Access to this full-text is provided by Springer Nature.
Content available from Scientific Reports
This content is subject to copyright. Terms and conditions apply.
1
SCientifiC REPORTS | 7: 12790 | DOI:10.1038/s41598-017-13007-w
www.nature.com/scientificreports
Imaging Atomic Scale Dynamics
on III–V Nanowire Surfaces During
Electrical Operation
J. L. Webb1, J. Knutsson1, M. Hjort
1, S. R. McKibbin1, S. Lehmann2, C. Thelander2,
K. A. Dick2, R. Timm
1 & A. Mikkelsen1
As semiconductor electronics keep shrinking, functionality depends on individual atomic scale surface
and interface features that may change as voltages are applied. In this work we demonstrate a novel
device platform that allows scanning tunneling microscopy (STM) imaging with atomic scale resolution
across a device simultaneously with full electrical operation. The platform presents a signicant step
forward as it allows STM to be performed everywhere on the device surface and high temperature
processing in reactive gases of the complete device. We demonstrate the new method through proof of
principle measurements on both InAs and GaAs nanowire devices with variable biases up to 4 V. On InAs
nanowires we observe a surprising removal of atomic defects and smoothing of the surface morphology
under applied bias, in contrast to the expected increase in defects and electromigration-related failure.
As we use only standard fabrication and scanning instrumentation our concept is widely applicable
and opens up the possibility of fundamental investigations of device surface reliability as well as new
electronic functionality based on restructuring during operation.
Signicant changes in performance during operation have been observed in electronic devices for many years,
although the precise microscopic structural origin oen remains unclear1. As component dimensions progress to
the atomic scale, the geometry and electrical response of single atom defects will signicantly inuence perfor-
mance2,3. Especially important are surface properties as even small changes of structural features (during electri-
cal operation) can alter device function4–6 due to the large surface to bulk ratio of nanoscale structures, novel 2D
materials7,8 or single molecular9 based devices.
Techniques that can relate surface geometry and electronic structure to electrical performance are thus highly
important, of which scanning tunnelling microscopy (STM) uniquely allows imaging of both surface geometry
and electronic structure down to individual atoms and as such already play a prominent role in the strive towards
atomic scale semiconductor electronics2,3,9–11. In particular, the surfaces of III–V semiconductors are a proven
base for STM studies of fundamental atomic scale semiconductor processes2,12–18. For larger (micrometre sized)
surfaces considerable structural rearrangements have been observed by STM under applied bias19–21. However,
the high resolution capabilities of STM have not been extended to the case of electrically active nanostructured
semiconductor surfaces.
In realising such studies, III–V nanowires have a number of important features. ey can be tailored into a
wide variety of atomically precise axial and radial heterostructures with well dened composition, crystal struc-
ture and morphology22,23. is makes them highly useful both as components in future (opto)electronics24–26 as
well as versatile systems for fundamental studies of semiconductor nanoscale conned surfaces. e 1D geometry
allows local control of the voltage and thermal proles along the wire, which have proven dicult for larger thin
lms19. Importantly, it has been demonstrated that STM/scanning tunnelling spectroscopy(STS) studies to the
atomic scale are possible on a wide variety of nanowires4,27,28 giving both structural arrangements and electronic
properties such as bandgap and band alignment.
e most common device for studies of nanowire device functionality has a lateral design whereby the wires
are deposited on insulating SiO2 and subsequently top contacted by electron beam lithographically (EBL) and
thermal deposition of metal29–31. is structure can be passivated, oxidised and top gated through lithographic
patterning and can be probed at lower resolution by alternative scanning probe techniques32,33. It has been used
1Division of Synchrotron Radiation Research, Lund University, Lund, Sweden. 2Division of Solid State Physics, Lund
University, Lund, Sweden. Correspondence and requests for materials should be addressed to J.L.W. (email: james.
webb@sljus.lu.se)
Received: 4 May 2017
Accepted: 12 September 2017
Published: xx xx xxxx
OPEN
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
2
SCientifiC REPORTS | 7: 12790 | DOI:10.1038/s41598-017-13007-w
in our previous work utilising combined AFM/STM scanning to perform measurements on the device surface31
However, in trying to extend directly on this work to obtain true atomic resolution on controlled nanowire surfaces
we have identied several serious shortcomings for atomic scale STM. Firstly, on this standard device only the metal
contacts and the nanowire itself are conducting, limiting STM scanning to these areas. e tip can oen move onto
the insulating SiO2 where it is rendered permanently useless by surface crash. Secondly, the height of the contacts
precludes scanning near the important nanowire-contact region. irdly and most seriously, the strong mechanical
grip of overlaid contact electrodes can cause the nanowire to fracture during heating and cooling procedures above
300 C34. is has been observed with the device electrically isolated, ruling out electromigration-related failure.
A prerequisite for any successful STM study to the atomic scale is the presence of a well dened crystal-
line surface with a limited (controllable) number of structural defects. To achieve this surface treatments above
400–500 C in reactive gases such as atomic hydrogen4,35 must be possible on the device platform. In contrast the
common nanowire device has an oxidised surface due to transport at ambient air conditions which in reality
precludes atomic resolved studies using STM (or AFM). Previous studies have thus been limited to signicantly
lower imaging resolution. Importantly such uncontrolled surface oxide conditions also makes it impossible to
determine the role of individual specic types of defects. Having the oxide free crystalline surface as a starting
point allows for the introduction and study of specic defects (through additional overgrowth or gas dosing for
example) while applying bias and studying surface properties.
