12
Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces D. Vu, 1 S. Arscott, 2 E. Peytavit, 2 R. Ramdani, 3,4 E. Gil, 3,4 Y. André, 3,4 S. Bansropun, 5 B. Gérard, 6 A. C. H. Rowe, 1 and D. Paget 1, * 1 Laboratoire de Physique de la Matière Condensée, Ecole Polytechnique–CNRS, 91128 Palaiseau Cedex, France 2 Institut d’Electronique de Microélectronique et de Nanotechnologie (IEMN), University of Lille–CNRS, Avenue Poincaré, Cité Scientifique, 59652 Villeneuve d’Ascq, France 3 LASMEA, Clermont-Université–Université Blaise Pascal, BP 10448, F-63000 Clermont-Ferrand, France 4 LASMEA, CNRS, UMR 6602, F-63177 Aubière, France 5 THALES Research and Technology France, Route départementale 128, 91767 Palaiseau Cedex, France 6 ALCATEL-THALES III-V Lab, THALES Research and Technology France, Route départementale 128, 91767 Palaiseau Cedex, France Received 21 May 2010; revised manuscript received 26 July 2010; published 29 September 2010 The tunnel photocurrent between a gold surface and a free-standing semiconducting thin film excited from the rear by above band-gap light has been measured as a function of applied bias, tunnel distance, and excitation light power. The results are compared with the predictions of a model which includes the bias dependence of the tunnel barrier height and the bias-induced decrease in surface recombination velocity. It is found that i the tunnel photocurrent from the conduction band dominates that from surface states. ii At large tunnel distance, the exponential bias dependence of the current is explained by that of the tunnel barrier height while at small distance, the change in surface recombination velocity is dominant. DOI: 10.1103/PhysRevB.82.115331 PACS numbers: 85.30.Hi, 85.75.d, 73.40.c I. INTRODUCTION Spin injection from GaAs under light excitation into a magnetic metal is of fundamental interest for spintronics 1 and spin-polarized scanning tunneling microscopy. 2 In con- trast with injection from magnetic tips, 3 the use of semicon- ducting injectors permits rapid optical control of the spin of the injected electrons and minimizes the magnetic interac- tions between the injector and the surface. Some attempts at spin injection from GaAs tips into magnetic surfaces have already been made 4,5 but these studies used direct light exci- tation of the tip apex and a parasitic dependency of the in- jected current on the light helicity as high as several percent was observed. 6 This deleterious effect, attributed to helicity- dependent light scattering in the tunneling gap, seriously lim- ited the use of the GaAs-tip injectors. It has since been pro- posed that spin injectors should operate in transmission mode, with light excitation incident on the planar back sur- face of the injector. 7,8 In order to understand the features of spin injection, it is first necessary to understand the mechanisms of charge in- jection via tunneling from a photoexcited semiconductor into a metal. To our knowledge, despite the large number of stud- ies of photoelectric effects in metal-semiconductor junctions 9,10 and tunnel microscopes, 11 a complete under- standing of photoelectrical processes is still lacking. A pre- vious study using silicon tips found that the dominant pro- cess is a Fowler-Nordheim-type one. 12 Perhaps the most detailed investigation of tunneling of photoexcited electrons into a metal was performed by Jansen et al. 13,14 These work- ers considered light excitation at the front surface and as- sumed that electrons tunnel from midgap surface states thereby obtaining good agreement with experimental results for values of the bias applied to the metal smaller than about 0.5 V. These studies considered an energy-independent den- sity of surface states and bias-independent surface recombi- nation and tunnel barrier height. The surface quasi-Fermi level was determined using charge conservation and the tun- nel photocurrent was obtained, using current conservation, as a balance between the injected photocurrent and the Schottky current. Injection of free carriers across a semiconductor- liquid interface has also considered, 15 with tunneling from the conduction or valence band accounted for. However, in this case the only applied bias was the constant photovoltage. In this work, the tunnel photocurrent into a gold surface is measured as a function of bias, film/surface distance, and light excitation power. The GaAs film is a free-standing can- tilever having a thickness of a few micrometers. 16 We con- sider thin GaAs films photoexcited from the rear face and held at a controlled distance from the metal. This configura- tion brings two simplifications to the understanding of the results: i since light excitation is performed from the rear of the film, the injected photocurrent originates from electrons created near this surface which have been diffused to the front surface. Unlike front surface excitation, 17 this photocur- rent does not directly depend on the width of the depletion layer. ii The use of a film rather than a tip or a sharp point ensures that the contact surface is relatively large thereby avoiding the effects of a complex electric field distribution near the tip apex. 13 Analysis of the results has allowed us to eliminate the effect of possible distance inhomogeneities so that, after correction, the metal-semiconductor interface can be considered as planar, in the sense of a parallel plate ca- pacitor. The results are analyzed using a new model which incor- porates photovoltage, 10 bias dependence of surface recombination 18 and tunnel barrier height 19 and energy de- pendence of the density of surface states 20 For a gold non- magnetic surface, the interpretation of the results is simpler as the density of empty states in the metal depends only weakly on energy. 21 The quantitative agreement between the model and the experimental data demonstrates that, in the present case, the tunnel photocurrent originates from the con- PHYSICAL REVIEW B 82, 115331 2010 1098-0121/2010/8211/11533112 ©2010 The American Physical Society 115331-1

Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces

  • Upload
    d

  • View
    214

  • Download
    1

Embed Size (px)

Citation preview

Page 1: Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces

Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces

D. Vu,1 S. Arscott,2 E. Peytavit,2 R. Ramdani,3,4 E. Gil,3,4 Y. André,3,4 S. Bansropun,5 B. Gérard,6 A. C. H. Rowe,1 andD. Paget1,*

1Laboratoire de Physique de la Matière Condensée, Ecole Polytechnique–CNRS, 91128 Palaiseau Cedex, France2Institut d’Electronique de Microélectronique et de Nanotechnologie (IEMN), University of Lille–CNRS, Avenue Poincaré,

Cité Scientifique, 59652 Villeneuve d’Ascq, France3LASMEA, Clermont-Université–Université Blaise Pascal, BP 10448, F-63000 Clermont-Ferrand, France

4LASMEA, CNRS, UMR 6602, F-63177 Aubière, France5THALES Research and Technology France, Route départementale 128, 91767 Palaiseau Cedex, France

6ALCATEL-THALES III-V Lab, THALES Research and Technology France, Route départementale 128, 91767 Palaiseau Cedex, France�Received 21 May 2010; revised manuscript received 26 July 2010; published 29 September 2010�

The tunnel photocurrent between a gold surface and a free-standing semiconducting thin film excited fromthe rear by above band-gap light has been measured as a function of applied bias, tunnel distance, andexcitation light power. The results are compared with the predictions of a model which includes the biasdependence of the tunnel barrier height and the bias-induced decrease in surface recombination velocity. It isfound that �i� the tunnel photocurrent from the conduction band dominates that from surface states. �ii� At largetunnel distance, the exponential bias dependence of the current is explained by that of the tunnel barrier heightwhile at small distance, the change in surface recombination velocity is dominant.

