You are on page 1of 16

Applied Surface Science 250 (2005) 252–267

www.elsevier.com/locate/apsusc

Surface structure and composition of flat titanium thin films


as a function of film thickness and evaporation rate
Kaiyong Cai a, Michael Müller a, Jörg Bossert a,
Annett Rechtenbach b, Klaus D. Jandt a,*
a
Institute of Materials Science and Technology, Friedrich-Schiller-Universität Jena, Löbdergraben 32, Jena 07743, Germany
b
Department for Materials Technology, University of Applied Sciences Jena, Carl-Zeiss-Promenade 2, Jena 07745, Germany
Received 28 September 2004; received in revised form 5 January 2005; accepted 6 January 2005
Available online 16 February 2005

Abstract

To correlate flat titanium film surface properties with deposition parameters, titanium flat thin films were systematically
deposited on glass substrates with various thicknesses and evaporation rates by electron-beam evaporation. The chemical
compositions, crystal structure, surface topographies as well as wettability were investigated by using X-ray photoelectron
spectroscopy (XPS), X-ray diffraction (XRD), atomic force microscopy (AFM) and water contact angle measurement,
respectively. The films consisted mainly of TiO2. Small percentages of Ti2O3 and metallic Ti were also found at the film
surface using high-resolution XPS analysis. Quantitative XPS showed little differences regarding elemental compositions
among different groups of films. The films were obtained by varying the deposition rate and the film thickness, respectively.
XRD data showed consistent reflection patterns of the different titanium samples deposited using different film thicknesses.
Without exception measurements of all samples exhibited contact angles of 808  58. Quantitative AFM characterization
demonstrated good correlation tendency between surface roughness and film thickness or evaporation rate, respectively. It is
important to notice that titanium films with different sizes of grains on their surfaces but having the same chemistry and film bulk
structure can be obtained in a controllable way. By increasing the film thickness and evaporation rate, the surface roughness
increased. The surface morphology and grain size growth displayed a corresponding trend. Therefore, the control of these
parameters allows us to prepare titanium films with desired surface properties in a controllable and reproducible way for further
biological investigations of these materials.
# 2005 Elsevier B.V. All rights reserved.

PACS: 81.15; 68.35.B; 61.16.C; 79.60

Keywords: Titanium thin film; Surface characterization; XPS; AFM; XRD; Contact angle; Nanostructure

1. Introduction
* Corresponding author. Tel.: +49 3641 947730;
fax: +49 3641 947732. Titanium is a successful biocompatible material.
E-mail address: k.jandt@uni-jena.de (K.D. Jandt). It has been extensively used in dental, orthopedic

0169-4332/$ – see front matter # 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.apsusc.2005.01.013
K. Cai et al. / Applied Surface Science 250 (2005) 252–267 253

and cardiovascular fields for manufacturing bone- including the adsorbed amount and conformation
anchoring devices, such as dental implants or hip-joint [18–20]. Therefore, it is important to investigate the
replacement devices as well as heart-valves and chemical compositions and structures of the titanium
cardiac pacemakers [1,2]. For example, commercially thin films. XPS was performed in this study to
pure titanium (cp Ti) has been successfully used in determine the surface chemical composition, the
dental prosthetic applications [3,4]. Its bulk and surface oxidation state of titanium in its oxide layers and the
properties have advantages compared with other metals elemental distribution as functions of depth. AFM
used in implants: in particular, a relatively low modulus can be used to measure large areas ranging from
of elasticity, a high strength and fatigue-resistance, approximately 1 nm  1 nm to 250 mm  250 mm
excellent resistance to corrosion and superior biological simultaneously quantifying and statistically analysing
performance. topographic features ranging in height from 0.1 nm to
Although titanium implants have been successfully 10 mm [21].
used in clinical applications, the mechanism of how The aim of the present study is to investigate the
adsorbed proteins and host tissue interact with the surface properties (chemical composition, surface
titanium surface on a micro/nano scale still needs to be topography as well as microstructure) as functions of
further investigated. With the advent of atomic force film thickness and evaporation rate. The aim is to
microscopy (AFM), direct observation of individual develop a correlation between the surface properties
protein molecules with nanometer resolution has been and the parameters of the film deposition. This will be
achieved [5–9] on biomaterials. Due to the size of a useful for investigations of the potential correlation of
single protein molecule of a few 10 nm, AFM can only the surface properties and the biological reactions
be performed to image proteins on relatively smooth (protein adsorption and cell growth) in the near future.
surfaces in order to resolve the molecular protein The production and characterization of the flat Ti
conformation. To this end, relatively flat model films will be of future importance in the preparation of
surfaces are required for these investigations. Flat self-assembled monolayers using titanium rather than
titanium films can be prepared via several techniques, gold substrates.
such as sputtering [10–12], chemical vapor deposition
[13], template stripping [8] as well as anodic growth
method [14]. In this study, the flat titanium thin films 2. Materials and methods
were prepared by the electron-beam evaporation
method. 2.1. Sample preparation
The biological performance of implants strongly
depends on the first interaction occurring when Titanium was purchased from Mateck-Material-
implant surfaces come into contact with a biological Technologie & Kristalle GmbH, Germany, with purity
environment [15]. Surface modification is a usual of 99.99%. BOROFLOAT1 B33 glass slides were
way to control such interaction by modifying the purchased from Jena 4 H Engineering GmbH, Jena,
surface properties, such as the chemistry, surface Germany. Titanium thin films were evaporated on
charge and the topography of the material surface. freshly cleaned glass slides by electron-beam eva-
Flat titanium films are good substrates to create porator (Univex 350, Leybold, Germany) with various
a nanostructure to investigate the effects of topo- evaporation rates and film thicknesses (see Table 1) at
graphy on protein adsorption at the nanometre a vacuum of 8  106 mbar. Film thickness was
scale. Galli et al. [16,17] created nanostructural monitored by a quartz balance installed inside the
titanium surfaces with sizes comparable to protein evaporation chamber, which was adjusted to an
dimensions by local anodic oxidation using atomic accuracy of 5 nm. The evaporation rate was kept
force microscopy without changing its chemical constant and controlled by a deposition monitor with an
composition, and then investigated protein adsorption error range of 0.01 nm/s. The glass was cleaned with a
behaviours. seven steps procedure. Briefly, glass substrates were
It is well known that surface physical properties have immersed in 1% (v/v) Blanchipon (Hans Rosemann,
great influence on the protein adsorption behaviour Berlin) solution for 5 min; rinsed three times with
254 K. Cai et al. / Applied Surface Science 250 (2005) 252–267

