growth, intermixing, and surface phase formation for zinc ... · growth, intermixing, and surface...

10
Growth, intermixing, and surface phase formation for zinc tin oxide nanolaminates produced by atomic layer deposition Citation for published version (APA): Hägglund, C., Grehl, T., Tanskanen, J. T., Yee, Y. S., Mullings, M. N., Mackus, A. J. M., ... Bent, S. F. (2016). Growth, intermixing, and surface phase formation for zinc tin oxide nanolaminates produced by atomic layer deposition. Journal of Vacuum Science and Technology A: Vacuum, Surfaces, and Films, 34(2), [021516]. https://doi.org/10.1116/1.4941411 DOI: 10.1116/1.4941411 Document status and date: Published: 08/02/2016 Document Version: Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication: • A submitted manuscript is the version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website. • The final author version and the galley proof are versions of the publication after peer review. • The final published version features the final layout of the paper including the volume, issue and page numbers. Link to publication General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal. If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please follow below link for the End User Agreement: www.tue.nl/taverne Take down policy If you believe that this document breaches copyright please contact us at: [email protected] providing details and we will investigate your claim. Download date: 18. May. 2020

Upload: others

Post on 19-May-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Growth, intermixing, and surface phase formation for zinc ... · Growth, intermixing, and surface phase formation for zinc tin oxide nanolaminates produced by atomic layer deposition

Growth, intermixing, and surface phase formation for zinc tinoxide nanolaminates produced by atomic layer depositionCitation for published version (APA):Hägglund, C., Grehl, T., Tanskanen, J. T., Yee, Y. S., Mullings, M. N., Mackus, A. J. M., ... Bent, S. F. (2016).Growth, intermixing, and surface phase formation for zinc tin oxide nanolaminates produced by atomic layerdeposition. Journal of Vacuum Science and Technology A: Vacuum, Surfaces, and Films, 34(2), [021516].https://doi.org/10.1116/1.4941411

DOI:10.1116/1.4941411

Document status and date:Published: 08/02/2016

Document Version:Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Please check the document version of this publication:

• A submitted manuscript is the version of the article upon submission and before peer-review. There can beimportant differences between the submitted version and the official published version of record. Peopleinterested in the research are advised to contact the author for the final version of the publication, or visit theDOI to the publisher's website.• The final author version and the galley proof are versions of the publication after peer review.• The final published version features the final layout of the paper including the volume, issue and pagenumbers.Link to publication

General rightsCopyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright ownersand it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, pleasefollow below link for the End User Agreement:www.tue.nl/taverne

Take down policyIf you believe that this document breaches copyright please contact us at:[email protected] details and we will investigate your claim.

Download date: 18. May. 2020

Page 2: Growth, intermixing, and surface phase formation for zinc ... · Growth, intermixing, and surface phase formation for zinc tin oxide nanolaminates produced by atomic layer deposition

Growth, intermixing, and surface phase formation for zinc tin oxide nanolaminatesproduced by atomic layer depositionCarl Hägglund, Thomas Grehl, Jukka T. Tanskanen, Ye Sheng Yee, Marja N. Mullings, Adriaan J. M. Mackus,Callisto MacIsaac, Bruce M. Clemens, Hidde H. Brongersma, and Stacey F. Bent

Citation: Journal of Vacuum Science & Technology A: Vacuum, Surfaces, and Films 34, 021516 (2016); doi:10.1116/1.4941411View online: http://dx.doi.org/10.1116/1.4941411View Table of Contents: http://avs.scitation.org/toc/jva/34/2Published by the American Vacuum Society

Articles you may be interested inIncomplete elimination of precursor ligands during atomic layer deposition of zinc-oxide, tin-oxide, and zinc-tin-oxideThe Journal of Chemical Physics 146, 052802 (2016); 10.1063/1.4961459

Tin oxide atomic layer deposition from tetrakis(dimethylamino)tin and waterJournal of Vacuum Science & Technology A: Vacuum, Surfaces, and Films 31, 061503 (2013);10.1116/1.4812717

Atomic layer deposition of tin oxide films using tetrakis(dimethylamino) tinJournal of Vacuum Science & Technology A: Vacuum, Surfaces, and Films 26, 244 (2008); 10.1116/1.2835087

Surface chemistry of atomic layer deposition: A case study for the trimethylaluminum/water processJournal of Applied Physics 97, 121301 (2005); 10.1063/1.1940727

Atomic layer deposited zinc tin oxide channel for amorphous oxide thin film transistorsApplied Physics Letters 101, 113507 (2012); 10.1063/1.4752727

In situ synchrotron based x-ray techniques as monitoring tools for atomic layer depositionJournal of Vacuum Science & Technology A: Vacuum, Surfaces, and Films 32, 010801 (2013);10.1116/1.4851716

Page 3: Growth, intermixing, and surface phase formation for zinc ... · Growth, intermixing, and surface phase formation for zinc tin oxide nanolaminates produced by atomic layer deposition

Growth, intermixing, and surface phase formation for zinc tin oxidenanolaminates produced by atomic layer deposition

Carl H€agglunda)

Department of Chemical Engineering, Stanford University, Stanford, California 94305 and Departmentof Engineering Sciences, Division of Solid State Electronics, Uppsala University, 75121 Uppsala, Sweden