DOI: 10.1103/PhysRevB.82.115331 PACS number�s�: 85.30.Hi, 85.75.�d, 73.40.�c

I. INTRODUCTION

Spin injection from GaAs under light excitation into amagnetic metal is of fundamental interest for spintronics1

and spin-polarized scanning tunneling microscopy.2 In con-trast with injection from magnetic tips,3 the use of semicon-ducting injectors permits rapid �optical� control of the spin ofthe injected electrons and minimizes the magnetic interac-tions between the injector and the surface. Some attempts atspin injection from GaAs tips into magnetic surfaces havealready been made4,5 but these studies used direct light exci-tation of the tip apex and a parasitic dependency of the in-jected current on the light helicity as high as several percentwas observed.6 This deleterious effect, attributed to helicity-dependent light scattering in the tunneling gap, seriously lim-ited the use of the GaAs-tip injectors. It has since been pro-posed that spin injectors should operate in transmissionmode, with light excitation incident on the planar back sur-face of the injector.7,8

In order to understand the features of spin injection, it isfirst necessary to understand the mechanisms of charge in-jection via tunneling from a photoexcited semiconductor intoa metal. To our knowledge, despite the large number of stud-ies of photoelectric effects in metal-semiconductorjunctions9,10 and tunnel microscopes,11 a complete under-standing of photoelectrical processes is still lacking. A pre-vious study using silicon tips found that the dominant pro-cess is a Fowler-Nordheim-type one.12 Perhaps the mostdetailed investigation of tunneling of photoexcited electronsinto a metal was performed by Jansen et al.13,14 These work-ers considered light excitation at the front surface and as-sumed that electrons tunnel from midgap surface statesthereby obtaining good agreement with experimental resultsfor values of the bias applied to the metal smaller than about0.5 V. These studies considered an energy-independent den-sity of surface states and bias-independent surface recombi-

nation and tunnel barrier height. The surface quasi-Fermilevel was determined using charge conservation and the tun-nel photocurrent was obtained, using current conservation, asa balance between the injected photocurrent and the Schottkycurrent. Injection of free carriers across a semiconductor-liquid interface has also considered,15 with tunneling fromthe conduction or valence band accounted for. However, inthis case the only applied bias was the constant photovoltage.

In this work, the tunnel photocurrent into a gold surface ismeasured as a function of bias, film/surface distance, andlight excitation power. The GaAs film is a free-standing can-tilever having a thickness of a few micrometers.16 We con-sider thin GaAs films photoexcited from the rear face andheld at a controlled distance from the metal. This configura-tion brings two simplifications to the understanding of theresults: �i� since light excitation is performed from the rear ofthe film, the injected photocurrent originates from electronscreated near this surface which have been diffused to thefront surface. Unlike front surface excitation,17 this photocur-rent does not directly depend on the width of the depletionlayer. �ii� The use of a film rather than a tip or a sharp pointensures that the contact surface is relatively large therebyavoiding the effects of a complex electric field distributionnear the tip apex.13 Analysis of the results has allowed us toeliminate the effect of possible distance inhomogeneities sothat, after correction, the metal-semiconductor interface canbe considered as planar, in the sense of a parallel plate ca-pacitor.

The results are analyzed using a new model which incor-porates photovoltage,10 bias dependence of surfacerecombination18 and tunnel barrier height19 and energy de-pendence of the density of surface states20 For a gold �non-magnetic� surface, the interpretation of the results is simpleras the density of empty states in the metal depends onlyweakly on energy.21 The quantitative agreement between themodel and the experimental data demonstrates that, in thepresent case, the tunnel photocurrent originates from the con-

PHYSICAL REVIEW B 82, 115331 �2010�

1098-0121/2010/82�11�/115331�12� ©2010 The American Physical Society115331-1

Page 2: Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces

duction band. For a large tunnel distance, the observed ex-ponential relationship of the tunnel photocurrent is caused bythe bias-dependent tunnel barrier height. At short distances,the dependence becomes nonexponential and is determinedby the bias-induced change in the surface recombination ve-locity.

This paper is organized as follows. Section II details themodel while the experimental results and procedure are pre-sented in Sec. III. Comparison between the model and theexperimental data is found in Sec. IV while Sec. V presentsa general discussion.

II. THEORY

The metal-semiconductor structure, described in Fig. 1, iscomposed of a p-type semiconductor film �thickness �, bandgap EG� and a metal to which a potential V is applied, sepa-rated by an insulating layer of thickness d and dielectricconstant �t. Light excitation from the rear of the semiconduc-tor creates a population of photoelectrons in the conductionband, with a fraction of these electrons being injected intothe space-charge region at the interface.

A. Generalities

The potential barrier at the semiconductor surface, definedby the energy difference between the top of the valence band

in the bulk and the bulk Fermi level at the surface, is givenby

�b = �0 + �� − qVs. �1�

Here �0 is the equilibrium value of �b, qVs is defined as theenergy difference between the electron quasi-Fermi level atthe surface and the bulk Fermi level, caused by light excita-tion and by the application of a bias. �q is the negative elec-tronic charge.� The energy �� is the shift of the electronquasi-Fermi level with respect to midgap caused by thechange in the surface charge. While a general calculation canreadily be performed, it will be assumed that �0 is equal tohalf the band-gap energy and that the density of surfacestates peaks at midgap.

In order to calculate ��, qVs, and the electron concentra-tion n0 at the onset of the depletion region, three conserva-tion equations are used. The first is the charge-neutralityequation,

�Qm + �Qsc + �Qss = 0, �2�

where the three terms are the departures from equilibrium ofthe charge densities at the metal, at the semiconductor sur-face, and in the semiconductor depletion layer. The two con-servation equations for the electron and hole current densi-ties are

Jp − Jte = Jr �3�

and

Js = J0 exp�−��

kT��exp�qVs

kT� − 1� = Jr − Jth, �4�

where Jp is the photocurrent density injected into the deple-tion region and Jr is the current density for electron-holesurface recombination. The tunnel current densities Jte andJth describe electron tunnel processes from the semiconduc-tor to the metal and hole processes from the metal to emptystates of the semiconductor, respectively. Js is the majoritycarrier Schottky current.9 Here J0=A��T2 exp�−

�0

kT � is theusual saturation current density where A�� is the effectiveRichardson constant, T is the temperature, and k is the Bolt-zmann constant.

B. Calculation of photoelectron concentrations at the surface

The electronic concentrations at the surface depend on ��and qVs for photoelectrons trapped in surface states while theconcentration ns of photoelectrons in the conduction banddepends on the concentration n0 at the onset of the depletionregion. The purpose of the present section is to calculatethese quantities from the conservation equations �Eqs.�2�–�4��. The quantities ns, n0, and qVs will be expressed as afunction of �� using the current conservation equationswhile �� will be found using the charge conservation equa-tion. This calculation proceeds in three distinct steps.

ld

ξsqV

qV

χ

pJ

sJ

FnE

FhE

GE

W

ϕϕ ∆+0

Light excitation

SemiconductorTunnelinggap

Metal

FE

ld

ξsqV

qV

χ

pJ

sJ

FnE

FhE

GE

W

ϕϕ ∆+0

Light excitation

SemiconductorTunnelinggap

Metal

FE

ld

ξsqV

qV

χ

pJ

sJ

FnE

FhE

GE

W

ϕϕ ∆+0

Light excitation

SemiconductorTunnelinggap

Metal

FE

FIG. 1. �Color online� Metal-semiconductor structure excited byabove band-gap light from the rear and for a positive bias V appliedto the metal. Also shown are the surface density of states, peaking atmidgap, and the energy difference �� between the electron quasi-Fermi level EFn at the surface and the Fermi level EF0 far from thejunction in equilibrium. The shaded areas are the surface stateslying between the electron and hole quasi-Fermi levels �for whichthe energy difference is qVs�. Also shown are the metal work func-tion, the semiconductor affinity, and the photocurrent �Jp� andSchottky current �Js�.

VU et al. PHYSICAL REVIEW B 82, 115331 �2010�

115331-2

Page 3: Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces

1. Dependence of n0 and of Jp on the surface recombinationvelocity S

As shown in Appendix A, these expressions, obtainedfrom a resolution of the one-dimensional diffusion equationin the semiconductor bulk are given by

n0 = �N0 �5�

and

Jp = qN0S� , �6�

respectively, with

� = �1 + S/vd�−1. �7�

The effective electron concentration N0 is proportional to thelight excitation power and the diffusion velocity vd is pro-portional to the ratio of diffusion constant and diffusionlength. Neither quantity depends on the surface recombina-tion velocity S or bias. Their expressions are given in Appen-dix A.

