Thissuppresses the switching mechanism along the extended defects and confines the switching toa region between the topmost conducting clusters in the surface layer, where a nano-filament isformed. According to this description, the Nb itself only plays a minor role during the switchingprocess.

The existence of such titanium oxide phases would also explain the characteristictemperature-dependent resistance curves measured after electrodegradation on the macro- andthe nanoscale showing semiconducting behaviour up to 250 ◦C and metallic behaviour at highertemperatures (figures 2 and 3). On the one hand, this behaviour could be explained classicallyby a serial connection of the semiconducting contact resistance and the metallic bulk, but we

New Journal of Physics 15 (2013) 103017 (http://www.njp.org/)

13

believe, especially because we found the same temperature dependence of the resistance onthe macro- and on the nanoscale, that it represents a Mott transition between the insulator andmetal. It is well known that such a transition occurs in Magneli phases and especially that Ti2O3

undergoes a Mott transition between 200 and 300 ◦C [42], which leads us to conclude that thisphase was created during the electrodegradation and then affected the resistance of the wholesample. Hence, we regard the measurement of the characteristic resistance curve as a strongindication of the formation of Ti-rich phases in SrTiO3:Nb under electrical gradients.

In summary, the resistive switching in the nominally metallic system SrTiO3:Nb canonly be understood if the special role of the intrinsic Sr-rich semiconducting surface layer isconsidered. On the nanoscale, conducting and switchable clusters are present indicating that theswitching mechanism in SrTiO3:Nb differs from the switching of extended filaments in undopedSrTiO3. We proposed a model suggesting that switching is a very complex local phenomenonrelated to the presence of conducting clusters and the generation of secondary phases as nano-filaments. Such a description is not limited to SrTiO3:Nb but may also be relevant for otheroxides, in which a local filamentary switching mechanism is present.

Acknowledgments

We thank R Borowski, J Friedrich, M Gerst, M Grates, S Masberg and T Possinger fortechnical support. Furthermore, we thank A Besmehn, M Ermrich and E Wurtz for contributingmeasurements. Finally, we gratefully acknowledge J Mayer and P Meuffels for fruitfuldiscussions. This work was supported in part by the Deutsche Forschungsgemeinschaft(SFB 917).

References

[1] Schooley J F, Hosler W and Cohen M 1964 Phys. Rev. Lett. 12 474–5[2] Beck A, Bednorz J G, Gerber C, Rossel C and Widmer D 2000 Appl. Phys. Lett. 77 139–41[3] Oligschlaeger R, Waser R, Meyer R, Karthauser S and Dittmann R 2006 Appl. Phys. Lett. 88 042901[4] Szot K, Speier W, Bihlmayer G and Waser R 2006 Nature Mater. 5 312–20[5] Tufte O N and Chapman P W 1967 Phys. Rev. 155 796–802[6] Nakamura H, Takagi H, Inoue I H, Takahashi Y, Hasegawa T and Tokura Y 2006 Appl. Phys. Lett. 89 133504[7] Munstermann R, Dittmann R, Szot K, Mi S, Jia C L, Meuffels P and Waser R 2008 Appl. Phys. Lett. 93 023110[8] Sim H et al 2005 IEDM Tech. Dig. 777–80[9] Seong D J, Jo M, Lee D and Hwang H 2007 Electrochem. Solid-State Lett. 10 H168–70

[10] Lee J et al 2010 Curr. Appl. Phys. 10 e68–70[11] Rana K G, Khikhlovskyi V and Banerjee T 2012 Appl. Phys. Lett. 100 213502[12] Park C, Seo Y, Jung J and Kim D W 2008 J. Appl. Phys. 103 054106[13] Shang D S, Sun J R, Shi L and Shen B G 2008 Appl. Phys. Lett. 93 102106[14] Zhang X T, Yu Q X, Yao Y P and Li X G 2010 Appl. Phys. Lett. 97 222117[15] Seong D J, Lee D, Pyun M, Yoon J and Hwang H 2008 Japan. J. Appl. Phys. Part 2 47 8749–51[16] Kan D, Kan K and Shimakawa Y 2010 Thin Solid Films 518 3246–49[17] Chen Y L, Wang J, Xiong C M, Dou R F, Yang J Y and Nie J C 2012 J. Appl. Phys. 112 023703[18] Chen X G, Ma X B, Yang Y B, Chen L P, Xiong G C, Lian G J, Yang Y C and Yang J B 2011 Appl. Phys.