Thomas GrehlION-TOF GmbH, Heisenbergstraße 15, 48149 M€unster, Germany

Jukka T. TanskanenDepartment of Chemical Engineering, Stanford University, Stanford, California 94305

Ye Sheng YeeDepartment of Electrical Engineering, Stanford University, Stanford, California 94305

Marja N. Mullings, Adriaan J. M. Mackus, and Callisto MacIsaacDepartment of Chemical Engineering, Stanford University, Stanford, California 94305

Bruce M. ClemensDepartment of Material Science and Engineering, Stanford University, Stanford, California 94305

Hidde H. BrongersmaION-TOF GmbH, Heisenbergstraße 15, 48149 M€unster, Germany

Stacey F. Bentb)

Department of Chemical Engineering, Stanford University, Stanford, California 94305

(Received 20 November 2015; accepted 25 January 2016; published 8 February 2016)

A broad and expanding range of materials can be produced by atomic layer deposition at relatively

low temperatures, including both oxides and metals. For many applications of interest, however, it

is desirable to grow more tailored and complex materials such as semiconductors with a certain

doping, mixed oxides, and metallic alloys. How well such mixed materials can be accomplished

with atomic layer deposition requires knowledge of the conditions under which the resulting films

will be mixed, solid solutions, or laminated. The growth and lamination of zinc oxide and tin oxide

is studied here by means of the extremely surface sensitive technique of low energy ion scattering,

combined with bulk composition and thickness determination, and x-ray diffraction. At the low

temperatures used for deposition (150 �C), there is little evidence for atomic scale mixing even

with the smallest possible bilayer period, and instead a morphology with small ZnO inclusions in a

SnOx matrix is deduced. Postannealing of such laminates above 400 �C however produces a stable

surface phase with a 30% increased density. From the surface stoichiometry, this is likely the

inverted spinel of zinc stannate, Zn2SnO4. Annealing to 800 �C results in films containing crystal-

line Zn2SnO4, or multilayered films of crystalline ZnO, Zn2SnO4, and SnO2 phases, depending on

the bilayer period. VC 2016 American Vacuum Society. [http://dx.doi.org/10.1116/1.4941411]

I. INTRODUCTION

Zinc tin oxide (ZTO) is a wide gap semiconductor of in-

terest as an earth abundant, nontoxic transparent conductive

oxide.1 It is a promising alternative to cadmium sulfide

buffer layers in thin film solar cells,2,3 as a channel layer in

thin film transistors,4,5 and for photocatalytic applications.6

It is also employed in varistors and gas sensors.7

The deposition of ZTO by atomic layer deposition (ALD)

is desirable as this technique allows for uniform and confor-

mal growth on three dimensional, nanoscale topographies.8

It is further of much interest to be able to produce ZTO at

low temperatures making the processing compatible with a

broader range of substrate materials and applications, and

hence enabling lower production costs. To this end, the

production of ZTO of varying (average) stoichiometry has

previously been investigated by stacking thin ZnO and SnOx

layers by low temperature ALD.9–11 In this way, a homoge-

neous ternary compound has been approached by repeating a

supercycle of the individual Zn and Sn oxide ALD sequen-

ces. The atomic scale control of the oxide sublayer thick-

nesses provided by the design of the supercycle allows for

fine tuning the compound properties toward specific condi-

tions. Critical questions for this approach are to what extent

the ZnO and SnOx sublayers mix, and if a solid solution can

be formed at low temperature.

In this work, stacks of zinc and tin oxide layers were de-

posited using supercycles with bilayer periods ranging from

a few cycles up to 800 cycles. Based on measurements by

spectroscopic ellipsometry (SE), inductively coupled plasma

optical emission spectroscopy (ICP-OES), and low energy

ion scattering (LEIS), the type of growth, the degree of mix-

ing, and the mechanism by which mixing takes place were

a)Electronic mail: [email protected])Electronic mail: [email protected]

021516-1 J. Vac. Sci. Technol. A 34(2), Mar/Apr 2016 0734-2101/2016/34(2)/021516/8/$30.00 VC 2016 American Vacuum Society 021516-1

Page 4: Growth, intermixing, and surface phase formation for zinc ... · Growth, intermixing, and surface phase formation for zinc tin oxide nanolaminates produced by atomic layer deposition

analyzed. The results are compatible with a substrate inhib-

ited growth of ZnO on SnOx, associated with island type

growth, and a more layer-by-layer like growth of SnOx on

ZnO. Annealing above 400 �C leads to the formation of a

dense and stable surface phase, identified as an inverted spi-

nel of zinc stannate (Zn2SnO4), while annealing to 800 �C is

required to produce crystalline Zn2SnO4 in the bulk of the

film. The appearance of the surface phase is of importance

for a more detailed understanding of the effect of

postannealing.

II. EXPERIMENT

A. Atomic layer deposition

A custom built,12 hot-wall laminar flow reactor was

employed for the ALD. Diethyl zinc (DEZn; Sigma-Aldrich)

and tetrakis(dimethylamido)tin (TDMASn; Strem, >99%

purity) were used as metal precursors with deionized water

as the oxygen source. The precursor dosing was controlled

by means of needle valves and the pulse time of computer

controlled air-actuated valves. The water and DEZn sample

vials were kept at room temperature, whereas the TDMASn

vial was maintained at 45 �C. The gas composition was con-

tinuously monitored by means of an in situ mass spectrome-

ter mounted with an orifice downstream the sample position.