The calculation can be simplified if Jte�Jp. This assump-tion is valid provided the tunnel gap is not too small and willbe justified below by comparison with the experimental re-sults. In the opposite extreme case, the tunnel photocurrent isequal to the injected photocurrent �Jp� and the photovoltageis small. Similarly, it will be assumed that Jth�Js, so thatEqs. �3� and �4� give Jp=Js. Using this equation and furtherassuming in Eq. �4� that eqVs/kT�1, Eq. �6� becomes

qVs = qVs� + �� + kT ln�1 − �� . �8�

The quantity Vs�, defined by qVs

�=kT ln�qvdN0 /J0� is theusual value of the photovoltage �kT ln�Jp /J0�� in the limitwhere S�vd. With respect to most studies performed usinglight excitation at the front surface,10 the transmission geom-etry strongly simplifies the expression for the surface recom-bination dependence of the photovoltage. Assuming thermo-dynamic equilibrium between bulk and surface, it isstraightforward to calculate the electron concentration ns atthe surface. This concentration mostly lies at the energy ofthe lowest quantized state in the surface depletion layer. Herethis energy lies above the bottom of the conduction band andis written as f��b, where22

f� 1

�b�q22Eef f

2

2m� �1/3

�3/4�2/3 �9�

and Eef f is the surface electric field. Assuming that the quasi-Fermi-level position is independent of position10,23 and usingEqs. �4� and �5� one finds

ns =A��T2

qS� qSn0

A��T2� f�

. �10�

2. Dependence of the surface recombination velocity on��

Neglecting recombination in the depletion layer and con-sidering the usual Stevenson-Keyes expression for Jr,

18,24

Eq. �3� becomes, again assuming that Jte�Jp

N0S� = EFh

EFn

NT�E��n�pvnvpni

2�eqVs/kT − 1��nvn�ns + nts� + �pvp�ps + pts�

dE .

�11�

It is assumed that the density of surface states NT�E� has amaximum NT�0� at midgap and a typical width 0.2 eV.20

Here, �n and �p are the electron and hole capture cross sec-tions, of respective velocities vn and vp and ni is the intrinsicelectron concentration. The quantities ps, nts, and pts are,respectively, the surface hole concentrations and the valuesthat ns and ps would have if the surface Fermi level were atenergy E.

As bulk and surface are in thermodynamic equilibrium,the second term in the denominator of Eq. �11� is generallysmaller than the first one. This also implies that hole recom-bination processes are less efficient than electron ones andthat the occupation probability is close to unity for all stateslying between the two quasi-Fermi levels.25 As a result, theonly states which contribute to surface recombination are ina relatively narrow range of typical width kT situated nearEFn. Using the standard room-temperature value of the in-trinsic density of states of the conduction band, one finds thatns�nts so that

S = S0 exp�−��

kT�/D���� , �12�

where D���� is the relative surface state density at energy��. The equilibrium surface recombination velocity is givenby S0= �J0

2 /qNT�Jr0�e�0/kT, where NT

� =NT�0�kT /a is an equiva-lent volume concentration of defects and Jr0=qvpni�ani�p�.The thickness a of the surface only plays a role for the ho-mogeneity of NT

� and Jr0.

3. Calculation of ��

Expressing �Qm using Gauss’s theorem, �Qss by an inte-gration on surface states, and taking account of the contribu-tion to �Qsc of conduction electrons,10 the charge-neutralityequation, Eq. �1�, becomes

Cm�V −�b − �0

q� + qW0NA����b

�0��1 +

ns

NA

kT

�b� − 1�

+ qNT�0�EFh

EFe

D���d� = 0, �13�

where W0 is the equilibrium value of the depletion layerwidth, NA is the acceptor concentration, and Cm=�t /d is thecapacitance per unit area of the tunnel gap. Since qVs and theS are expressed as a function of ��, �Eqs. �8� and �12�,respectively� this energy is the relevant quantity for calculat-ing the tunnel currents and is found by numerically solvingEq. �13�.

C. Calculation of the photoassisted tunnel currents

The tunnel photocurrent density Jt is generally the sum ofthree contributions describing, respectively, tunneling ofphotoelectrons from the conduction band �Jtb�, from surface

PHOTOASSISTED TUNNELING FROM FREE-STANDING… PHYSICAL REVIEW B 82, 115331 �2010�

115331-3

Page 4: Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces

states �Jts�, and of the current from the valence band �Jtv�.Since the tunnel photocurrents are assumed negligible withrespect to the photocurrent, each of these contributions canbe calculated separately. Electron tunneling between the con-duction band and the metal occurs by conservation of theparallel electronic momentum and of the total energy.26 Inaddition to the electronic perpendicular momentum, the tun-nel probability depends on the spatial average of the tunnelbarrier, which itself depends on bias.19 As discussed in Ap-pendix B, the tunnel probability is a function of energyabove the conduction-band edge. The majority of tunnelingelectrons have a nonzero energy, written as f�b, where f� f� is a number a priori distinct from f� and defined by thelowest quantized state �Eq. �9��. For Jtb one then obtains

Jtb = Jtb0 N��S�exp�−

qV

kT� , �14�

where

=d

2d0

kT

��b�

, �15�

N��S� =A��T2

qSW0� qSn0

A��T2� f+ �1−2�f�

, �16�

and

Jtb0 = Kb�m�E�exp�−

��b�/4 − �1 − 2f��0�

kT� �17�

are, respectively, the reduced distance, the surface electronconcentration, and a factor independent of bias and light ex-citation. E=Eg− �1− f��b+qV is the energy of tunnelingelectrons with respect to the metal Fermi level. The quantity�, defined in Eq. �B4�, is the fraction of perpendicular kineticenergy to total kinetic energy. Kb, defined by Eq. �B5�, is aconstant. The other quantities are d0= /�2m and�b

�= ��m+�−EG+ �1−2f��0� /2. The exponential factor inEq. �14� is due to the bias-dependent tunnel barrier while thebias dependence of N��S� reflects the changes in the surfacerecombination velocity.

The tunnel current from surface states is obtained by in-tegration over energy � with respect to midgap between theelectron quasi-Fermi level and the metal Fermi level. Onefinds

Jts = NT�0�A exp�− s

kT�q�V − Vs� + ���

���+q�V−Vs�

��

K�Es��m�Es�D���exp�2� s

kT�d� ,

�18�

where �s�= 1

2 ��m+�+�0�, s= d2d0

kT��s

�, A=exp�−2d��s

� /d0�,and Es=�−��+qVs. Taking account of Eq. �8�, this expres-sion becomes

Jts = NT�0�A exp�−q sV

kT�

��qSn0

J0� s

��+q�V−Vs�

��

Ks�E��m�E�D���exp�2� s

kT�d� .

�19�

In this expression it is noted that the dependence on lightexcitation power is mostly contained in the term �qSn0 /J0� s.

The tunnel current from the valence band can also bemodulated under light excitation since the photovoltagemodulates the energy of the top of the valence band at thesurface. This equivalent photocurrent appears as soon asqV��b and is given by

Jtv = Jts0 exp�−

vqV

kT��qSn0

J0� v

0

qV−�b

�Gv��v��m�E�D��v�exp�−4 v�v�

kT�d�v, �20�

where v= d2d0

kT��v

�, �b

�= ��m+�+EG+�0� /2, and Jtv0

=Kv exp�−2d��v

d0�. The density of states per unit surface is

D��v�=�c�2m� /2�3/2��v, where �c is the coherence length.Kv gives a measure of the tunnel matrix element, and G��v�is a slowly varying function similar to that defined in Appen-dix B for conduction electrons. In the same way as for sur-face states, the power dependence of this current is given bythe third factor of Eq. �19� and is of the form N0

v, where vis smaller than s because of the larger value of the tunnelbarrier.