In this work we demonstrate via a series of proof of principle measurements on III–V nanowire devices that we
can achieve atomic scale resolution imaging of InAs and GaAs surfaces by ultra high vacuum (UHV) STM with
simultaneous electrical operation of the nanowire device. We demonstrate four major capabilities of our method:
atomic resolution during operation, atomic level surface control, nanoscale imaging in combination with large volt-
age variations and nally high resolution electrical potential mapping. We show that atomic defects on the InAs
surfaces move, reorder and vanish under applied bias and with larger nanoscale changes in surface structure while
local variation in surface potential are extracted from topographic images across a crystal stacking fault.
Results
The new device platform. We have developed a new device design suitable for STM imaging as shown
in Fig.1, consisting of a semiconductor nanowire suspended on top of two conducting electrodes that can be
independently biased and which are electrically isolated from the substrate. Unlike the >100 nm SiOx thermally
grown insulating oxide, the Si substrate has only a thin native oxide layer suciently conductive to allow STM
scanning on it. A key point is the design of the layers of the contacts as shown in Fig.1b), where the SiOx insulat-
ing layer isolating the metal electrodes from the conductive substrate is only found underneath the contact and
not visible to the STM tip. is permits scanning anywhere on the sample surface whilst electrically insulating
the nanowires from ground and allowing a bias to be applied only across them. Placing the nanowires on top of
the electrodes also reduces strain in the nanowire during heating, preventing fracture of the wires due to surface
cleaning. To prevent problems with alloying between contacts and the nanowire while retaining an Ohmic, low
resistance junction, a thin Au capping layer (<5 nm) was used, as opposed to thicker Au contacts used in more
standard designs36. e detailed fabrication and experimental layout is described in Methods. It retains the stand-
ard processing techniques and materials (Si, Ti, Au) to provide a realistic alternative that can be easily fabricated
and keep many of the properties of the conventional lateral design.
A top-down SEM image of the new device structure aer complete processing can be seen in Fig.1c) which
was transported in air to a UHV Omicron STM system. Oxide-free, crystalline nanowire surfaces were obtained
by atomic hydrogen cleaning at 400–500 C (see Methods) inside an adjacent preparatory chamber under UHV,
with immediate transfer into the STM chamber. Large scale STM imaging as seen in the inset of Fig.1c) was used
to locate a given device nanowire (the full location procedure is outlined in ‘Methods’). We could scan across the
nanowire despite the 90–100 nm height dierence between the Si substrate and the top of the nanowire. All areas
of the device could be scanned without degradation of the STM tip. Figure1d shows an STM image of part of a
nanowire contacted by the Ti/Au electrode and suspended above the Si substrate (black), used for positioning
of the STM at any specic distance from the contacts or any other relevant (hetero)structure along a nanowire.
Atomically resolved imaging during device operation. As a proof-of-principle of our method for
samples with many dierent structural features, we imaged a GaAs nanowire surface that had a mixture of sur-
face steps, free crystalline surface and defects (largely remaining oxide that can be removed by further cleaning
cycles27). In Fig.2a) we show a STM image of this surface while there is simultaneously a Vb = −3 V bias along the
nanowire. is demonstrates that true atomically resolved imaging (even in the presence of steps and defects) is
possible with a large voltage drops over the nanowire device. To show that high quality measurements are possi-
ble across dierent materials and crystal structures we show an atomically resolved image from a wurtzite InAs
nanowire in the device conguration in Fig.2b. is is not a trivial observation, as we will now demonstrate how
applied bias can seriously aect the semiconductor surfaces.
We applied the atomically resolved imaging to study surface defect changes and mobility during operation
across a uniform InAs nanowire. For these measurements we chose to apply the bias and then image at Vb = 0 in
order to fully and unambiguously distinguish the eect of applied bias from the eect of tip scanning on the sur-
face, which could otherwise potentially generate and rearrange defects independently to the applied bias.
Figure3a) shows the rst image in one session of such STM imaging with a high resolution image taken on the
{1010}
facet on an InAs nanowire. We observe an unreconstructed at surface with the same quality as in previous
STM studies of grounded InAs nanowires on an InAs substrate28. As labelled on the gure, we observe randomly
distributed point defects, some resembling As vacancies as previously seen on InAs(110)37 as well as crystal stack-
ing faults.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
3
SCientifiC REPORTS | 7: 12790 | DOI:10.1038/s41598-017-13007-w
Studying defect dynamics requires imaging stability which we tested by scanning the same area of the surface
repeatedly for more than 1 hour with no bias along the nanowire. Figure3b) shows the same position aer this
period of continuous scanning at Vb = 0 V. At Vb = 0 V the small point defect vacancies and the stacking fault do
not alter. Only a few mobile clusters of atoms on the surface moved. Aer checking that the tip did not signicantly
inuence surface structure, a bias of Vb = −2 V was applied along the wire for 10 min. e image in Fig.3c) taken at
Vb = 0 V aer this period shows the dramatic eect of the applied bias. e defect density was signicantly reduced
with many point defects now absent (example indicated by red arrow). e stacking fault remained essentially unal-
tered, but two vacancies in it were eliminated along with several larger adsorbate features. In all three images atomic
resolution is retained, but the tip changes somewhat (as is oen the case for extended imaging periods on volatile
sample surfaces). is could potentially also render some types of defects dicult to observe and present a general
problem. However, both before and aer device bias have been applied we can observe some of the same vacancies
and adsorbates (example indicated by the blue arrow) while other defects have gone. If the disappearance of some
defects were due to a tip change rendering them invisible one would expect that all of these defects would no longer
be visible. Since this is not the case we can conclude that the observation of defect changes is possible on III–V sur-
faces even under structural changes. We address this issue of tip change further through Supplementary Information
Figs2, 3 and 4 where we indicate defects present in scan images before and aer bias, an area of surface reconstruc-
tion and show that these can still be observed (and observed to have moved or disappeared) even aer a tip change.