B. Spectroscopic ellipsometry

Standard ellipsometric measurements were performed on

a Woollam M2000 system over wavelengths from 210 to

1700 nm and angles of incidence of 65�, 70�, and 75�. The

data were analyzed using an oscillator model for the

unknown optical properties of the ZTO layer and tabulated

data for the Si substrate. In the simplest model, the ZTO

stack was assumed to have isotropic and effectively homoge-

neous optical properties throughout. The total ZTO thickness

was fitted together with a single polynomial spline function

to represent the imaginary part of the permittivity in the

measured range. The real part of the permittivity was

obtained from a Kramers-Kronig transformation, with

absorption outside the measured range taken into account

through the addition of poles in the ultraviolet and infrared,

respectively. In all, ten parameters plus the ZTO thickness

were fitted. In a more refined repeated bilayer model, dis-

crete but isotropic and homogeneous sublayers were

assumed. The ZnO and SnOx sublayer properties were fitted

separately by two different polynomial spline oscillator

functions. All sublayers of the same type were assumed to

have identical thicknesses and properties in the stack. In all,

20 parameters of the oscillator functions plus two sublayer

thicknesses were then fitted.

C. Inductively coupled plasma optical emissionspectroscopy

Films on Si substrates were dissolved in 1 ml concen-

trated aqua regia (HCl:HNO3 3:1), then diluted and soni-

cated. The concentration of dissolved zinc and tin ions in

solution were analyzed using an Inductively Coupled Plasma

Spectrometer (Thermo Scientific, ICAP 6300 Duo View).

D. Low energy ion scattering

ZTO films deposited on Si were cleaned by atomic O to

remove organic surface contaminations resulting from sam-

ple exposure to air. The atomic O was generated by dissocia-

tion of molecular oxygen in a plasma. The radicals

generated were filtered, so that only low kinetic energy

atomic O and no charged or energetic particles could reach

the sample. For the analysis, an ION-TOF Qtac 100 was

used with 3 keV He (for survey spectra) and 5 keV Ne scat-

tering (for quantification of metals). Ion fluencies resulted in

sputtering of less than 1% of the outer monolayer (“static

analysis”). This means that the surface was analyzed in its

nearly pristine state. Depth profiles were obtained by inter-

leaving spectra under static conditions with 500 eV Ar sput-

tering phases. During each sputtering phase, a dose of

approximately 4� 1014 ions/cm2 was applied, removing

about 1=4 of a monolayer.

E. Grazing incidence x-ray diffraction

Grazing incidence x-ray diffraction (GIXRD) was carried

out at beamline 11–3 at the Stanford Synchrotron Radiation

Lightsource (SSRL), using 12.7 keV x-rays. The MAR345

2D imaging detector was positioned 150 mm away from the

sample, which was calibrated with a LaB6 sample, and the

incidence angle was set to 2�. The data from the 2D detector

were converted into XRD patterns as a function of Q¼ 2p/dusing the WxDiff software package. The powder diffraction

files used for reference were PDF 00-036-1451 for ZnO,

PDF 00-041-1445 for SnO2, and PDF 00-024-1470 for

Zn2SnO4, respectively.

III. RESULTS AND DISCUSSION

A custom built12 ALD system was employed to grow

alternating layers of zinc oxide (ZnO) and tin oxide (SnOx,

1< x< 2). In this process, diethyl zinc13 and tetrakis(dime-

thylamino)tin14 were used as the zinc and tin precursors,

respectively, with water as the counter-reactant.2,9 We use

here the term supercycle to describe the sequence of ALD

cycles that produces a single ZnO-SnOx bilayer that is then

repeated in the stack. The total number of ALD cycles in

each supercycle is the bilayer period. Samples of varying

sublayer thicknesses (i.e., the thickness of a ZnO or SnOx

layer within a single bilayer) were produced on Si(100) sub-

strates using supercycles with a fixed ratio of 1:3 ZnO:SnOx

ALD cycles. This cycle ratio corresponds roughly to a 50:50

ratio of elemental Zn to Sn based on the growth rates of each

individual oxide.9 As discussed further below, however, the

growth rates vary substantially depending on the type and

condition of the growth surface. The shortest bilayer period

consisted of four cycles, of which one was a ZnO cycle

and three were SnOx cycles. Bilayer periods of 4, 8, 20, 60,

200, and 800 cycles were studied, resulting in a broad

range of sublayer thicknesses. The substrate was maintained

at a temperature of 150 �C, and the precursor doses were

021516-2 H€agglund et al.: Growth, intermixing, and surface phase formation 021516-2

J. Vac. Sci. Technol. A, Vol. 34, No. 2, Mar/Apr 2016

Page 5: Growth, intermixing, and surface phase formation for zinc ... · Growth, intermixing, and surface phase formation for zinc tin oxide nanolaminates produced by atomic layer deposition

continuously monitored during the deposition process by

means of an in situ mass spectrometer.