III. EXPERIMENTAL

A. Experimental system and procedure

We have used free-standing, 3-�m-thick, cantileverpatches of p+ GaAs �doping 1018 cm−3, stiffness 21 N/m�deposited on fused silica substrates with the cantilever over-hanging the substrate. These devices have been fabricatedusing an original microfluidic assembly process developedby some of the authors.16 No preliminary surface treatmentwas given to the cantilevers before the experiment. As shownin the top panel of Fig. 2, the cantilevers were excited by alaser diode at 1.59 eV, of power 5 mW, focused to a spot of20 �m diameter. The laser beam reflected by the cantileverwas also detected by a quadrant photodiode which permitsthe measurement of the force between cantilever and the sur-face and therefore to characterize the mechanical contact.Freshly made, atomically smooth, Au surfaces, fabricated us-ing an electrochemical technique described elsewhere,27

were used for the experiments. All experiments were madeunder air ambient at room temperature.

The experimental procedure, described in the lower panelof Fig. 2, is similar to that used before.12 The current wasstabilized in the dark to a value

Idark�Vset� = 10nAx exp�Idark� � �21�

for a cantilever bias Vset �set here to −1.5 V�. The reduceddark current Idark

� , imposed by the feedback control system

VU et al. PHYSICAL REVIEW B 82, 115331 �2010�

115331-4

Page 5: Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces

determines the tunnel distance if quantities such as the di-electric constant of the tunnel gap are constant. After stabi-lization, the feedback loop was opened and two rapid biasscans were performed, one in the dark and the other oneunder illumination. This procedure allows us to measure boththe dark current Idark and the tunnel photocurrent Iph definedas the additional tunnel current induced by light excitation.

In the following we show the bias dependences as a func-tion of Idark

� rather than of the tunnel distance which is notknown accurately. Figure 3 shows the dark current as a func-tion of bias while Fig. 4 shows the absolute value of theadditional photocurrent. For a positive bias, the photocurrenthas the opposite sign and is due to tunneling of holes tooccupied states of the metal. This process, which will not bediscussed here, compensates the electron photocurrent at abias of about 0.2 eV which is therefore not directly related tothe standard photovoltage. Figure 5 shows the atomic forcebetween the cantilever and the metal surface.28 Finally, thephotocurrent as a function of light excitation power, for anapplied bias of −1.5 V, is shown in Fig. 6. This figure showsa power law with an exponent which slightly increases withdistance from 0.44 �curve d� to 0.66 �curve a�.

B. Analysis

The experimental results of Figs. 3–5 allow us to distin-guish two regimes as a function of Idark

� . As seen in Fig. 4, forIset�0, the photocurrent behavior is very close to exponen-

tial, with a slope which decreases with distance betweencurves a and b and increases again between curves b and c.For larger values of Idark

� , the photocurrent increases moreslowly than exponential. The limit between the two behav-iors approximately coincides with the onset of mechanicalcontact which, as seen in Fig. 4, occurs between Idark

� =0 andIdark

� =0.5. In forward �positive� bias, it is possible to definean ideality factor n since the dark current exhibits exponen-tial behavior according to exp�qV /nkT�. Figure 3 shows thatfor Idark

� �0, in agreement with Ref. 38, the ideality factor

cantilever

light excitation

lensbeamsplitter

Quadrantphotodiode

SamplePiezo-electric

Tube #1

Piezo-electricTube #2

FIG. 2. �Color online� The top panel shows the experimentalsetup. The tipless cantilever is fixed at the end of a piezoelectrictube, and the light reflected from its end is analyzed by a quadrantphotodiode for atomic force microscopy-like investigations. Thebottom panel shows the experimental procedure: after a stabiliza-tion period in the dark, the feedback loop is opened and two rapidscans of the cantilever bias are performed, one in the dark and oneunder light excitation. The cantilever current is monitored.

-1.6 -1.2 -0.8 -0.4 0.1 0.2 0.3 0.4 0.50.01

0.1

1

10

100

1000

c

de

fg

h

b

aDar

kcu

rren

t(n

A)

Bias (qV)

FIG. 3. �Color online� Measured dark current versus bias forreduced values of the dark current imposed by the feedback loopIdark� equal to �a� −1.5, �b� −1, �c� −0.5, �d� 0, �e� 0.25, �f� 0.75, �g�

1, and �h� 1.5. �Idark� is defined by Eq. �21�.� The exponential bias

dependence of the current for a forward �positive� bias gives theideality factor, which decreases up to Idark

� =0 and stays approxi-mately constant for larger values of Idark

� . The curves were rigidlyshifted for clarity by a factor of �d� 2, �e� 4, �f� 8, �g� 16, and �h� 30.

-1.6 -1.2 -0.8 -0.4 0.0 0.4

1

10

100

g

cd

f

e

h

b

a

Tu

nn

elP

ho

tocu

rren

t(n

A)

Bias (qV)

FIG. 4. �Color online� Tunnel photocurrent versus bias, definedas the additional tunnel current under light excitation. For �a�Idark� =−1.25, �b� −1, and �c� −0.5, the dependence at reverse �nega-

tive� bias is exponential. Progressive departure from purely expo-nential behavior occurs for �d� Idark

� =0, �e� 0.25, �f� 0.75, �g� 1, and�h� 1.5.

PHOTOASSISTED TUNNELING FROM FREE-STANDING… PHYSICAL REVIEW B 82, 115331 �2010�

115331-5

Page 6: Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces

decreases with increasing Idark� . For Idark

� �0 the slope of theexponential is constant which shows that the capacitance Cmof the tunnel gap is constant. The overall variation in theideality factor is from 2.7 to 1.5.

These results show that possible inhomogeneities of thetunnel distance between the tipless cantilever and the metaldo not play a role or can be corrected. As summarized in theinset of Fig. 5, before mechanical contact the tunnel distanceand therefore the ideality factor decrease with increasingIdark

� . As seen from the exponential behavior of the tunnelphotocurrent, the results can be interpreted by a well-definedtunnel distance, corresponding to the smallest value of the

tunnel gap. It will be shown that, once mechanical contact isestablished the bias dependence of the tunnel photocurrentcan be reduced to a sum of a contact contribution, character-ized by a fixed distance, together with a homogeneous non-contact contribution. The relative importance of each contri-bution depends on the ratio of the two areas and thus on Idark

� .Support for this hypothesis is shown in Fig. 7 where forIdark

� �0, the contribution of the contact to the tunnel photo-current and the dark current are expressed by

Icontact�Idark� � = I�Idark

� � − �I�− 0.5� , �22�

where I stands for tunnel photocurrent or dark current, � isan adjustable parameter which gives a measure of the rela-tive area of the noncontact part to the contact one. The valueof −0.5 chosen for Idark

� in this correction is the highest valuegiving an exponential bias dependence of the tunnel photo-current and corresponding to a noncontact regime.

It has been possible to find values of � such that �i� inreverse �negative� bias, both the tunnel photocurrent and thedark current bias dependences are nearly independent ofIdark

� . �ii� The dark current is now exponential as a function offorward �positive� bias over as much as 4 orders of magni-tude. These values are given in Table I, and as expected,decrease upon increasing Idark

� .In contact it is interesting to note that a bistability of the

atomic force is observed �see Fig. 5�. This bistability is cor-related with a bistability of the tunnel photocurrent and canbe corrected in the same way as above using two distinctvalues of � as shown in Table I for Idark

� =1.5. For the fol-lowing analysis only the smallest value of � will be consid-ered for each value of Idark

� .The corrected results are summarized in Fig. 8 which

shows the bias dependence of the tunnel photocurrent in thecontact regime and in the noncontact regime as a function ofIdark

� . The ideality factor increases from a nearly constantvalue of 1.5 in contact to 2.7 at large distances. These resultsare free of possible contact inhomogeneities arising from thelarge contact surface area and reveal the tunnel characteris-tics of a purely two-dimensional contact considered in Sec.II.