Lett. 98 122102[19] Gwon M, Lee E, Sohn A, Bourim E M and Kim D W 2010 J. Korean Phys. Soc. 57 1432–6[20] Bourim E M and Kim D W 2013 Curr. Appl. Phys. 13 505–9

New Journal of Physics 15 (2013) 103017 (http://www.njp.org/)

14

[21] Shen J X, Qian H Q, Wang G F, An Y H, Li P G, Zhang Y, Wang S L, Chen B Y and Tang W H 2013 Appl.Phys. A 111 303–8

[22] Zhong S and Cui Y 2013 Curr. Appl. Phys. 13 913–8[23] Buzio R, Gerbi A, Gadaleta A, Anghinolfi L, Bisio F, Bellingeri E, Siri A S and Marre D 2012 Appl. Phys.

Lett. 101 243505[24] Sun J, Jia C H, Li G Q and Zhang W F 2012 Appl. Phys. Lett. 101 133506[25] Zhang P, Meng Y, Liu Z, Li D, Su T, Meng Q, Mao Q, Pan X, Chen D and Zhao H 2012 J. Appl. Phys.

111 063702[26] Sawa A 2008 Mater. Today 11 28–36[27] Han L Q, Kaimai A, Yashiro K, Nigara Y, Kawada T, Mizusaki J, Chen P, Zhong C and Higuchi T 2004 Solid

State Ion. 175 431–5[28] Haruyama Y, Aiura Y, Bando H, Suzuki H and Nishihara Y 1997 Physica B 237–8 380–2[29] Kanzig W 1955 Phys. Rev. 98 549–50[30] Toyoda H and Itakura M 1962 J. Phys. Soc. Japan 17 924–31[31] Zhu Y, Ma X, Li D, Lu H, Chen Z and Yang G 2005 Acta Mater. 53 1277–84[32] Tokuda Y, Kobayashi S, Ohnishi T, Mizoguchi T and Shibata N 2011 Appl. Phys. Lett. 99 033110[33] Akhtar M J, Akhtar Z, Jackson R A and Catlow C R 1995 J. Am. Ceram. Soc. 78 421–8[34] Szot K, Speier W, Herion J and Freiburg C 1997 Appl. Phys. A 64 55–9[35] Gunhold A, Gomann K, Beuermann L, Frerichs M, Borchardt G, Kempter V and Maus-Friedrichs W 2002

Surf. Sci. 507 447–52[36] Blennow P, Hagen A, Hansen K K, Wallenberg L R and Mogensen M 2008 Solid State Ion. 179 2047–58[37] Witek S, Smyth D M and Pickup H 1984 J. Am. Ceram. Soc. 67 372–5[38] Fujimoto M and Watanabe M 1985 J. Mater. Sci. 20 3683–90[39] Hung K and Yang W 2003 Mater. Sci. Eng. A 351 70–80[40] Gunhold A, Beuermann L, Frerichs M, Kempter V, Gomann K, Borchardt G and Maus-Friedrichs W 2003

Surf. Sci. 523 80–8[41] Kolodiazhnyi T and Petric A 2005 J. Electroceram. 15 5–11[42] Szot K, Rogala M, Speier W, Klusek Z, Besmehn A and Waser R 2011 Nanotechnology 22 254001[43] Necas D and Klapetek P 2012 Cent. Eur. J. Phys. 10 181–8[44] Menzel S, Waters M, Marchewka A, Bottger U, Dittmann R and Waser R 2011 Adv. Funct. Mater. 21 4487–92[45] Muenstermann R, Menke T, Dittmann R and Waser R 2010 Adv. Mat. 22 4819[46] Guisinger N P, Santos T S, Guest J R, Chien T Y, Bhattacharya A, Freeland J W and Bode M 2009 ACS Nano