Two series of samples were produced with the set of

bilayer periods as described, but with surface terminations of

either ZnO or SnOx layers as illustrated in Fig. 1. The cases

with a single terminating ALD cycle are of particular interest

for studying the nucleation process. However, since a mini-

mum of 3 SnOx cycles was used in the samples, an additional

sample was prepared where the termination was with a sin-

gle SnOx cycle but the bilayers underneath contained 1

ZnOþ 3 SnOx cycles. Hence, the bilayer period of 4 cycles

in the bulk of the stack is duplicated for the SnOx terminated

series.

The total thicknesses of the deposited ZTO laminates

were examined by SE with the results shown in Figs. 1(a)

and 1(b). Individual layer thicknesses were not determined

by this analysis, as the stack was approximated as an opti-

cally homogeneous layer with effective optical constants in

the ellipsometry model. This approach led to relatively large

mean squared deviation from the measured ellipsometry data

for the samples with the thickest sublayers. In these cases, a

model using a repeated bilayer structure where the ZnO and

SnOx sublayer properties were distinguished was able to

reduce the error substantially. However, this refinement did

not make any significant difference to the total thickness

when compared to the simpler homogeneous film model.

The relative bulk fractions (defined as the elemental mole

fraction of Zn or Sn relative to the summed mole fractions of

Zn and Sn in the sample) of the laminates were subsequently

determined using ICP-OES, as described in Sec. II. From

these measurements, the total amounts of Zn and Sn in the

samples were obtained. Assuming that the densities of the

deposited ZnO and SnOx phases corresponded to the bulk

densities of hexagonal ZnO and rutile SnO2, respectively,

the total volume of ZnO and SnOx in the laminates was esti-

mated. If this material is distributed in perfectly discrete sub-

layers corresponding to the ALD sequences, the stacks

illustrated in Figs. 1(a) and1(b) would result. The actual

sample structures are an intermediate between these per-

fectly layered stacks and a homogeneous mixture of the

elements.

The amount (expressed as sublayer thickness) of ZnO and

SnOx deposited per supercycle is shown as a function of the

number of sublayer cycles in Fig. 1(c). By interpolating

these data with cubic spline functions, differentiation gives

an estimate of the instantaneous growth rates for the binaries

[Fig. 1(d)]. Because the initial growth is not linear with the

number of cycles per sublayer, this instantaneous rate differs

from the average growth rate obtained by dividing the esti-

mated sublayer thicknesses by the number of cycles per

sublayer.

The details of the extracted sublayer growth rate are sen-

sitive to the boundary conditions imposed on the spline fit at

either end of the experimental cycle number range, espe-

cially at the boundary at the high cycle number end (large

sublayer thickness), where the interpolation is based on rela-

tively few data points. The natural boundary condition (zero

second derivative) was used for the fits shown. Whether nat-

ural boundary conditions or conditions based on other deriv-

ative orders (>0) are used, however, it is a robust finding

that the ZnO growth rate is suppressed for short bilayer peri-

ods, while for intermediate bilayer periods it substantially

overshoots the steady state ZnO growth rate previously

reported for thicker layers.9,15,16 Suppressed initial growth

and overshooting behavior has previously been associated

with island type growth.17 Mechanistically, this may be

understood as a consequence of a limited number of nuclea-

tion sites found on the SnOx surface, which initially sup-

presses the growth rate of the ZnO. Eventually the surface

area of the growing ZnO islands surpasses that of a corre-

sponding planar growth surface. Temporarily, the enhanced

ZnO surface area then leads to a higher growth rate com-

pared to steady state. As the islands coalesce into a film after

additional sublayer cycles, the steady state growth rate will

be approached. The latter is illustrated by the drawn, dotted

part of the curve in Fig. 1(d). The obtained growth rates also

show that the initial SnOx growth exceeds the steady state

growth rate observed for thicker films.12,14 All of this is

FIG. 1. (a) Layer stacks for ZnO terminated samples, illustrated for simplic-

ity for the special case of perfectly flat, unmixed sublayers. The stack height

represents total film thicknesses, obtained from spectroscopic ellipsometry

for each sample. The sublayers [see panel (c) for thicknesses] are shown to

scale. (b) Same as in panel (a), but for the series of samples terminated by a

SnOx sublayer. (c) The average sublayer thickness within each sample stack,

vs the number of cycles per sublayer. This was obtained by combining the

total stack thickness with the number of sublayers and the elemental frac-

tions measured by ICP-OES and by further assuming sublayer densities

according to crystalline ZnO and SnO2, respectively. The solid curves are

cubic spline interpolations to the data points. (d) Instantaneous growth rates

obtained by differentiation of the spline interpolants in (c). The near steady

state binary oxide growth rates observed for 300 or more ALD cycles are

indicated by dashed horizontal lines. The black dotted curve is an extrapo-

lated (“guide to the eye”) behavior for ZnO in thicker layers.

021516-3 H€agglund et al.: Growth, intermixing, and surface phase formation 021516-3

JVST A - Vacuum, Surfaces, and Films

Page 6: Growth, intermixing, and surface phase formation for zinc ... · Growth, intermixing, and surface phase formation for zinc tin oxide nanolaminates produced by atomic layer deposition

consistent with the recent demonstration that ZnO initially

grows more slowly on SnOx than on ZnO, and conversely

that SnOx grows faster on ZnO than on SnOx, due to varying

densities of hydroxylated surface sites available on these

surfaces18 and possibly surface poisoning by persisting

TDMASn precursor ligands.