IV. INTERPRETATION

In this section, the experimental behavior of the tunneldark current and photocurrent are compared with the predic-tions of the model for the current densities since the contactarea is not well known.29 Since the model describedin Sec. II contains a relatively large number of parameters,we have chosen reasonable values of several parametersfrom the literature and no attempts have been made to adjustthem. The values of the surface recombination velocityS0=107 cm /s �Ref. 30� and the bulk recombination time �=2.6 ns �Ref. 31� have been measured independently. Thediffusion constant D=37 cm /s is calculated from the mobil-ity value for minority carriers,32 using the Einstein relation.Using D and �, one estimates a minority carrier diffusionlength of 3.1 �m. The values of N0 and vd, calculated usingEqs. �A3� and �A4� were found equal to 2�1022 m−3 and1600 m s−1, respectively. The energy dependence D���� of

Non contact Approach Contact

-1.5 -1.0 -0.5 0.0 0.5 1.0 1.5

0

1

2

3

4

Repulsive

Attractive

Ato

mic

Forc

e( µµ µµ

N)

I*dark

Non contact Approach ContactNon contact Approach Contact

-1.5 -1.0 -0.5 0.0 0.5 1.0 1.5

0

1

2

3

4

Repulsive

Attractive

Ato

mic

Forc

e( µµ µµ

N)

I*dark

FIG. 5. �Color online� Atomic force between the cantilever andmetal, as a function of Idark

� measured in the same experiment as thetunnel currents. For negative values of Idark

� , the atomic force isapproximately constant and taken as zero. As shown in the inset,mechanical contact occurs for positive values of Idark

� situated be-tween 0 and +0.75. Here a clear mechanical bistability gives rise totwo distinct values of the atomic force.

1 10

d

c

b

a Jph

=P0.5

Tu

nnel

Ph

oto

curr

ent

(arb

.un

its)

(at

V=-

1.5

V)

Light Power (mW)

FIG. 6. �Color online� Measured photocurrent versus lightpower at an applied bias of −1.5 V for �a� Idark

� =−1.5, �b� −1, �c�−0.5, and �d� �1.5�. Also shown for reference is a power law ofexponent 0.5.

VU et al. PHYSICAL REVIEW B 82, 115331 �2010�

115331-6

Page 7: Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces

the density of surface states is approximated by a Gaussianprofile of width �, estimated to be 0.20 eV,20,33 whereas for�� larger than �, the tails of the valence and conductionbands are approximated by parabolas. The density of surfacestates NT�0� has been found to range from several1017 eV−1 m−2 up to larger than 1018 eV−1 m−2.20,33 Here wetake NT�0�=6�1018 eV−1 m−2 as implied by the slopes ofthe bias dependences at large distance and discussed in Sec.IV B below. Since Cm and depend on the width of thetunnel gap, their values are adjusted for each spectrum whilemaintaining constant values of �b

� and �s�. Using Eq. �16�,

we take the exponent of the experimental power dependenceof the photocurrent for the factor f�0.4�. For simplicity wetake f�= f , thus assuming that tunneling of photoelectronsoccurs from the first quantized state in the depletion layer.The parameter values used in the model are summarized inTable II. In the following, we outline the physical mecha-nisms underlying the tunnel photocurrent from a semicon-ductor into a metal.

A. Tunnel currents from surface states and from theconduction band

The relative values of Jts, Jtv, and Jtb depend on the tunnelmatrix elements which are unknown so that the magnitudesof these currents cannot be conclusively determined. How-ever, for the present system, the experimental evidence pre-sented here is that the tunnel photocurrent from surface statesand from the valence band are negligible with respect to thatfrom the conduction band.34,35 This is most apparent for thefollowing two reasons:

�a� The predicted excitation power dependences of Jts andJtv are very weak and cannot explain the experimental re-sults. Recalling that ��1 at large distance, it is seen fromEqs. �19� and �20� that these dependences are dominated bythat of N0

s and N0 b. The exponents s and b are on the

order of 1�10−2 and are more than one order of magnitude

smaller than the experimental values. Even larger discrepan-cies are found in contact.

�b� The bias dependences of Jts and Jtv cannot interpretthe data for all values of Idark

� . This is shown in Fig. 9 for theextreme case of curve a and curve d in Fig. 8. The biasdependence of Jtv exhibits a threshold near −0.4 V and non-exponential behavior which does not interpret the data atlarge distance. At large distance, the bias dependence of Jts isfound exponential, in agreement with the data. However, incontact, because of the nonlinear integral of Eq. �18�, Jts isalmost independent of bias and cannot interpret the experi-mental data. Even a strong modification of the tunnel param-eters cannot account for the experimental results.

B. Tunnel current from the conduction band

Comparison of the experimental results with the bias de-pendences of Jtb, calculated using Eq. �14� is shown in Fig.8. Very good agreement with the experimental results is ob-tained. The values of Cm and used in the comparison aregiven in Table II. Both of them increase with decreasingIdark

� , which reveals an increase in the tunnel distance. Figure10 shows curves a and d from Fig. 7 along with the calcu-lated bias dependences of N��S� and exp�− qV /kT� which

TABLE I. Experimentally measured tunnel currents in contactwere corrected by subtracting a fraction � of the tunnel currentcorresponding to the largest value of Idark

� out of contact. This tableshows the values of � as a function of Idark

� . Bistability of theatomic force �see Fig. 4� in contact is correlated with two distinctvalues of � shown here for Idark

� =1.5.

Idark� 1.5 1.25 1 0.75 0

� 1 0.8 0.85 1 1

0.7

-1.6 -1.2 -0.8 -0.4 0.0 0.1 0.2 0.3 0.4 0.5

0.01

0.1

1

10

100

e

d

c

ba

Darkcu

rrent

(nA

)

Bias (qV)

-1.6 -1.2 -0.8 -0.4 0.0 0.4

0.1

1

10

100

1000

e

d

c

b

a

Tu

nn

elP

ho

tocu

rren

t(n

A)

Bias (qV)

-1.6 -1.2 -0.8 -0.4 0.0 0.1 0.2 0.3 0.4 0.5

0.01

0.1

1

10

100

e

d

c

ba

Darkcu

rrent

(nA

)

Bias (qV)

-1.6 -1.2 -0.8 -0.4 0.0 0.4

0.1

1

10

100

1000

e

d

c

b

a

Tu

nn

elP

ho

tocu

rren

t(n

A)

Bias (qV)

FIG. 7. �Color online� Bias dependences of the dark and tunnel photocurrents in mechanical contact after the correction defined by Eq.�22� in order to take account of inhomogeneities of tunnel distance. The curves correspond to �a� Idark

� =0, �b� 0.75, �c� 1, �d� 1.25, and �e�1.5 and were multiplied for clarity by a factor of 2 for curve c in the dark, 4 for curve d, and 10 for curve e. The bias dependence of thecurves depends very little on Idark

� and, along with the improved exponential character at forward �positive� bias, shows that each curvecorresponds to a constant tunnel distance.

PHOTOASSISTED TUNNELING FROM FREE-STANDING… PHYSICAL REVIEW B 82, 115331 �2010�

115331-7

Page 8: Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces

appear in Eq. �14�. While the exponential factor accounts forthe bias dependence of the tunnel barrier height, N��S� ex-presses the bias-induced decrease in the surface recombina-tion velocity which, according to Eq. �9�, produces an in-crease in the concentration of tunneling electrons.