In addition we show the alteration and reconstruction in the direction of current ow of a large scale defect in SI
Fig.5 at atomic resolution with no change in the tip before and aer imaging.
We have performed similar experiments to the GaAs(110) surface as seen in Fig.2. Unlike the observations
on InAs we found limited change in defect and surface structure under the same level of applied bias. We attrib-
ute this to the higher resistivity and lower current densities in GaAs alongside the generally stronger bonding of
Figure 1. Nanowire device for atomically resolved STM. (a) Schematic of the STM device experiment. e
nanowire (in green) has been cleaned and atomic resolution is possible as illustrated in the zoom-in. Vb is
dened as the bias applied across the nanowire. Vtip the bias on the STM tip with respect to the ground. e
nanowire contacts can be independently grounded which prevents electrical shorting and denes a reference
ground level (and reference height for the STM tip). (b) Prole of the nanowire-contact area in (a). e
nanowire crosses a conducting Si trench contacted by Ti/Au electrodes and separated from the conducting Si
substrate by thermally deposited SiO (or sputter grown SiO2). e Si substrate can either be grounded or used
as a back gate. e nanowire can be held in position by an optional thin Ti ‘holding’ layer deposited on top of it
parallel to the trench, but signicantly away from the trench edge. (c) A top down SEM image of the completed
device, showing the nanowire above a 1 μm wide Si trench and contact electrodes (labelled). Inset: Fast scan
overview STM image of Ti/Au contacts and nanowire, the image is 3 × 2 μm. (d) Higher resolution STM image
of a nanowire device, showing the Ti/Au contact, nanowire above the Si substrate. All regions of the device can
be scanned by the STM tip at tip bias of −1.5 V in constant current mode at setpoint 60 pA.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
4
SCientifiC REPORTS | 7: 12790 | DOI:10.1038/s41598-017-13007-w
the GaAs lattice38. is self-cleaning eect upon supplying a voltage through the InAs nanowire is unexpected
as prolonged STM scanning and exposure to vacuum leads to the formation of more defects on some III–V sur-
faces17. As signicant Joule heating eects can be excluded39 and any conceivable temperature rise would lead to
formation of more vacancies, we must conclude that it is the electric current or voltage gradient through the wire
surface that directly leads to the removal of defects. is will be discussed in more detail below.
Larger scale reconstruction observed after applying bias. In order to further explore the opportu-
nities for studies of structural changes and since surface steps can signicantly impact nanowire device function
(through charging and acting as recombination centres) we investigated larger nanoscale changes in nanowire
surface morphology. We therefore investigated larger nanoscale changes in nanowire surface structure. Figure4
shows a larger scale section of the InAs nanowire surface, imaged aer successive application of Vb ranging from
0 to −3 V for 5–10 min at each bias. We initially observed a surface with a number of larger islands, separated by
defects ranging in size from small point defects to the large features separating surface islands of atoms.
Recording many repeated images with no bias applied along the nanowire and we observed some mobile
surface clusters that could be picked up and relocated by the tip. is was also the case for biases Vb < −2 V in
Fig.4a,b where only a few clusters have moved with no evidence of signicant changes induced by nanowire bias.
Aer Vb < −2 V, signicant structural change in the surface was observed as seen in Fig.4c,d. Most prominently
the island morphology completely changes and the surface becomes smoother, the larger features between islands
Figure 2. STM imaging with atomic resolution on both InAs and GaAs nanowire devices. (a) 20 × 20 nm2 STM
image of the (110) surface of a GaAs nanowire with zincblende structure. e gure inset more clearly shows
the atomic resolution from the top layer As atoms on the crystalline surface with the individual As atoms of the
surface (bright) and a defect consisting of several As vacancies (dark). e image is recorded with Vb = −3 V
applied across the nanowire. (b) 20 × 20 nm STM image of a InAs
{1120}
facet showing the atomic resolved zig-
zag pattern observed on the surface of a wurtzite InAs nanowire device. In this specic session the surface was
imaged for about 3 hours under varying bias conditions. Here Vtip = −2 V in constant current mode at setpoint
90 pA.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
5
SCientifiC REPORTS | 7: 12790 | DOI:10.1038/s41598-017-13007-w
changing shape and lling. Eventually a more defect free and stable surface appears as seen in Fig.4d. At this
point, few changes occur even when applying −3 V along the nanowire for 5 min. We speculate that the changes
observed are from either direct migration of In/As atoms to ll the defects and regions between islands due to
the strong surface electric eld under applied bias, or indirectly due to scattering from current ow through the
nanowire.
Electromigration of In atoms on the surface of semiconductors is a phenomenon that has been observed pre-
viously40,41. However, this should be a process that by denition produces more (vacancy) defects in a localised
area and severely compromises the crystal structure of the nanowire ultimately ending in its structural failure42.
In the work by Chao Zhang et al. using TEM for example43 this occurs close to the electrical contact, at a similar
position to our STM scan location. It is thus surprising to see the opposite eect occurring. We account for this
via two possible scenarios: rstly, that defect generation (and failure) occurred elsewhere (such as the opposite
contact), liberating atoms to move along the surface and ll defects in the scan area or that secondly we were
below the critical threshold for this to occur, with the surface rearrangement occurring due to the presence of
the strong surface electric eld. To exclude in the former case scans were performed moving along the full length
of the nanowire observing no such defect accumulation and failure region. Our previous measurements31 sug-
gest the clean, oxide free surface of InAs to have higher electrical conductance than the bulk of the nanowire - a
change to a defect free, smooth surface should thus produce an increase in conductance.