Although the ZnO to SnOx cycle ratio was the same for

all samples, one of the consequences of the nonlinear growth

rates is that the elemental ratios vary quite substantially as a

function of the bilayer period. This is shown by the ICP-

OES data presented in Fig. 2(a), where the relative Zn frac-

tion is plotted versus the number of terminating cycles, that

is, the number of ALD cycles of the final binary material.

Another source of variation in the overall composition is due

to the terminating layer. For instance, samples terminated by

a ZnO layer will show a significantly higher Zn fraction on

the average when containing only a few, thick bilayers.

As neither the ellipsometry nor the ICP-OES measure-

ments yield information about the spatial distribution of the

elements in these layered stacks, LEIS was employed to

analyze the surfaces and depth profiles of the ZTO films.

LEIS is an extremely surface sensitive technique, capable of

determining the elemental composition of the very outermost

atomic layer of a sample.19 It thus sensitively detects varia-

tions due to, for instance, the surface termination of the sam-

ples. All samples were cleaned using atomic oxygen for

15 min prior to analysis, to remove organic contamination

resulting from exposure to air.20 Spectra were taken using

3 keV He and 5 keV Ne scattering. The He spectra show all

elements, while the Ne spectra resolve heavier elements bet-

ter and were therefore used for quantification of surface cov-

erage here. Examples of LEIS spectra using Ne ions are

shown in Figs. 3(a) and 3(b).

The abundance of Zn and Sn on the surface can be quanti-

fied by comparison of the integrated areas of the associated

FIG. 2. (a) Bulk fraction of elemental Zn (mole fraction of Zn relative to the

summed mole fractions of Zn and Sn, in the bulk), determined by ICP-OES

in samples with varying number of terminating cycles of either ZnO (left

panel) or SnOx (right panel). (b) Surface fraction of elemental Zn (mole

fraction of Zn relative to the summed mole fractions of Zn and Sn, at the

surface) determined by LEIS for the same set of samples as in (a). In both

(a) and (b), vertical bars show the spread of the measured values.

FIG. 3. (Color online) (a) LEIS spectra using Ne scattering for samples ter-

minated with varying numbers of ZnO cycles, indicated in the legend. (b)

LEIS spectra for samples terminated with varying number of SnOx cycles.

(c) Correlation plot used to quantify the oxide surface coverage by means of

a linear fit. The red data points show the result for an uncleaned sample with

a bilayer period of 60 cycles and a 45 cycle SnOx termination, upon anneal-

ing in air for different temperatures and times. Some representative data

points are labeled. Points along the line of constant composition (red arrows)

indicate a density increase for the longest annealing times at 400 �C.

021516-4 H€agglund et al.: Growth, intermixing, and surface phase formation 021516-4

J. Vac. Sci. Technol. A, Vol. 34, No. 2, Mar/Apr 2016

Page 7: Growth, intermixing, and surface phase formation for zinc ... · Growth, intermixing, and surface phase formation for zinc tin oxide nanolaminates produced by atomic layer deposition

peaks. For each sample, two Ne spectra were taken and the

integrals determined. The average peak areas obtained in

this way are proportional to the corresponding oxide surface

coverages on the sample. To calibrate the measurement, the

average Sn peak area is plotted versus the Zn peak area for

each sample, yielding the correlation plot shown in Fig.

3(c).21 The good fit of these data to a straight line demon-

strates the absence of matrix effects, such that the detection

sensitivity for a specific atom is not affected by surrounding

elements.19 Using this fit, the relative surface fractions of the

metallic elements are then estimated by assuming that the ra-

tio of the metal atom surface densities in the pure oxides are

the same as in the corresponding crystalline oxides (rutile

SnO2 and hexagonal ZnO, respectively).

The LEIS results are summarized in Fig. 2(b), which indi-

cates a somewhat higher abundance of Sn compared to Zn on

the surface (i.e., a relative Zn fraction below 50%) for short

bilayer periods up to about 15 ALD cycles, regardless of sur-

face termination. The surface layer thus appears fairly well

mixed for supercycles including such sequences. To quantify

this observation, we define a surface mixing parameter M based

on the relative surface and bulk fractions of the metal element

in the terminating oxide (fsurf and fbulk, respectively), as

M � 1� fsurf

1� fbulk

: (1)

This mixing parameter is zero when the surface contains

exclusively the terminating element, as it would after film

closure. It is unity when the elemental surface fraction is the

same as in the bulk, as it is for a perfectly mixed system.

Based on the LEIS data for the surface fractions and the

ICP-OES data for the bulk fractions, the plots shown in Fig.

4(a) are obtained. For cases with few terminating cycles of

ZnO, the mixing parameter takes values above 1. This may

seem contradictory at first, since it implies that the surface

fraction of the terminating ZnO is actually lower than the av-

erage bulk fraction. One possible explanation is an unfavora-

ble surface termination of ZnO relative to SnO2. Indeed the

surface energies of the lowest energy planes of SnO2 (Ref.