3 4132–6[47] Szot K and Speier W 1999 Phys. Rev. B 60 5909–26[48] Szot K, Speier W, Breuer U, Meyer R, Szade J and Waser R 2000 Surf. Sci. 460 112–28[49] Meyer R, Szot K and Waser R 1999 Ferroelectrics 224 323–9[50] Meyer D, Levin A, Leisegang T, Gutmann E, Paufler P, Reibold M and Pompe W 2006 Appl. Phys. A 84 31–5[51] Szot K, Pawelczyk M, Herion J, Freiburg C, Albers J, Waser R, Hulliger J, Kwapulinski J and Dec J 1996

Appl. Phys. A 62 335–43[52] Fuks D, Dorfman S, Piskunov S and Kotomin E 2005 Phys. Rev. B 71 14111[53] Rossell M D, Ramasse Q M, Findlay S D, Rechberger F, Erni R and Niederberger M 2012 ACS Nano

6 7077–83

New Journal of Physics 15 (2013) 103017 (http://www.njp.org/)

Supplementary information

to

Cluster-like resistive switchingof SrTiO3:Nb surface layers

C. Rodenbucher, W. Speier, G. Bihlmayer, U. Breuer, R. Waser & K. Szot

2

I. LITERATURE MODELS OF RESISTIVE SWITCHING IN SrTiO3:Nb

Based on investigations of the resistive switching effect in SrTiO3 thin films and single crystals

various models were developed but the actual nature of the switching process is still under dis-

cussion. On metal/SrTiO3:Nb junctions some groups reported an area-dependence of the resistive

switching1,2 and they tried to explain the switching by the change of the Schottky barrier induced

by charge trapping at the interface3–5. Other groups presented measurements that exclude such

a Schottky model and they proposed local models such as the formation of conducting paths by

oxygen vacancies or inhomogeneities in the chemical composition6–10. Furthermore, it has been

suggested that an interface layer is responsible for the switching11, which has also been found in

contacts with superconductors12,13.

II. CRYSTALLOGRAPHIC STRUCTURE

FIG. S 1. Lattice parameter as a function of doping concentration obtained by Rietveld refinement of X-raydiffraction patterns.

As the starting point for the investigation of the single crystals, we analysed the influence

of niobium doping on the crystallographic structure by performing XRD measurements of single

crystalline powders with three different doping concentrations 0.2 at%, 1.4 at% and 10.1 at%. It was

found that all samples showed a pattern, which is typical of the perovskite structure representing

a cubic system with the space group Pm3m and the powder diffraction file (PDF) number 010-70-

8508. The lattice parameter, which was calculated by a refinement procedure, is in general higher

than for the undoped SrTiO3 and increases linearly with niobium doping (Fig. S 1).

3

III. SWITCHING OF CLUSTERS

FIG. S 2. LC-AFM scan before and after switching in a single spot.

In Figure 3b it is shown that applying a voltage to one spot results in the switching of an area

with a diameter of approx. 100 nm. If we have a closer look at the surface before and after this

switching (Fig. S 2), it becomes apparent that the switched area consists of several clusters. For

our model, this implies that during the switching not only the bridge between the surface cluster

and bulk cluster but also the bridges between several surface clusters can change their phase into

the metallic state. To estimate the local resistance, linescans in the centres of the images were

obtained with a scan voltage of 0.1 V in the OFF state and 0.01 V in the ON state showing that

the resistance switches by more than a factor of 104.

4

IV. SWITCHING OF AN ARRAY OF CLUSTERS

FIG. S 3. Time and temperature dependence of the array of clusters (see description in main article). (a)LC-AFM scans. (b) Arrhenius plot of the resistance of ON and OFF states calculated from the LC-AFMmaps. (c) Switching back to the OFF state at 175 ∘C as a function of time. (d) Resistance of the as-receivedsample showing semiconducting behaviour.