Some slight image dri was observed, potentially indicative of heating of the surface. However this dri disap-
peared rapidly (within a single 1–2 min image). Typically for samples heated above room temperature we would
observe several hundred nm of dri over such a period. e current (maximum < 1–4 μA) was signicantly
below that required to heat the nanowire signicantly and any heat dissipation from the nanowire should have
been rapid due to the good thermal contact with the large metallic electrodes on either end of the nanowire.
Recent measurements of the heating of InAs nanowires due to a driving voltage by Menges et al.39 have shown
that heating eects are very small. At the maximum voltages used in the present study a temperature rise of <60 C
would be expected. is temperature is too low to explain the observed changes in the surface as Joule heating, in
particular since heating at these temperatures of similar surfaces leads to vacancy formation and not removal17.
Generally vacancy formation has been observed even at room temperature for some III–Vs. However, at the rel-
evant temperatures of this study we only observe As vacancy formation not removal. is conclusion is reached
both based on theoretical studies that show how defects are more stable at the surface than in the bulk44 and from
many experimental observations of vacancy formation of III–V surfaces13 and also on nanowires35. Works by
others also generally nd that the STM tip might in some cases generate more vacancies, but not fewer.
An important benchmark for the usefulness of our method towards dynamical studies is that imaging could
be resumed and nanowire structural changes observed rapidly aer signicant changes in the voltage applied
across the nanowire, for example within only a few minutes aer up to −3 V had been applied to the wire for
10 min and then switched o abruptly. e ability to image immediately aer large changes of device voltage
contrasts previous studies of electromigration using STM19. In this work signicant time was needed (>30 min)
before the surface became stable enough to perform imaging on due to heating and charging eects. at we are
able to measure again so quickly further corroborates the excellent heat and charge dissipation in our setup.
Figure 3. Demonstration of the removal of surface defects by bias application. (a–c) STM images on an InAs
nanowire approximately 200 nm from the biased contact, showing the
{1010}
facet with a stacking fault (SF)and
numerous point defects and surface features. We probe lled states, with As surface atoms appearing as bright
protrusions. All images were taken at Vb = 0. On (a), label 1) is a mobile cluster of surface In atoms, 2) an As
vacancy defect. (b) Imaging the same position at Vb = 0 V aer 6 scans over one hour at zero applied bias across
the nanowire with little tip-induced motion of vacancies. (c) Imaging at Vb = 0 V aer Vb = −2 V had been
applied for 10 min, showing substantial motion and elimination of surface point defects. e nanowire growth
direction (y, long axis of the nanowire) across which the potential is applied is indicated by an arrow. In all three
images both vacancies and adsorbates can be clearly identied (example indicated by blue arrow) even though
tip images conditions changed slightly. However, in the third image many defects present previously were no
longer visible (example indicated by red arrow). Here Vtip = −1.5 V in constant current mode at setpoint 80 pA.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
6
SCientifiC REPORTS | 7: 12790 | DOI:10.1038/s41598-017-13007-w
Probing the eect of surface defects on nanowire conductance. An interesting feature of our setup
is the ability to probe the surface potential at any position on the nanowire surface. As we change the applied bias
Vb along the nanowire, the variation in surface potential VL alters tunnel current to the tip that in turn will change
the z-position Δd due to the response of the STM feedback loop in constant current mode. Using established
models for this z-dependence45 we can estimate how changes in apparent height variation relate to changes in VL.
is allows direct comparison between the surface electrical conduction at a given point along the nanowire and
the surface structure. e concept is based on our ability to vary Vb during continuous STM imaging of the same
area. Assuming that changes in VL depend on the local conductance this method further allows an estimate of
the eect on conductance of a specic local nanoscale surface feature, which is not possible by standard electrical
transport measurements. A detailed explanation of this method is given in Supplementary Information.
In order to provide a proof of principle demonstration of this method we study a region with a stacking fault in
an InAs nanowire, imaged in Fig.5a (top le) at Vb = 0 V. is is the same type of stacking fault as seen in Fig.3a
Figure 4. STM imaging of large scale re-arrangements due to applied bias. (a–d) Shows 40 × 40 nm2 STM
image scans taken on a wurtzite crystal structure InAs nanowire aer successive application of bias ranging
from Vb = 0 to −3 V. (a) Taken before application of any bias, (b) aer application of Vb = −0.5 V, (c) aer
application of −2 V and d) aer −3 V. Images were recorded with Vb = 0 V. e faint equally spaced lines seen
across all images are the atomic rows of the
{1010}
surface facet, indicating atomic resolution despite the surface
being changed signicantly due to the applied bias along the nanowire. An approximate common position on
each image is indicated by a red dot and the nanowire growth direction (long axis of the nanowire) across which
the potential is applied is indicated by arrow y. Here Vtip = −1.5 V in constant current mode at setpoint 90 pA.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
7
SCientifiC REPORTS | 7: 12790 | DOI:10.1038/s41598-017-13007-w
with atomic resolution. ese stacking faults are important common atomic scale structural features in nanow-
ires46 that can inuence conductance and are clearly visible and in the present case do not change with applied
bias along the wire.