22) and ZnO (Ref. 23) are 0.87 and 1.41 J/m2, respectively,

which should favor the driving of Zn (ZnO) to the subsurface

of SnO2. On the other hand, it is uncertain whether such

swapping would take place at the low temperatures used

here (150 �C), or if the activation energies would be too high

for this process to be relevant. Given the growth behavior

observed for ZnO on SnO2, another feasible explanation is

due to island type ZnO growth versus a layer-by-layer like

growth of the SnO2. Because ZnO island type growth will

initially leave a substantial fraction of surface Sn atoms

exposed, the relative bulk fraction of ZnO in the sample may

well exceed the relative surface fraction. For example, the

surface structure shown in Fig. 4(b) has a Zn surface cover-

age of 1/3 available for ion scattering (at near normal inci-

dence), while the average volume fraction is 1/2, making the

mixing parameter 1.33. Apart from an outlier for the case of

15 terminating cycles of SnOx, the data in Fig. 4(a) further

indicate that the mixing parameter declines much more rap-

idly with SnOx cycles terminating the surface compared to

ZnO terminated surfaces. SnOx has previously been observed

to form smooth layers with an x-ray amorphous phase,12 in

support of a conformal, layer-by-layer growth mode favoring

a rapid film closure, as illustrated in Fig. 4(c). In contrast,

ZnO is known to form crystalline grains already at low tem-

peratures.24,25 Taken together, the data suggest a ZnO island

growth mechanism combined with a layer-by-layer SnOx

growth. After repeated supercycles this leads to a film of

ZnO nanoparticles in a SnOx matrix, which is further in line

with recent observations by TEM for a related system.11 At

some point between 15 and 50 ZnO sublayer cycles, how-

ever, ZnO film closure is observed and a laminated ZnO-

SnOx structure then results. In both cases, the mixing of the

two phases occurs on the nanoscale rather than on the atomic

scale. A homogeneous ZTO phase is thus not formed by

ALD at 150 �C—at least not with the precursor system

employed here.

To investigate the effect on mixing by a postannealing

step, a sample terminated by 45 cycles of SnOx was heated

in a 200 mbar oxygen atmosphere and intermittently meas-

ured by LEIS. This sample was not cleaned, which meant

that after 15 min at 200 �C, a relatively weak Sn signal was

observed [Fig. 3(c)]. After 15 min at 300 �C, this signal

increased to the same level as the cleaned samples. At

400 �C, Zn appears on the surface, and its abundance

increases with annealing time up to about 60 min. At this

point, the surface Sn and Zn composition stabilizes while

the LEIS peak intensity increases by about 30% [Fig. 3(c)].

This peak intensity increase occurs while the ratio of the

peak areas remains constant, which can be interpreted as an

increased surface density of the oxide. Converted to ele-

mental surface fractions of Zn, the composition changes

according to Fig. 5(a). Interestingly, the composition

FIG. 4. (a) Elemental mixing as represented by the parameter M, for ZnO

and SnOx terminations, respectively. Values of zero corresponds to full film

closure of the terminating layer, while a value of unity indicates a surface

fraction equal to the average sample fraction. (b) Suggested island type

growth mode of ZnO on SnOx, leading to a mixing parameter exceeding

unity. (c) Layer-by-layer type growth of SnOx on ZnO.

021516-5 H€agglund et al.: Growth, intermixing, and surface phase formation 021516-5

JVST A - Vacuum, Surfaces, and Films

Page 8: Growth, intermixing, and surface phase formation for zinc ... · Growth, intermixing, and surface phase formation for zinc tin oxide nanolaminates produced by atomic layer deposition

saturates at an intermediate level of Sn and Zn, of close to

50%. In contrast, the average bulk fraction of Zn is much

lower in this sample [45 cycles SnOx termination in Fig.

3(a)]. This accumulation of Zn on the surface in excess of

what would be expected for a perfectly mixed sample sug-

gests that diffusion alone cannot be responsible for this. It

appears that some stable (surface) phase is formed, since

otherwise the binary oxide which minimizes the surface

energy would be driven to the surface until a high coverage

is established. A similar surface composition is observed

for the sample terminated by 150 SnOx cycles, after anneal-

ing at 500 �C for 45 min.

To further investigate the eventual presence of a surface

phase, sputtering of the latter sample was performed before

and after annealing. Before annealing, a pure and homogene-

ous SnOx layer of approximately 13 nm thickness is

observed, followed by a ZnO dominated layer, see Fig. 5(b).

After annealing, some intermixing of the two materials is

observed. Importantly, however, the sputtering profile [Fig.

5(c)] differs markedly at the outermost surface layers (sput-

ter ion fluence approaching zero). The sharp Zn gradient

observed near the surface suggests the formation of a stable

surface phase, since such a depth profile cannot result from

diffusion intermixing. Even if ZnO would minimize the sur-

face energy, in contradiction to previous observations,22,23

the surface composition would not be stable given that the

bulk composition is far from homogeneous after the per-

formed annealing [Fig. 5(c)].

The reaction between SnO2 and ZnO is exothermic. It

forms Zn2SnO4, which is known to have an inverse spinel

structure. It has been shown that the surface of powders of

spinels exposes preferentially crystallographic plane(s)

which contain the cations in the octahedral sites. For spinels,

the cations in the tetrahedral sites are not in the outer surface

detected by LEIS.26 For Zn2SnO4, this implies a structure

Znt [Sn Zn]o O4 where superscripts t and o indicate tetrahe-

dral and octahedral sites, respectively. This means that the

outer surface should contain equal amounts of Sn and Zn,

explaining the saturated levels observed here.