At large distance, the bias dependence of the tunnel cur-rent is due to that of the tunnel barrier height as the surfacerecombination velocity only weakly depends on bias. Indeed,Eq. �12� simplifies into

�� − �t�q�V − Vs

�� = −Cm

q2NT�0�q�V − Vs

�� . �23�

�� is smaller than the width � of the surface density ofstates so that the electron quasi-Fermi level is still pinned

near midgap. The linear bias dependence of �� induces anexponential dependence of the tunnel photocurrent, propor-tional to exp�−V /Vph�, where

kT

qVph= �t

� + . �24�

The second term of Eq. �24�, given by Eq. �16�, is propor-tional to d and expresses the bias dependence of the tunnelbarrier. The first term, which is proportional to d−1, reflectsthe dependence of the tunnel barrier on ��. The observeddecrease in the slope for Idark

� increasing between −1.25 and−1 implies that the exponential increase in the tunnel currentis determined by the bias dependence of the tunnel barrierand that �t

�� . The subsequent increase between −1 and

TABLE II. Values of parameters used for the analysis of thecurves of Fig. 7. �a� Common parameters and �b� adjustableparameters.

�a� Parameter Value

S0 /vd 62.5

Jsat �A /m2� 6�1010

NA �m−3� 1024

NT�0� �eV−1 m−2� 6�1018

NTD�0� �eV−1 m−2� 8�1017

� �eV� 0.20

N0 �m−3� 2�1022

f 0.38

�b� Iset Contact −1000 −2000 −2500

�0 /Cm �nm� 0.009 0.18 0.47 0.85

0.011 0.017 0.027 0.043

�d� �10−3 V−1� 30 �2 �2 Irrelevant

-1.6 -1.2 -0.8 -0.4 0.00.01

0.1

1

10

100

Darkcu

rrent

(nA

)

d

c

b

a

Bias (qV)-1.6 -1.2 -0.8 -0.4 0.0

0.1

1

10

100

1000

c

d

b

a

Tu

nn

elP

ho

tocu

rren

t(n

A)

Bias (qV)-1.6 -1.2 -0.8 -0.4 0.0

0.01

0.1

1

10

100

Darkcu

rrent

(nA

)

d

c

b

a

Bias (qV)-1.6 -1.2 -0.8 -0.4 0.0

0.1

1

10

100

1000

c

d

b

a

Tu

nn

elP

ho

tocu

rren

t(n

A)

Bias (qV)

FIG. 8. �Color online� Summary of the experimental bias dependences of the tunnel photocurrent �left panel� and dark current �rightpanel�. The out-of-contact dependences have been obtained for �a� Idark

� =−1.25, �b� −1, and �c� −0.5. The contribution of the contact to thebias dependence �d� was taken from the data obtained for Iset=1.5 after the correction defined by Eq. �22�. For clarity the data for the tunnelphotocurrent were multiplied by �b� 2, �c� 4, and �d� 10 while the multiplication factors for the dark current were �b� 2, �c� 3, and �d� 4. Notethat the dark current is about one order of magnitude smaller than the tunnel photocurrent. Lines show the calculated currents found usingthe parameters in Table II.

-1.6 -1.2 -0.8 -0.4 0.0

I*dark

=-1.25

Bias (qV)

Tu

nn

elP

ho

tocu

rren

t(ar

b.un

its)

Jts

Jtv

Contact

102

100

104

102

100

104

102

100

104

102

100

104

FIG. 9. �Color online� Tunnel photocurrents from surface statesand from the valence band calculated for Idark

� =−1.25 using Eqs.�19� and �20�, respectively. The other fixed parameter values thathave been used are shown in Table II. Also shown by symbols arethe corresponding experimental data.

VU et al. PHYSICAL REVIEW B 82, 115331 �2010�

115331-8

Page 9: Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces

−0.5 suggests that d is now small enough so that �t�� . The

condition �t� implies that the value of is given by the

measured exponential slope at large distance. Using the val-ues of Cm and given in Table II, one finds that qNT�0�should be of the order of several 1018 eV−1 m−2 which isindeed the case.

At small distances, �� increases because of the largervalue of the tunnel capacitance Cm. �� can become largerthan the width � of the band of surface states which inducesan unpinning of the surface Fermi level and a decrease in thesurface recombination velocity. The bias dependence of thetunnel current is now caused by the increase in the electronconcentration ns which, as seen in the top panel of Fig. 10 isas large as three orders of magnitude.

V. DISCUSSION

A. Effects of interface chemistry

The values of the parameters used in the analysis suggestthat the natural oxide layer, originally present at the surface,has been at least partially removed. For a Schottky barriercomposed of gold deposited on naturally oxidized GaAs, onefinds a value of �0 /Cm=d /�t1.5, about two orders of mag-nitude larger than the one measured here in contact.36 Asshown for InP, the oxide may have been removed by anelectrochemical reaction at cathodic potentials.37

Taking �b�4 eV in Eq. �15�, one finds that the distance

d ranges between 1.1 and 0.45 nm in the noncontact regimeand is about 0.28 nm in contact. The resulting values of thedielectric constant of the interfacial layer �t are shown in Fig.11 as a function of distance. �t is equal to ��30 in contactwhich suggests the partial formation of a molecular film ofwater �dielectric constant 80 and thickness d�0.28 nm� be-tween the semiconductor and the metal. If d�d�, one ex-

pects the effective dielectric constant to be given by�t=d���d�+ �d−d�����−1. As shown in Fig. 11 for the non-contact regime, the correspondence between the calculateddependence of �t and the data is unexpectedly good, giventhe uncertainties in most parameters used in the calculation.

B. Dark current

For a forward �positive� biases, including the contribution�d of residual processes such as image charge effects andtunneling of majority carriers, the ideality factor is givenby9,38

1

n= 1 −

�s/W + q�1 − ��NT�0�Cm + �s/W + qNT�0�

− �d � − �d, �25�

where � and 1−� are the fractions of the total number ofstates for which follow the metal statistics and thesemiconductor statistics, respectively. Comparison of themodel with the data as discussed in Sec. IV B shows that�s /W�q�1−��NT�0� and Cm�qNT�0� which leads to theapproximate expression in Eq. �25�. Figure 11 shows thedependence of n−1 on �0 /Cm, where the value near zero cor-responds to the contact situation.

Under reverse �negative� bias, �� and qVs are found bynumerically solving the current and charge conservationequations and the dark tunnel current from surface states isgiven by Eq. �18�.39 Current conservation implies that thetunnel and Schottky currents are equal. For the Schottky cur-rent, in order to take account of additional processes contrib-uting to the ideality factor, one replaces �0 by �0

� whichdepends on the barrier change ��−qVs. For a large biasrange we write to second order

�⎠⎞

�⎝⎛−

kT

qVϖexp

102

100

104

102

100

104

101

100

101

100

Tun

nelp

hot

ocu

rren

t(a

rb.u

nits

) Contact

-1.6 -1.2 -0.8 -0.4 0.0

N*(S)I*

dark=-1.25

Bias (qV)

FIG. 10. �Color online� Explanation for the distinct bias depen-dences before and after mechanical contact. This figure shows themeasured bias dependence of the photocurrent for Idark

� =−1.25�circles� and for mechanical contact �squares� as well as the calcu-lated bias dependence of the second and third terms of Eq. �14� ascalculated using the parameter values shown in Table II. Beforecontact the bias dependence of the tunnel photocurrent is deter-mined by that of the tunnel barrier. After contact the bias depen-dence of the surface recombination velocity plays a dominant role.