A bias dependent dierence in absolute STM tip height is observed across the stacking fault where for negative
bias the position of the tip was higher below (bottom of the image) as compared to above the stacking fault (top of
image) for negative voltages and reversed for positive voltages. is can be seen in Fig.5 for Vb = 0.2 V, −2 V and
−4 V. We see an increase in the gradient of tip height change from higher (darker blue) to lower (darker red) as
Vb is increased from 0 V to −4 V. is is as expected for a local drop in VL across the stacking fault that increases
in gradient with Vb. e change in local potential induced by the bias along the nanowire can be seen in the
potential drop across the 20 nm section of the InAs nanowire as approximately 0.04 V for Vb = −4 V. A nanowire
feature such as a stacking fault, large defect or heterojunction has a nite electrical resistance and hence can aect
the nanowire conductance. Indeed, while the change in VL was approximately linear either side of the stacking
fault - Fig.5d - as would be expected from two regions of InAs nanowire with similar resistance, a rather dierent
potential behaviour is seen across the stacking fault.
We rule out several other explanations for the eects observed. Firstly, the eect of changing Vb (and thus local
surface potential) on tip height is somewhat equivalent to changing the tip bias Vtip. At dierent Vtip, features may
appear more or less prominent. ese eects should be uniform across the image, being a product of the absolute
change in tip height as a result in change in tip potential and should show no dependence in the direction of
current ow/applied bias (in y, along the nanowire). Figure5b shows subtracted height data for images taken at
Vtip = −1.7 V and Vtip = −1.3 V with the nanowire electrically grounded. No clear dierence across the stacking
fault was observed, with no dependence on y and a uniform or no height increase across the area of interest.
Secondly, a change in the tip conguration could also induce changes between subsequent subtracted images. In
Fig.5c we show 2 consecutive subtracted images taken at the same applied (Vb = −1 V) and tip bias (−1.5 V) in
order to demonstrate that that this cannot account for the eects observed. irdly, in addition to the robust eect
across the stacking fault we observe localised changes in d(x,y) corresponding to mobile surface features - mobile
clusters of surface atoms moved and redeposited by the tip from image to image.
Figure 5. Local nanowire surface potential measured by STM topography. In (a) (top le) we show a
20 × 20 nm2 STM topography scan image on an InAs nanowire of a stacking fault (SF) at Vb = 0 V, where y
indicates the direction along the length of the nanowire between the contact electrodes. is is shown along
with (top right, bottom) the subtracted dierence between extrapolated local potential VL between Vb = 0 and
Vb = +0.2, −2 V and −4 V (labelled) with Vtip = −1.3 V, showing the eect of increasing applied bias. A clear
trend was observed of increasing dierence in VL in y across either side of the stacking fault with increasing Vb.
As a control to show this is not a tip artefact we show in (b) 20 × 20 nm subtracted dierence in extrapolated
local potential between scans of this area at Vtip = −1.7 V and Vtip = −1.3 V at Vb = 0 V, with the change of
tip bias having eects an order of magnitude below that in (a). We also show another control in (c) with a
20 × 20 nm subtraction of two consecutive scans at the same Vb = −1 V, with no gradiation in y of the local
potential observed. (d) shows a 2D plot of Vb = −4 V data in (a), taking the average calculated VL across a 20 nm
section. Here the stacking fault position is indicated by an arrow, with the red/blue arrows referring to the
colourscheme used in (a) to better relate STM topography and VL.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
8
SCientifiC REPORTS | 7: 12790 | DOI:10.1038/s41598-017-13007-w
Our method has advantages over alternatives such as scanning tunnelling spectroscopy (STS)31 and scanning
probe potentiometry (SPP)47,48 since it does not require a xed sample-tip height or a at surface and can be per-
formed at a single constant tip and sample bias.
Discussion
We have demonstrated the rst simultaneous direct atomic scale STM and electrical operation on semiconductor
nanowire devices. is is a signicant step forward over previous SPM studies of biased nanostructures32,33,49
done at lower resolution, with a limited materials choice or on much larger thin lm surfaces19–21. It is also an
advance over our previous studies with combined AFM/STM measurements that we can remove surface oxide
through atomic hydrogen cleaning without destroying the device, that we can scan on the nanowire without
crashing onto an adjacent oxide layer and that STM oers considerably improved stability and resolution over
AFM scanning on a curved nanowire surface. We combine all the properties needed for studies of the structural
dynamics on nanoscale surfaces during device operation. is includes fast identication of a device and long
term stable scanning even with voltage changes across the devices of several volts. Surface preparation can be
performed with reactive gas and at elevated temperatures, a prerequisite for obtaining well dened surfaces. In
contrast to our previous work, the method can be implemented in standard STM-only systems as well as low
temperature systems without requiring additional in-situ microscopy tools such as SEM. Device fabrication use
only commonly available processing technology.
Because InAs and GaAs have very dierent thermal, electrical and mechanical properties50, our high quality
imaging on both materials (and on dierent crystal structures) indicate that our concepts should be applicable
on many other nanostructured materials. e new device design also limits detrimental surface contamination
of the wires due to processing steps as the nanowires can be deposited in the last step. Its fully conducting nature
(when viewed from above) eliminates problems with charging and can therefore also be used in other important
(electron based) microscopies such as surface sensitive photoelectron emission microscopy (PEEM)51 or SEM.
Our initial studies already reveal the interesting phenomena of device bias self-healing on InAs nanowire sur-
faces, as atomic scale defects were removed and morphology smoothed. is was an extremely surprising result
since we would expect well known electromigration of In atoms in the material to produce and accumulate more
defects and eventual structural failure of the nanowire. We observed no clear failure region nor did we observe
an increase in electrical resistance associated with nanodevice failure, instead seeing reduced electrical resistance
with defect removal and surface smoothing. We speculate that this is due to electric-eld driven reordering (‘eld
annealing’) of the surface at below the threshold required for the nanowire to fail. Future measurements beyond
the scope of the present proof-of-principle work will be needed to elucidate these eects in more detail.