The high surface mobility of cations has been shown to

promote the formation of a “surface spinel.”26 The low sur-

face energy of the plane with the octahedral sites may in the

present case lead to nucleation of the inverse spinel at the

surface, while the bulk remains amorphous at these tempera-

tures (see below).

Only at much higher temperatures can the entire film be

converted to the ternary spinel phase. This is confirmed by

synchrotron-radiation GIXRD on thicker films (deposited

by 100 supercycles). Figure 6(a) shows XRD patterns of a

44 nm thick sample annealed in air at various temperatures.

The as-deposited material appears to be amorphous up to

400 �C, with a crystalline phase identified as ZnO appear-

ing in samples annealed at 600 �C. In the film annealed at

800 �C, the diffraction patterns changes to that of the spi-

nel zinc stannate Zn2SnO4 phase with some weak lines of

rutile SnO2. This indicates that an annealing temperature

exceeding 600 �C is required to form crystalline ZTO,

which is in agreement with the previous observation of

bulk crystallization of ZTO around 750 �C.4 The Sn/

(SnþZn) fraction is 0.46 as determined by ICP-OES,

while Zn2SnO4 has a Sn fraction of 0.33. This accounts for

the observed SnOx.

The effect of annealing to 800 �C was further analyzed

for films of various bilayer periods. Figure 6(b) shows the

presence of diffraction patterns of ZnO and SnO2 for films

with long bilayer periods but mostly Zn2SnO4 for shorter

bilayer periods, indicative of better sublayer mixing for

shorter bilayer periods. XRD peaks from crystalline ZnO are

only clearly observed for the two longest bilayer periods of

160 and 800 cycles, while rutile SnO2 peaks appear for

FIG. 5. (a) Elemental surface fraction of Zn (mole fraction of Zn relative to

the summed mole fractions of Zn and Sn, on the surface) for varying anneal-

ing time and temperature. (b) LEIS sputtering profile before annealing. (c)

LEIS sputtering profile after annealing.

021516-6 H€agglund et al.: Growth, intermixing, and surface phase formation 021516-6

J. Vac. Sci. Technol. A, Vol. 34, No. 2, Mar/Apr 2016

Page 9: Growth, intermixing, and surface phase formation for zinc ... · Growth, intermixing, and surface phase formation for zinc tin oxide nanolaminates produced by atomic layer deposition

bilayer periods of 8 and higher. The data may be interpreted

such that for the longest bilayer periods, the sample is com-

prised of thick layers of crystalline ZnO and SnO2 with only

thin layers of crystalline Zn2SnO4 at the sublayer interfaces.

The pattern for the shortest bilayer period of 4 cycles appears

to be “pure” crystalline Zn2SnO4 (without any significant

contribution of crystalline SnO2 or ZnO phases), suggesting

that with thinner bilayers the majority of the deposited ZnO

and SnO2 is converted into crystalline Zn2SnO4. However, it

should be noted that the films deposited using bilayer periods

of 8 and 40 cycles also contain rutile SnO2, which is likely

due an excess of Sn in these films. The film prepared using a

bilayer period of 160 is an outlier as it is more crystalline as

compared to the other films. From the width of the XRD

peaks, it can be concluded that the films consist of relatively

small grains.

The picture emerging from these results is thus that ALD

deposited ZnO/SnOx laminates do not yield ternary oxide

phases until annealed at about 400 �C or above, and then

only in the form of a surface phase in the form of the

inverted spinel Zn2SnO4. In the range of 600–800 �C this

spinel phase evolves predominantly around ZnO/SnOx inter-

faces of the laminate to eventually penetrate throughout the

bulk—when the stoichiometry is right.

IV. SUMMARY AND CONCLUSIONS

A combined analysis of data from spectroscopic ellipsom-

etry, optical emission spectroscopy, x-ray diffraction, and

low energy ion scattering was performed for ZTO thin film

laminates. It is found that ZnO follows a substrate inhibited,

overshooting growth rate behavior on SnOx, indicative of

island type growth. SnOx displays a more conformal layer-

by-layer like growth on ZnO. This results in a morphology

for short bilayer periods (<20 cycles) characterized by ZnO

inclusions (islands, clusters) in a SnOx matrix, in line with

recent TEM observations.11 For bilayer periods exceeding

approximately 20 cycles, more discrete layers are formed.

There are no clear indications of atomic scale mixing and

formation of a solid solution for ZTO stacks grown by ALD

at 150 �C; however, postannealing at 400 �C creates a stable

surface phase with equal amounts of Zn and Sn in the outer-

most atomic layer. We suggest that this corresponds to the

inverted spinel Zn2SnO4. The LEIS data indicate that the

density of the surface phase is 30% higher than the as-

deposited film. The bulk phase of zinc stannate is further

confirmed by XRD to require a much higher annealing tem-

perature (>600 �C) for its formation. This study thus pro-

vides insight into the complex characteristics of mixed metal

oxides at the nanoscale, and points toward the surface

inverted spinel as a first ternary phase developed in the ZnO/

SnOx system prior to the Zn2SnO2 bulk phase forms at

higher temperature.