In Figure S 3, the temperature-dependent measurement of the switched array of clusters is

investigated in more detail. We extracted the resistance from the ON and the OFF regions of the

LC-AFM scans (Fig. S 3a) and plotted this as an Arrhenius graph. The resistance of the ON state

is almost constant up to 125 ∘C indicating metallic behaviour. In contrast to this, the as-received

cleaved sample also shows conducting clusters (cf. Fig. S 4), but they are n-type semiconducting

over the entire temperature range with an activation energy of 0.41 eV as shown in Figure S 3d,

which proves that a semiconductor-to-metal transition takes place during the switching process.

In Figure S 3c, the resistance of the switched area at 175 ∘C is presented as a function of time

illustrating the re-switching. The resistance increases exponentially in the first 6 minutes after

reaching the temperature when most of the clusters switch back. After this, the slope of the graph

changes representing the behaviour of the clusters in the upper part of the scan which remain in

the ON state for quite a long time. The different behaviour of the clusters could be related to the

initial writing scan, in which the upper part of the area was switched first by the scanning tip.

5

V. CLUSTERS ON CLEAVED SURFACE

FIG. S 4. LC-AFM scan (U = 0.1 V) on cleaved surface showing regularly aligned clusters.

LC-AFM measurements on the cleaved surface were conducted to investigate the origin of the

conducting clusters. Therefore the single crystal was cleaved ex situ perpendicular to the (100)

plane and afterwards the local conductivity was investigated under vacuum conditions. As shown

in Figure S 4, conducting clusters with a size of 40 - 50 nm are visible, which is comparable to

the distribution of the conductivity on the epi-polished surface (Fig. 3). This measurement clearly

proves that the clusters already develop during crystal growth and are not caused by surface

preparation such as polishing. Furthermore, the regular alignment of the clusters along the crystal

axes, which we presented for the etched (Fig. 4) and the switched surface (Fig. 3), can also be

observed on the cleaved surface.

6

VI. SWITCHING OF SINGLE DISLOCATIONS

FIG. S 5. Switching of a single dislocation in the centre of an etch pit.

In undoped SrTiO3 it was found that single dislocations can be switched by applying a voltage

via the AFM tip. As presented in the main article by etching experiments, a large number of dis-

locations are also present in Nb-doped SrTiO3. Most of the etch pits which evolve during etching

with hydrofluoric acid show a cluster-like conductivity pattern at the sides, but due to inhomo-

geneities of the crystal there are some few etch pits where the sides are only poorly conducting,

which gives us the opportunity to investigate the dislocation itself. As shown in Figure S 5, the

centre of these etch pits can be switched by applying a voltage between the ON state and the OFF

state because of the movement of ions (e.g. Sr vacancies) along the dislocation serving as an easy

diffusion path. This measurement clearly shows that the behaviour of the dislocations in Nb-doped

SrTiO3 is very similar to that in undoped SrTiO315 but this switching mechanism is superimposed

by the presence of conducting clusters.

7

VII. NIOBIUM SEGREGATION: EDX

FIG. S 6. Sr/Ti and Nb/Ti ratios obtained by EDX measurements at different positions from the edge (x= 0 mm) to the centre of the crystal.

As described in the main article, we assume that the clusters arise in the crystal during growth

because of segregation effects of Nb. Although we have not had the opportunity to measure

the atomic composition of the clusters on the nanoscale, we were able to demonstrate that such

segregation effects exist in principle especially on the macroscale. A closer look at the edge of

the crystal revealed that it was transparent and became darker towards the centre of the sample

(photographic image in Fig. S 6). Using energy-dispersive X-ray spectroscopy (EDX) to investigate

the chemical composition at different distances from the edge, it became apparent that the colour

is directly related to the Nb content. In the transparent region, the Nb/Ti ratio was smaller by

a factor of three than in the dark-blue part. The origin of this segregation could be temperature

gradients between the shell and the core of the crystal during growth.