0.4 0.8 1.21

10

εε εε t

Distance (nm)

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

εεεε0/C

m

1/n (Experimental)ηηηη1-αααα

d

η − αη − αη − αη − αd

0.4 0.8 1.21

10

εε εε t

Distance (nm)

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

εεεε0/C

m

1/n (Experimental)ηηηη1-αααα

d

η − αη − αη − αη − αd

FIG. 11. Calculated dependence on distance of the dielectricconstant out of contact, assuming that the metal is covered by a thinlayer of thickness smaller than the tunnel distance. The data points�full squares� correspond to the values used in the analysis of theout of contact curves. The inset shows various parameters used inthe calculation as a function of �0 /Cm. Image charge effects andtunneling of majority carriers contribute to an ideality factor�1−�d�−1 and only play a significant role in contact.

PHOTOASSISTED TUNNELING FROM FREE-STANDING… PHYSICAL REVIEW B 82, 115331 �2010�

115331-9

Page 10: Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces

�0� = �0 + �d�qVs − ��� + �d��qVs − ���2. �26�

In the charge-neutrality equation, in addition to ns=0, theterm �Qss must take account of the two types of surfacestates used in forward �positive� bias. The dark current andideality factor depend on the following additional param-eters: �i� �, �ii� �d and �d�, and �iii� the tunnel matrix elementKs defined in Eq. �18�. Since this equation uses the productKsNT�0�, the quantity NT�0� will be replaced by an effectivedensity of states NT

d�0� taken here as 8�1017 eV−1 m−2.The dependences of the dark current under reverse bias

were calculated using the same parameter values as for thephotocurrent as well as imposing �−�d=n−1 from Eq. �24�.The dependences of � and �d on �0 /Cm are shown in Fig. 10.As shown in Table II, �d� is only significant in contact andhas very small values on the order of 10−3 V−1. The biasdependences of the dark current under reverse bias, shown inFig. 8, account very well for the experimental results. Thecalculations also suggest that, as expected,40 the quantity �decreases with increasing distance from a value of about 0.84while the residual ideality factor �1−�d�−1 decreases from1.20 to 1.04.

C. Validity of the approximations made

It has been assumed that the electrons in the conduction-band tunnel from the first quantized level �f f��. The powerdependence of the tunnel photocurrent gives f 0.4. The biasdependence of f� was calculated using Eq. �9�, neglecting themodification of the surface electric field due to the photo-electrons in the depletion layer: f� is approximately constantand varies from 0.38 �a value close to f� to about 0.25 as afunction of bias. In view of the numerous quantities whichplay a role in defining the value of f it is concluded thattaking f f� is a valid approximation.

At small distances, image charge effects might furthermodify the bias dependence of the tunnel photocurrent.19

However, the characteristic energy for evaluating the magni-tude of these effects �=q ln�2� / �8�td�, on the order of0.4 eV for �t=�0 and d=1 nm, is smaller by one order ofmagnitude than the effective tunnel barrier height �b

�. Asseen in Ref. 19, the tunnel barrier decreases with bias so thatimage charge effects should induce a superexponential in-crease in the tunnel photocurrent. This is at odds with theexperimental results obtained at small distance.

In order to obtain analytical expressions of the tunnel cur-rent, this current has been neglected with respect to the pho-tocurrent and Schottky currents. This assumption is certainlyvalid at large distance, in which case the tunnel photocurrentis small. In contact, the photocurrent JpqN0S /vd decreasesbecause of the reduced surface recombination velocity andcould become a lower limit value for the tunnel photocurrent�Jt=Jp�. However, the latter hypothesis can also be excludedbecause, in contradiction with the results of Fig. 6, the re-sulting power dependence of the tunnel current should bequite different from that obtained at large distance.

VI. CONCLUSION

We have developed a general model for describing thebias and distance dependence of the tunnel photocurrent

from a thin free-standing GaAs film photoexcited from therear surface and a metallic surface. Based on current andcharge conservation equations, this model predicts that thetunneling current can depend on bias via the bias dependenceof the tunneling barrier and also because application of biaschanges the position of the electronic quasi Fermi level at thesurface and therefore the effective density of states for sur-face recombination. Both the tunnel currents from the con-duction band and from surface states have been calculated.

This model was compared with experimental data of tun-neling injection into gold surfaces, for which the density ofempty states depends only weakly on energy. All results,including tunnel photocurrent, ideality factor, and dark cur-rent under reverse bias, are satisfactorily interpreted by iden-tical values of the parameters, close to the values found inthe literature. The values obtained for the width and dielec-tric constant of the tunnel gap are also reasonable. The modeland experimental results indicate that: �a� the dominant partof the tunnel photocurrent comes from the conduction band.�b� At large distance, the bias dependence of the tunnel cur-rent is interpreted as a bias dependence of the tunneling gapwhile, at smaller distance, the bias dependence of the surfacerecombination velocity plays a dominant role. The presentmodel can be used as a basis for the interpretation of futurespin-dependent tunnel photocurrent data.

ACKNOWLEDGMENTS

The authors acknowledge X. Wallart for the epitaxialgrowth. This work was partially funded by the ANRSPINJECT-06-BLAN-0253.

APPENDIX A: EXPRESSIONS OF vd AND N0

The charge diffusion equation is of the form

D�2n

�z2 −n

�+ g� exp�− �z� = 0, �A1�

where g is the density of impinging photons per unit time, �is the light absorption coefficient, � is the bulk photoelectronlifetime, and D is the diffusion constant. For a planar sampleof thickness �, the general solution of Eq. �A1� is

n+ + n− = Ae−z/L + Bez/L +g��

1 − ��L�2e−�z, �A2�

where L=�D� is the electron diffusion length. Using� �n�z �0=S�n�0� and D� �n

�z ��−W=−Sn0 as boundary conditions,one finds that the electronic concentration n0 at z=�−W andthe photocurrent are given by Eqs. �5� and �6�, respectively,where

N0 =g��

��L�2 − 1

��L − � + �S�L/D��� − ��L��S�L/D�Ch��/L� + Sh��/L�

, �A3�

vd =D

L

�S�L/D�Ch�d/L� + Sh�d/L�Ch�d/L� + �S�L/D�Sh�d/L�

, �A4�

where the quantities � and � are given by

VU et al. PHYSICAL REVIEW B 82, 115331 �2010�

115331-10

Page 11: Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces

� = 1 − e−��Ch��/L� � = e−��Sh��/L� . �A5�

For an unpassivated rear surface, one has S�L /D�Th�� /L�and S�L /D� �Th�� /L��−1 and

vd = �D/L��Th��/L��−1. �A6�

Further assuming that �L�1 and neglecting for a large valueof �� the light absorption at the front surface, one has�1 and �=0 and

N0 g�

LSh��/L�. �A7�

APPENDIX B: TUNNEL CURRENT FROM THECONDUCTION BAND

We first write, for a given energy �c above the bottom ofthe conduction band, the conservation of the perpendicularelectronic momenta k�, in the conduction band i�� in thetunnel gap and k�� in the metal � is related to the electron

mass m by 2�2 /2m=�̄−�c� where �c�=2k�2 /2m is the

fraction of the energy �c above the bottom of the conductionband corresponding to a kinetic energy perpendicular to the

surface. The spatially averaged value of the tunnel barrier �̄for electrons at the bottom of the conduction band dependson bias.19 Neglecting image charge effects, it is given by

�̄ = ��m + � − Eb + qV�/2, �B1�

where �m and � are, respectively, the metal work functionand the semiconductor affinity and Eb is the energy of thebottom of the conduction band at the surface. The momen-tum k�� is obtained by expressing conservation of energy andof parallel momentum. Assuming that exp�−2�d��1, onefinds that the tunnel probability is proportional toG��c�exp�−2�d�, where

G��c� =k

k�

k�2

�k� + k�� �2k��2 + �� + k�k�� /��2 . �B2�

The tunnel current also depends on the productK��m�E�ns��c�W��c�, where K� is a constant and �m�E� is themetallic density of states at the corresponding energy E.Here ns��c� and W��c� are the volume density of the electronconcentration and the width of the depletion zone at energy�c. Thus, ns��c�W��c� is a concentration per unit area. One

has finally, to first order in �c /�̄,

Jtb = K�ns exp�− 2d��̄/d0�0

�b

�m�E�W��c�

�G��c����c�exp� �c�d

d0��̄

−�c

kT�d�c, �B3�

where d0= /�2m. For simplicity we only retain here theelectrons which have the largest contribution to the integralof Eq. �B3�. Because W��c�G��c� increases with �c and be-cause of quantization of electronic states near the surface, theenergy of these electrons is nonzero and will be written asf�b where f is quite generally larger than f� defined in Eq.�9�. Using Eqs. �1� and �8�, one finds exp�−

�c

kT �exp�−f�b

kT �= �

qSn0

A��T2 � f. The first exponential factor in the integral of Eq.