ACKNOWLEDGMENTS

This work was supported by the Department of Energy

under Award No. DE-SC0004782. C.H. acknowledges

financial support from the Marcus and Amalia Wallenberg

Foundation. A.J.M.M. is grateful for financial support from

the Netherlands Organization for Scientific Research (NWO-

Rubicon 680-50-1309). The GIXRD measurements were

carried out at the SSRL, a Directorate of SLAC National

Accelerator Laboratory and an Office of Science User

Facility operated for the U.S. Department of Energy Office

of Science by Stanford University. The authors would like to

thank beam line engineers Chris Tassone and Chad Miller

for assistance during these measurements.

FIG. 6. (Color online) GIXRD patterns of ZTO films, deposited with a 1:3 ZnO:SnOx cycle ratio. (a) A 44 nm thick sample with a bilayer period of 8, annealed

at different temperatures. (b) A series of films with 800 total cycles and varying bilayer periods, annealed at 800 �C. Film thicknesses decrease with decreasing

bilayer period/increasing number of supercycles (from 97 to 43 nm) due to pronounced nucleation effects when switching from SnOx to ZnO.

021516-7 H€agglund et al.: Growth, intermixing, and surface phase formation 021516-7

JVST A - Vacuum, Surfaces, and Films

Page 10: Growth, intermixing, and surface phase formation for zinc ... · Growth, intermixing, and surface phase formation for zinc tin oxide nanolaminates produced by atomic layer deposition

1M. Tadatsugu, Semicond. Sci. Technol. 20, S35 (2005).2A. Hultqvist, M. Edoff, and T. T€orndahl, Prog. Photovoltaics Res. Appl.

19, 478 (2011).3X. Wu, Sol. Energy 77, 803 (2004).4J. Heo, S. Bok Kim, and R. G. Gordon, Appl. Phys. Lett. 101, 113507 (2012).5H. Q. Chiang, J. F. Wager, R. L. Hoffman, J. Jeong, and D. A. Keszler,

Appl. Phys. Lett. 86, 013503 (2005).6B. Sunandan and D. Joydeep, Sci. Technol. Adv. Mater. 12, 013004 (2011).7C. A. Hoel, T. O. Mason, J.-F. Gaillard, and K. R. Poeppelmeier, Chem.

Mater. 22, 3569 (2010).8T. Tommi and K. Maarit, Semicond. Sci. Technol. 29, 043001 (2014).9M. N. Mullings, C. H€agglund, J. T. Tanskanen, Y. Yee, S. Geyer, and S.

F. Bent, Thin Solid Films 556, 186 (2014).10A. Hultqvist, C. Platzer-Bj€orkman, U. Zimmermann, M. Edoff, and T.

T€orndahl, Prog. Photovoltaics Res. Appl. 20, 883 (2012).11J. Lindahl, C. H€agglund, J. T. W€atjen, M. Edoff, and T. T€orndahl, Thin

Solid Films 586, 82 (2015).12M. N. Mullings, C. H€agglund, and S. F. Bent, J. Vac. Sci. Technol., A 31,

061503 (2013).13A. Yamada, B. Sang, and M. Konagai, Appl. Surf. Sci. 112, 216 (1997).14J. W. Elam, D. A. Baker, A. J. Hryn, A. B. F. Martinson, M. J. Pellin, and

J. T. Hupp, J. Vac. Sci. Technol., A 26, 244 (2008).

15V. Lujala, J. Skarp, M. Tammenmaa, and T. Suntola, Appl. Surf. Sci.

82–83, 34 (1994).16J. T. Tanskanen, J. R. Bakke, T. A. Pakkanen, and S. F. Bent, J. Vac. Sci.

Technol., A 29, 031507 (2011).17R. L. Puurunen and W. Vandervorst, J. Appl. Phys. 96, 7686 (2004).18J. T. Tanskanen, C. H€agglund, and S. F. Bent, Chem. Mater. 26, 2795

(2014).19H. H. Brongersma, M. Draxler, M. de Ridder, and P. Bauer, Surf. Sci.

Rep. 62, 63 (2007).20H. H. Brongersma, Characterization of Materials (Wiley, 2002).21P. A. J. Ackermans, G. C. R. Krutzen, and H. H. Brongersma, Nucl.

Instrum. Methods Phys. Res., Sect. B 45, 384 (1990).22C. Sun, A. Du, G. Liu, S. Qiao, G. Lu, and S. C. Smith, Solid State

Commun. 150, 957 (2010).23N. L. Marana, V. M. Longo, E. Longo, J. B. L. Martins, and J. R.

Sambrano, J. Phys. Chem. A 112, 8958 (2008).24N. Y. Yuan, S. Y. Wang, C. B. Tan, X. Q. Wang, G. G. Chen, and J. N.

Ding, J. Cryst. Growth 366, 43 (2013).25J. Malm, E. Sahramo, J. Per€al€a, T. Sajavaara, and M. Karppinen, Thin

Solid Films 519, 5319 (2011).26J. P. Jacobs, A. Maltha, J. G. H. Reintjes, J. Drimal, V. Ponec, and H. H.

Brongersma, J. Catal. 147, 294 (1994).

021516-8 H€agglund et al.: Growth, intermixing, and surface phase formation 021516-8

J. Vac. Sci. Technol. A, Vol. 34, No. 2, Mar/Apr 2016