Table 1
Quantitative XPS surface composition of the titanium films prepared using different deposition parameters (n = 4, mean  S.D.)
Group Evaporation rate (nm/s) Thickness (nm) Composition
Ti (at.%) O (at.%) C (at.%)
1 0.05 100 35.0  1.2 51.9  2.3 13.1  1.9
2 0.1 100 35.8  2.2 49.4  2.4 14.8  2.3
3 0.5 100 33.3  1.7 50.4  2.8 16.4  2.9
4 0.05 200 30.4  0.6 52.5  2.5 17.1  2.8
5 0.1 200 33.0  1.1 53.6  2.2 13.4  1.5
6 0.5 200 32.0  1.1 54.1  3.0 14.0  2.4
7 0.05 500 31.1  1.7 50.2  2.5 18.7  2.9
8 0.1 500 30.6  0.8 51.7  1.7 17.7  2.1
9 0.5 500 31.9  1.2 54.8  3.0 13.3  3.2

distilled water; then the samples were transferred to 5% Intensity ratios were converted into atomic concen-
(v/v) Optical II (Mineralöle Albert, Jena) and washed tration ratios by using the sensitivity factors proposed
3 min in an ultrasonic bath. Subsequently, the samples by the manufacturer. For quantitative chemical
were immersed into high purity ethanol and distilled composition analysis, four samples in each group
water for 10 min, respectively; followed by immersing were acquired and the survey scans were performed in
samples in 35–36% (v/v) HNO3 for another 30 s; then two different areas on each sample.
these cleaned samples were intensively washed with
double distilled water and stored until use. Glass 2.3. XRD
substrates were dried with nitrogen immediately before
being mounted for film deposition. A Siemens D5000 (Brucker AXS GmbH, Karls-
ruhe, Germany) diffractometer was employed to
2.2. XPS determine the structures of titanium films in this
study. The diffractometer was operated in the u–2u
For chemical composition analysis, specimens arrangement, where u is the angle between the incident
were characterized using XPS (Quantum 2000, PHI X-ray and sample surface normal and the diffraction
Co., Chanhassen, MN, USA) with a focused mono- angle; 2u, is the angle between the incoming X-ray
chromatic Al Ka source (1486.6 eV) for excitation beam and the diffracted beam. XRD measurements
and a 100-mm diameter beam. The electron take-off were carried out with a detector scan mode using Cu
angle was 458 and the analyser was operated in the Ka radiation. The incident angle was 18 and the
constant energy mode for all measurements. XPS effective penetration depth of the X-ray beam was
survey spectra over a binding energy range of 0– approximately 430 nm. Large-angle range survey
1000 eV were acquired using analyzer pass energy of spectra with 2u value ranging from 228 to 858 were
58.7 eV and 23.5 eV for high-resolution elemental recorded with step of 0.0158 and a step time of
scan. Argon ion sputtering was performed during 10 s/point. Diffrac-AT software was used to process
XPS to estimate the thickness and structure of the the results and subsequently compare them to the
surface layers. The vacuum pressure was around Powder Diffraction FileTM release 2002 (PDF-2)
7.5  108 mbar during spectral acquisition. Spectra database.
were peak fitted after background subtraction by
assuming a Gauss/Lorentz (90–70/10–30) peak shape 2.4. Contact angle
except for the asymmetric metallic Ti peak. Data
analysis was carried out with Multipak software The contact angle with respect to double-distilled
provided by the manufacturer. The hydrocarbon peak water was measured with a DSA 10 drop shape
maximum in the C 1s spectra was set as 284.8 eV to analysis system (Krüss GmbH, Hamburg, Germany).
reference the binding energy scales for the samples. The images were recorded by a camera, and then
K. Cai et al. / Applied Surface Science 250 (2005) 252–267 255