VIII. NIOBIUM SEGREGATION: ENERGY CALCULATION

Since the clusters were detected in SrTiO3:Nb and are related to differences in the conductivity,

we assume that the Nb content inside the cluster is higher than in the surrounding matrix. By EDX

measurements, we confirmed that such Nb segregation effects are present on the macroscale (cf.

Supplementary Figure S 6) and we performed ab initio calculations to check whether Nb clustering

would be possible in principle on the nanoscale. Therefore, we estimated the energy cost needed

8

FIG. S 7. Calculation of the energy for different inhomogeneous arrangements of Nb revealing the possibilityof Nb clustering. (NN denotes the number of nearest neighbours)

to create local inhomogeneities by performing DFT calculations in the local density approximation

(for a program description see http://www.flapw.de). In this study, the full-potential linearised

augmented planewave method as implemented in the Fleur-code16 was used. We assumed different

distributions of Nb in a 2x2x4 bulk unit cell of SrTiO3 at 0 K. Based on electrostatic arguments,

a homogeneous distribution of Nb atoms in SrTiO3 would be expected. Indeed, the calculations

show that configurations with many Nb atoms at nearest-neighbour positions are highest in energy

and spatially separated arrangements can be up to 300 meV lower in energy (Fig. S 7, circles). This

is, however, only valid as long as the SrTiO3 matrix is kept frozen in. When we allow the atoms

to relax around the impurities, we see that different configurations are affected quite differently

(Fig. S 7, diamonds), leading to quite similar energies for both separated Nb atoms and clustered

species. This can be rationalized considering that the strain field around a Nb cluster is smaller

than the accumulated strain fields of the same amount of isolated Nb atoms. In summary, without

taking temperature effects into account, we obtain a rather flat energy-profile for the Nb atoms that

allows substantial local inhomogeneities in the SrTiO3 matrix, which could be generated during

crystal growth. The importance of entropic contributions will, of course, depend on the actual

growth temperature.

9

IX. SURFACE TREATMENTS

FIG. S 8. XPS, LEED and AFM measurements of surfaces prepared by different techniques.

The influence of different preparation techniques on the surface was investigated using XPS

with an AlK𝛼 monochromator. In Figure S 8, additionally the LEED pattern and the topography

measured by AFM are shown. The topography shows huge changes of the surface layer. After

reduction, a stripe-like structure is visible, which is also present after oxidation, but additional

islands evolved resulting in the appearance of stripes in the LEED pattern. Etching leads to the

formation of etch pits, as already discussed in the main article. Leaching alone does not have a

great influence on the topography of the surface. Regarding the LEED pattern measured using an

10

electron beam of 96 eV, it can be seen that the 1 x 1 symmetry is present in all measurements

proving that the perovskite structure is very stable irrespective of the changes in the topography.

X. LOCAL DENSITY OF STATES

FIG. S 9. Local densities of states (LDOS) of a symmetric TiO2-terminated SrTiO3 film with Nb at thecentre as described in the text (A). The lower left panels show for each layer (starting from the surface (S)to the centre (C)) the LDOS of the atoms: Ti (grey), Nb (yellow), Sr (blue) and O (red). If two atoms ofthe same species are located in the same layer, the DOS of the second atom is shown as a dashed line. Incolumn B, the LDOS of a film with a SrO-SrO stacking fault is shown. Layers containing Nb are shownin the bottom panels, the surface LDOS at the top. The upper left panel gives a direct comparison ofthe surface Ti DOS in the film without (black) and with (grey, shaded) stacking fault. On the right, anillustration of the calculated cells is shown.

To investigate the influence of Nb doping on the electronic structure of SrTiO3, we performed

LDOS calculations as described in the main article. We simulated a symmetric, 13-layer TiO2-

terminated film with a 𝑐(2 × 2) in-plane unit cell, containing a single Nb atom at the centre that

replaces Ti. Structurally, the films were fully relaxed. As can be seen from the lower left panels of

Fig. S 9, the Fermi level is located in the conduction band where the Nb 𝑑-electrons are situated.