While the importance of in-situ studies of semiconductor nanowires during growth has been clearly shown52,
this demonstrates that surface structural change during operation can also be unusual and opens up possibilities
for creating highly ordered crystalline surfaces with the use of bias voltages. Atomic scale structural changes
due to induced voltages are central to devices that rely on defect diusion and very small electric eld induced
structural changes for their function, such as memristors53,54. Our ability to rapidly switch voltages, and the small
thermal eects can allow assessment of the impact of dierent suggested mechanism for structural change such
as heating, electric current and pure voltage gradient eects53. Understanding structural dynamics will be central
to go beyond trial-and-error in addressing the signicant problems of performance degradation and failure over
time found in many nanoscale structures.
Methods
Device manufacturing. Starting with n-type Si substrate HF etched free of thermally grown oxide both
front and backside, we patterned by EBL in ma-n2403 negative resist strips of similar width as the length of the
nanowires to be studied (and signicantly shorter than the total length of the nanowire). We then deposited
60 nm SiO or SiO2 followed by a thin 15 nm Ti layer and 2 nm Au capping layer onto this pattern, leaving a trench
in this material aer li-o and the Ti/Au layers on each side of it electrically separated from each other (typical
resistance >5 M) and the substrate. We choose to construct in this manner rather than the alternative of etching
a trench into Ti-covered SiO2 on Si due to a) the greater control, consistency and resolution the negative resist
EBL patterning aorded over wet etching, b) to prevent sidewall etching that could undercut the metal electrode
under the nanowire and c) due to the preference to use Ti as the contact metal (also etched by standard SiO2
HF-based wet etching). We also wished to ensure perfectly conducting Si at the bottom of the trench with no
patches of remaining residual oxide. We found the 2 nm Au cap under the nanowire to be suciently thin so as
to not destroy the nanowire by alloying eects, whilst ensuring a highly conductive oxide-free surface on which
to scan. We then deposited the nanowires by mechanical transfer, leaving a small number suspended between the
Ti/Au electrodes. A second EBL step could be used to pattern a Ti ‘holding’ layer onto the nanowire on each side
of the trench with thickness of 15 nm. e weak mechanical grip provided by the Ti holding layer was designed
to ensure that the STM tip could not pull the nanowire from the metal surface whilst not inducing strain in the
nanowire on heating. e thickness of this layer was empirically determined by thermally cycling a series of test
devices at dierent layer thickness.
Device preparation for STM. External electrical connection to the device was made by wire bonding to
the Ti layer each side of the trench, with the rear of the sample grounded or optionally connected to a gate bias.
Markers were patterned to allow the approximate initial positioning of the STM tip near a specic nanowire of
interest and subsequent location of the relevant nanowire device without use of additional microscopy such as
in vacuum Scanning Electron Microscopy (SEM). When bonded and mounted on a conducting sample plate the
result is a fully conductive surface for the STM tip. is permits scanning anywhere on the sample surface whilst
electrically isolating the nanowires from ground and allowing a bias to be applied only across them. e method
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
9
SCientifiC REPORTS | 7: 12790 | DOI:10.1038/s41598-017-13007-w
retains the standard fabrication techniques and materials (Si, Ti, Au) in order to provide a realistic alternative that
can be easily fabricated and keep many of the properties as the conventional lateral design.
Nanowire location procedure. Using images taken elsewhere by optical microscope for navigation and
identication of our position of the surface using the markers patterned on the surface by lithographic process-
ing, we approximately positioned the STM tip using the standard tip-sample coarse positioning telescope camera
(resolution 50 μm). is could locate the tip within 100 μm of the nanowire device. By moving in the direction
of travel towards the trench, taking 10 μm length STM scans in this direction, we could locate it’s position and
then follow it’s edge to navigate to the relevant nanowire. is procedure required between 30 minutes and 1 hour
to locate the desired nanowire within the STM scan window. rough further renement and reducing the scan
size, we then positioned the tip onto the nanowire on top of the contact, aer which we could move freely along
it’s length.
Device in-vacuum cleaning and STM imaging. To enable STM experiments on crystalline clean nano-
wire surfaces, we removed the native oxide from the nanowires by exposing the full device to atomic hydrogen
(H*) from a W lament thermal cracker at 1700 C for periods of 20 min at 350–450 C at chamber hydrogen
pressure 1 × 10−6 mbar as described in detail elsewhere27. All preparation and STM imaging was performed in
the same commercial Omicron system with ultra-high vacuum chamber (base pressure 1 × 10−10 mbar). We used
standard tungsten wire STM tips, cleaned before imaging by Ar+ sputtering at 5 × 10−6 mbar chamber pressure of
Ar. All images shown were recorded at negative sample voltages (lled state imaging) in the range between −1 to
−2 V and with currents 0.05–0.2 nA. Cleaning was performed with the device sample mounted and electrically
connected by wire bonding onto a sample plate with up to four Au electrical contacts that could be accessed from
outside the vacuum chamber while the sample was in the STM. No indications of damage could be observed aer
cleaning in the STM system or subsequent ex-situ SEM images. We note a side eect from the cleaning process
was to considerably enhance the conductance of the nanowire device (by several orders of magnitude) compared
to the as-fabricated version prior to cleaning in air. is is shown in the I–V characteristics of an example device
given in Supplementary Information Fig.6.