�B3� is written as��cd

d0��̄

, where

� = �c�/�c. �B4�

Since the barrier value will be found weakly dependent onboth bias and light excitation power, the productW��c�G��c����c� will be taken as constant and incorporatedinto the multiplicative constant, thus writing

Kb = K�W��c�G��c����c� . �B5�

Equation �14� is finally obtained by developing ��̄ tofirst order in qV.

*Corresponding author; [email protected] I. Žutić, J. Fabian, and S. Das Sarma, Rev. Mod. Phys. 76, 323

�2004�.2 D. T. Pierce, Phys. Scr. 38, 291 �1988�.3 M. Bode, Rep. Prog. Phys. 66, 523 �2003�.4 M. W. J. Prins, R. Jansen, R. H. M. Groeneveld, A. P. van

Gelder, and H. van Kempen, Phys. Rev. B 53, 8090 �1996�.5 Y. Suzuki, W. Nabhan, and K. Tanaka, Appl. Phys. Lett. 71,

3153 �1997�.6 R. Jansen, R. Schad and H. Van Kempen, J. Magn. Magn. Mater.

198, 668 �1999�.7 D. Paget, J. Peretti, A. C. H. Rowe, G. Lampel, B. Gérard, and S.

Bansropun, French Patent No. 05 05394 �2005�.8 D. Vu, R. Ramdani, S. Bansropun, B. Gérard, E. Gil, Y. André,

A. C. H. Rowe, and D. Paget, J. Appl. Phys. 107, 093712�2010�.

9 E. H. Rhoderick, Metal-Semiconductor Contacts �Clarendon,

Oxford, 1978�.10 L. Kronik and Y. Shapira, Surf. Sci. Rep. 37, 1 �1999�.11 S. Grafström, J. Appl. Phys. 91, 1717 �2002�.12 A. C. H. Rowe and D. Paget, Phys. Rev. B 75, 115311 �2007�.13 R. Jansen, M. W. J. Prins, and H. van Kempen, Phys. Rev. B 57,

4033 �1998�.14 M. W. J. Prins, R. Jansen, and H. van Kempen, Phys. Rev. B 53,

8105 �1996�.15 J. Reichman, Appl. Phys. Lett. 36, 574 �1980�.16 S. Arscott, E. Peytavit, D. Vu, A. C. H. Rowe, and D. Paget, J.

Micromech. Microeng. 20, 025023 �2010�.17 W. G. Gärtner, Phys. Rev. 116, 84 �1959�.18 D. Aspnes, Surf. Sci. 132, 406 �1983�.19 J. G. Simmons, J. Appl. Phys. 34, 1793 �1963�.20 E. W. Kreutz, Phys. Status Solidi A 56, 687 �1979�.21 D. A. Papaconstantopoulos, Handbook of the Band Structure of

Elemental Solids �Plenum Press, New York, 1986�.

PHOTOASSISTED TUNNELING FROM FREE-STANDING… PHYSICAL REVIEW B 82, 115331 �2010�

115331-11

Page 12: Photoassisted tunneling from free-standing GaAs thin films into metallic surfaces

22 J. He, M. Chan, and Y. Wang, IEEE Trans. Electron Devices 53,2082 �2006�.

23 C. G. B. Garrett and W. H. Brattain, Phys. Rev. 99, 376 �1955�.24 R. A. Smith, Semiconductors, 2nd ed. �Cambridge University

Press, Cambridge, 1978�.25 C. H. Henry, R. A. Logan, and F. R. Merritt, J. Appl. Phys. 49,

3530 �1978�.26 L. Landau and E. Lifchitz, Quantum Mechanics �Mir, Moscow,

1966�.27 P. Prod’homme, F. Maroun, R. Cortes, and P. Allongue, Appl.

Phys. Lett. 93, 171901 �2008�.28 This force is found constant during the bias scan, which implies

that the stiffness of the cantilever is large enough so that thesystem geometry is not modified by electrostatic forces duringthe scan.

29 The values of the tunnel photocurrents under contact, in the sameconditions of light excitation for which the contact area is rela-tively large are quite similar to those found in the same setup forsilicon tips �Ref. 12�. This implies that the feedback loop com-pensates for the change in tunnel area and that in identical lightexcitation conditions the tunnel area does not play a crucial role.

30 H. Ito and T. Ishibashi, Jpn. J. Appl. Phys. 33, 88 �1994�.31 R. J. Nelson and R. G. Sobers, J. Appl. Phys. 49, 6103 �1978�.32 J. R. Lowney and H. S. Bennett, J. Appl. Phys. 69, 7102 �1991�.33 Ş. Karataş and Ş. Altındal, Mater. Sci. Eng., B 122, 133 �2005�.34 Jts and Jtv are small with respect to Jtb because their respective

tunnel barriers are larger than for conduction electrons, or pos-sibly because of their relatively small tunnel matrix elementsand coherence length �c for Jtv in Eq. �20�. In agreement withthese conclusions, the tunnel current from the conduction bandof n-type GaAs is known to be larger than that from the valenceband �R. M. Feenstra, Phys. Rev. B 50, 4561 �1994��. Moreover,no tunnel current from defects is found on oxygen-covered

GaAs. �R. M. Feenstra and J. A. Stroscio, J. Vac. Sci. Technol. B5, 923 �1987��.

35 The difference between the present results and those of Refs. 13and 14 for injection from GaAs tips, which is not clear at thepresent time can have three explanations: �i� if the photoelectronpresence probability at the tip apex is smaller than for a planarsurface, the matrix element Kb will be reduced. �ii� Distinct dis-tribution and concentration of surface states. �iii� The thicknessof the interfacial layer may be smaller at the apex, which shouldfavor the tunnel photocurrent from surface states with respect tothat from the conduction band.

36 N. L. Dmitruk, O. Yu Borkovskaya, and O. V. Fursenko, Vacuum50, 439 �1998�.

37 N. C. Quach, N. Simon, I. Gérard, P. Tran Van, and A. Etche-berry, J. Electrochem. Soc. 151, C318 �2004�.

38 H. C. Card and E. H. Rhoderick, J. Phys. D 4, 1589 �1971�.39 This approach neglects the tunnel current from the valence band

Jtv. At short distance, the semiconductor band structure followsthe motion of the metal Fermi level �Vs=V� so that Jtv=0 sincevalence-band states lie below the metal Fermi level. At largedistance, one would expect in the same way as in Fig. 8, appear-ance of Jtv at a bias on the order of −0.6 V. Such effect is notobserved experimentally, probably because of the smallness ofthe tunnel matrix element or of the coherence length in the va-lence band which as seen in Eq. �18� appears in the two-dimensional density of states in the valence band.

40 The presence of two distinct types of surface states and the biasdependence of the barrier �0

� have not been taken into accountunder light excitation. This is reasonable since �i� photoelectroncapture processes increase the kinetics of establishment of equi-librium with the semiconductor, �ii� because of the photovoltage,the correction term, proportional to ��−qVs is smaller underlight excitation than in the dark.

VU et al. PHYSICAL REVIEW B 82, 115331 �2010�

115331-12