11

With increasing distance from the Nb, the local density of states (LDOS) drops quickly and is

reduced by two orders of magnitude at the surface.

From a similar calculation, with two SrO-SrO Ruddlesden-Popper-type stacking faults, which

modelled the Sr-rich surface of the as received sample, we can see that the Nb states become further

confined in the middle of the film and the DOS at the Fermi level is reduced even more strongly

with distance from the Nb defect (column B of Fig. S 9). A direct comparison of the Ti LDOS

at the surface (upper left panel of Fig. S 9) shows another reduction by an order of magnitude as

compared to the film without the SrO-SrO stacking fault. This proves that the donor doping effect

of the Nb can be effectively suppressed by an excess of Sr.

12

XI. ATOM PROBE

FIG. S 10. Atom probe measurements. (a) Maps of the atomic distribution measured by the atom probetechnique. (b) SEM image of the tip prepared by focused ion beam. (c) Proxigram at the position of theSr-rich cluster. (d) Illustration of Sr movement due to focused ion beam.

In our model, we assume that the semiconducting clusters can be transformed into metallic

clusters via a movement of Sr. Such Sr movements have been found by several experiments including

an atom probe (cf. Fig. 6). Here, we present this measurement in more detail. A nanoscale tip

of Nb-doped SrTiO3 was cut out of the single crystal by a focused ion beam. Then the atomic

composition of this tip was recorded by a CAMECA LEAP4000 X HR instrument using UV-laser

pulsing. As shown in Figure S 10, a Sr-rich region evolved at the end of the tip due to electrical

or thermal gradients during sample preparation, which is illustrated in the proxigram profile (Fig.

S 10c) showing the Sr/Ti ratio along a linescan from the inside of the cluster to the outside. Similar

behaviour was also found using SIMS-measurements on undoped SrTiO314. Furthermore, it can

be seen that the Nb content at the end of the tip was slightly decreased due to the evolution of the

Sr-rich region revealing that Nb segregation effects as assumed to be responsible for the existence

of the conducting clusters are in principle possible.

13

REFERENCES

1 Sim H et al. 2005 In IEDM Tech. Dig. 758 –612 Sawa A 2008 Mater. Today 11, 28–363 Seong D J, Jo M, Lee D and Hwang H 2007 Electrochem. Solid-State Lett. 10 H168–704 Lee J et al. 2010 Curr. Appl Phys. 10 e68–705 Rana K G, Khikhlovskyi V. and Banerjee T 2012 Appl. Phys. Lett. 100 2135026 Seong D J, Lee D, Pyun M, Yoon J and Hwang H 2008 Jpn. J. Appl. Phys., Part 2 47 8749–517 Shang D S, Sun J R, Shi L and Shen B G 2008 Appl. Phys. Lett. 93 1021068 Kan D, Kan K and Shimakawa Y 2010 Thin Solid Films 518 3246–499 Zhang X T, Yu Q X, Yao Y P and Li X G 2010 Appl. Phys. Lett. 97 222117

10 Chen Y L, Wang J, Xiong C M, Dou R F, Yang J Y and Nie J C 2012 J. Appl. Phys. 112 02370311 Park C, Seo Y, Jung J and Kim D W 2008 J. Appl. Phys 103 05410612 Mannhart J, Bednorz J G, Muller K and Schlom D 1991 Z. Phys. B 83 307–1113 Hasegawa H, Fukazawa T and Aida T 1989 Jpn. J. Appl. Phys., Part 2 28 L2210–214 Szot K, Speier W, Herion J and Freiburg C 1997 Appl. Phys. A 64, 55–5915 Szot, K., Speier, W., Bihlmayer, G. & Waser, R. 2006 Nat. Mater 5, 312–32016 Perdew J P and Zunger A 1981 Phys. Rev .B 23 5048–79


Recommended