References
1. Ueda, O. eliability issues in III–V compound semiconductor devices: optical devices and GaAs-based HBTs. Microelectronics
Reliability 39, 1839–1855 (1999).
2. oenraad, P. M. & Flatté, M. E. Single dopants in semiconductors. Nature Materials 10, 91–100 (2011).
3. Del Alamo, J. Nanometre-scale electronics with III–V compound semiconductors. Nature 479, 317–323 (2011).
4. Hjort, M. et al. Surface chemistry, structure, and electronic properties from microns to the atomic scale of axially doped
semiconductor nanowires. ACS Nano 6, 9679–9689, https://doi.org/10.1021/nn303107g (2012).
5. Johansson, S., Ghalamestani, S. G., Borg, M., Lind, E. & Wernersson, L. Temperature and annealing eects on InAs nanowire
MOSFETs. Microelectronic Engineering 88, 1105–1108, Proceedings of the 17th Biennial International Insulating Films on
Semiconductor Conference 17th Biennial International Insulating Films on Semiconductor Conference (2011).
6. Houssa, M., Chagarov, E. & ummel, A. Surface Defects and Passivation of Ge and III–V Interfaces. MRS Bulletin 34, 504–513
(2009).
7. Fang, H. et al. Degenerate n-Doping of Few-Layer Transition Metal Dichalcogenides by Potassium. Nano Letters 13, 1991–1995
(2013).
8. Balog, . et al. Bandgap opening in graphene induced by patterned hydrogen adsorption. Nature Materials 9, 315–319 (2010).
9. Martnez-Blanco, J. et al. Gating a single-molecule transistor with individual atoms. Nature Physics 11, 640–644 (2015).
10. Fuechsle, M. et al. A single-atom transistor. Nature Nanotec hnology 7, 242–246 (2012).
11. Mönch, W. Semiconductor Surfaces and Interfaces. Springer Series in Surface Sciences (Springer Berlin Heidelberg, 1993).
12. Ebert, H. Imaging defects and dopants. Materials Today 6, 36–43 (2003).
13. Ebert, P. Nano-scale properties of defects in compound semiconductor surfaces. Surface Science Reports 33, 121–303 (1999).
14. hajetoorians, A. A. et al. D etecting excitation and magnetization of individual dopants in a semiconductor. Nature 467, 1084–1087
(2010).
15. Mielsen, A. et al. Direct imaging of the atomic structure inside a nanowire by scanning tunnelling microscopy. Nature Materials
3, 519–523 (2004).
16. Xue, Q.-., Hashizume, T. & Sa urai, T. Scanning tunneling microscopy of III–V compound semiconductor (001) surfaces. Progress
in Surface Science 56, 1–131 (1997).
17. Mielsen, A. & Lundgren, E. Cross-sectional scanning tunneling microscopy studies of novel III–V semiconductor structures.
Progress in Surface Science 80, 1–25 (2005).
18. Timm, . et al. Self-organized formation of GaSb/GaAs quantum rings. Phys. Rev. Lett. 101, 256101 (2008).
19. Tao, C., Cullen, W. G. & Williams, E. D. Visualizing the Electron Scattering Force in Nanostructures. Science 328, 736 (2010).
20. Bondarchu, O. et al. Biased Surface Fluctuations due to Current Stress. Phys. Rev. Lett. 99, 206801 (2007).
21. Ho, M.-S., Hwang, I.-S. & Tsong, T. T. Direct Observation of Electromigration of Si Magic Clusters on Si(111) Surfaces. Phys. Rev.
Lett. 84, 5792–5795 (2000).
22. Lehmann, S., Wallentin, J., Jacobsson, D., Deppert, . & Dic, . A. A General Approach for Sharp Crystal Phase Switching in InAs,
GaAs, InP, and GaP Nanowires Using Only Group V Flow. Nano Letters 13, 4099–4105 (2013).
23. Hyun, J. ., Zhang, S. & Lauhon, L. J. Nanowire Heterostructures. Annual Review of Materials Science 43, 451–479 (2013).
24. Yang, P., Yan, . & Fardy, M. Semiconductor Nanowire: What’s Next? Nano Letters 10, 1529–1536 (2010).
25. Zota, C. B., oll, G., Wernersson, L.-E. & Lind, E. adio-Frequency Characterization of Selectively egrown InGaAs Lateral
Nanowire MOSFETs. IEEE Transactions on Electron Devices 61, 4078–4083 (2014).
26. Wallentin, J. et al. InP Nanowire Array Solar Cells Achieving 13.8% Eciency by Exceeding the ay Optics Limit. Science 339,
1057–1060 (2013).
27. Hjort, M. et al. Direct Imaging of Atomic Scale Structure and Electronic Properties of GaAs Wurtzite and Zinc Blende Nanowire
Surfaces. Nano Letters 13, 4492–4498 (2013).
28. Hjort, M. et al. Electronic and Structural Dierences between Wurtzite and Zinc Blende InAs Nanowire Surfaces: Experiment and
e ory. ACS Nano 8, 12346–12355 (2014).
29. Schroer, M. D. & Petta, J. . Correlating the Nanostructure and Electronic Properties of InAs Nanowires. Nano Letters 10, 1618–1622
(2010).
30. Miao, J. et al. Single InAs Nanowire oom-Temperature Near-Infrared Photodetectors. ACS Nano 8, 3628–3635 (2014).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
10
SCientifiC REPORTS | 7: 12790 | DOI:10.1038/s41598-017-13007-w
31. Persson, O. et al. Scanning Tunneling Spectroscopy on InAs–GaSb Esai Diode Nanowire Devices during Operation. Nano Letters
15, 3684–3691 (2015).
32. Zhou, X., Dayeh, S. A., Aplin, D., Wang, D. & Yu, E. T. Direct observation of ballistic and dri carrier transport regimes in InAs
nanowires. Applied Physics Letters 89, 053113 (2006).
33. Halpern, E., Elias, G., retinin, A. V., Shtriman, H. & osenwas, Y. Direct measurement of surface states density and energy
distribution in individual InAs nanowires. Applied Physics Letters 100, 262105 (2012).
34. Webb, J. et al. High resolution scanning gate microscopy measurements on InAs/GaSb nanowire Esai diode devices. Nano Research
7, 877–887 (2014).
35. Hilner, E. et al. Direct Atomic Scale Imaging of III–V Nanowire Surfaces. Nano Letters 8, 3978–3982 (2008).
36. elander, C. et al. Nanowire-based one-dimensional electronics. Materials Today 9, 28–35 (2006).
37. Melitz, W. et al. Scanning tunneling spectroscopy and elvin probe force microscopy investigation of Fermi energy level pinning
mechanism on InAs and InGaAs clean surfaces. Journal of Applied Physics 108, 023711 (2010).
38. ratzer, P., Penev, E. & Scheer, M. First-principles studies of inetics in epitaxial growth of III–V semiconductors. Applied Physics
A 75, 79–88 (2002).
39. Menges, F. et al. Temperature mapping of operating nanoscale devices by scanning probe thermometry. Nature Communications 7,
10874 (2016).
40. Wu, N.-J., Yasunaga, H. & Natori, A. Surface mass transport of Ag and In on vicinal Si(111). Surface Science 260, 75–84 (1992).
41. Shi, F., Shirai, I., Nagao, T. & Hasegawa, S. Diffusion anisotropy of Ag and In on Si(1 1 1) surface studied by UHV-SEM.
Ultramicroscopy 85, 23–33 (2000).
42. ang, D., im, T., Bae, C.-., Meyyappan, M. & Lee, J.-S. Investigation of electromigration in In2Se3 nanowire for phase change
memory devices. Applied Physics Letters 103, 233504 (2013).
43. Zhang, C. et al. In situ electrical characterization of tapered InAs nanowires in a transmission electron microscope with ohmic
contacts. Nanotechnology 26, 155703 (2015).
44. Höglund, A., Castleton, C. W. M., Göthelid, M., Johansson, B. & Mirbt, S. Point defects on the (110) surfaces of InP, InAs, and InSb:
A comparison with bul. Phys. Rev. B 74, 075332 (2006).
45. Pronschinse, A., Mardit, D. J. & Dougherty, D. B. Modeling the constant-current distance-voltage mode of scanning tunneling
spectroscopy. Physical Review B 84, 205427 (2011).
46. elander, C., Caro, P., Plissard, S., Dey, A. W. & Dic, . A. Eects of Crystal Phase Mixing on the Electrical Properties of InAs
Nanowires. Nano Letters 11, 2424–2429 (2011).
47. Dong, Y., Feenstra, . M., Hey, . & Ploog, . H. Scanning tunneling potentiometry of semiconductor junctions. Journal of Vacuum
Science and Technology B 20, 1677–1681 (2002).
48. Schneider, M. A., Wenderoth, M., Heinrich, A. J., osentreter, M. A. & Ulbrich, . G. Current transport through single grain
boundaries: A scanning tunneling potentiometry study. Applied Physics Letters 69 (1996).
49. Leroy, B. J., Heller, I., Pahilwani, V. ., Deer, C. & Lemay, S. G. Simultaneous Electrical Transport and Scanning Tunneling
Spectroscopy of Carbon Nanotubes. Nano Letters 7, 2937–2941 (2007).
50. Adachi, S. Physical Properties of III–V Semiconductor Compounds. Wiley-Interscience publication (Wiley, 1992).
51. Mielsen, A. & Lundgren, E. Surface science of free standing semiconductor nanowires. Surface Science 607, 97–105 (2013).
52. Jacobsson, D. et al. Interface dynamics and crystal phase switching in GaAs nanowires. Nature 531, 317–322 (2016).
53. Yang, J. J., Struov, D. B. & Stewart, D. . Memristive devices for computing. Nature Nanotechnolo gy 8, 13–24 (2013).
54. Sun, J., Lind, E., Maximov, I. & Xu, H. Q. IEEE Electron Device Letters 32, 131–133 (2011).
Acknowledgements
is work was performed within the NanoLund Center for Nanoscience at Lund University, and was supported
by the Swedish Research Council (VR), the Swedish Foundation for Strategic Research (SSF), the Crafoord
Foundation, the Knut and Alice Wallenberg Foundation, and the European Research Council under the European
Union’s Seventh Framework Programme Grant Agreement No. 259141.
Author Contributions
J.L.W., A.M., R.T. and C.T. designed the experiment and devices. J.L.W. fabricated the devices. S.L. and K.D.T.
designed and synthesized the nanowires. J.L.W., J.K., M.H., S.M. performed the measurements. J.L.W., A.M., R.T.,
J.K., C.T. analyzed and interpreted the data. J.L.W. and A.M. wrote the manuscript with contributions from all
authors. Correspondence to A. Mikkelsen.
Additional Information
Supplementary information accompanies this paper at https://doi.org/10.1038/s41598-017-13007-w.
Competing Interests: e authors declare that they have no competing interests.
Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre-
ative Commons license, and indicate if changes were made. e images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons license and your intended use is not per-
mitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the
copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
© e Author(s) 2017
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
Available via license: CC BY
Content may be subject to copyright.