analyzed using the software supplied by the manu- spectrum were attributed to TiO2, corresponding with
facturer. Four samples were measured in each group. Ti 2p3/2 at a binding energy (BE) of 458.0 eV and Ti
Three different points were measured for each sample. 2p1/2 at BE of 463.6 eV, being a typical value for TiO2
Every drop was deposited to the sample surface and [22–24]. The binding energy difference of 5.6 eV
after 15, an image of the drop was captured and then between these two peaks for TiO2 correlated well with
analyzed. All measurements were performed at previous studies (DBE from 5.5 eV to 5.8 eV) [24,25–
ambient temperature. 27]. The lowest energy peak was assigned to metallic
Ti with an observed BE for the Ti 2p3/2 of 453.1 eV. In
2.5. AFM addition, Ti3+ (Ti 2p3/2 at BE of 456.3 eV) from
substoichiometric oxide, Ti2O3, was present, as also
The AFM characterization was performed using a reported in previous XPS studies [25,27,28].
Dimension 3100 AFM (Digital Instruments, Veeco, The O 1s peak (Fig. 1b) shows three components:
CA) in tapping mode at ambient temperature. The the first, at 531.0 eV, was attributed to the titanium
topographic and phase images were recorded simul- oxide; the second peak at 532.3 eV was fitted and
taneously with a standard silicon tip on a cantilever assigned to hydroxide and hydroxyl species; the third
beam. The spring constant of the cantilever was oxygen peak at BE of 533.2 eV was associated with
between 20 and 100 N/m and the length was 125 mm. surface contamination, usually water and carbonates
Root mean square (RMS) surface roughness value was [22,24,25].
determined on an area of 50 mm  50 mm per sample. XPS survey spectrum demonstrates the carbon
RMS data were used for quantitative analysis. contamination on the titanium surface, which is
Roughness parameters RMS were calculated using typical for adventitious, unavoidable hydrocarbon
the roughness analysis software provided by the contamination, adsorbing spontaneously from ambi-
manufacturer of the instrument. ent air onto the surface [29–31]. From the high-
resolution spectrum of C 1s (Fig. 1c), three peaks were
fitted and assigned as follows: the main carbon peak at
3. Results and discussion 284.8 eV was assigned to hydrocarbon species (C–H);
the other two fitted peaks attributed to C–O (i.e.
3.1. XPS characterization alcohol) species at 286.1 eVand C O (i.e. carbonates)
species at 288.5 eV, respectively.
XPS is a sensitive surface analysis technique, Argon ion etching XPS depth profiles perpendi-
which is capable of providing both qualitative and cular to the sample surfaces show a carbon
quantitative information about the presence of contamination overlayer on the surface. This carbon
different elements at the surface. Table 1 lists the contaminant was removed after the first several
surface chemical compositions of the titanium films etching cycles (Fig. 2). About 45 at.% oxygen was
deposited with different evaporation rates and film present on the film surface, which came from the
thicknesses. In addition to the expected titanium and native oxide layer on the titanium film. The oxygen
oxygen, which were present in the titanium dioxide, concentration decreased dramatically after the initial
carbon was also observed on all samples. The presence etching cycles. At greater depth of about 8–10 nm,
of carbon originated from surface contamination since titanium and oxygen concentrations remained almost
the samples were exposed to air after e-beam constant at about 95 at.% and 5 at.%, respectively,
processing. Small differences in chemical composi- until the silicon substrate interface was reached. As for
tion on Ti, O or C were observed among different the remaining oxygen (approximately 5 at.%) inside
groups. This indicates that deposition parameters in the titanium film, there were three factors that might
this study have relatively little influence on the contribute to this. The first was the environmental
chemical composition of these films. oxygen presenting in the chamber used for titanium
The original high-resolution XPS-spectrum of the sample preparation. Titanium is highly reactive with
Ti 2p peak (Fig. 1a) shows three peaks. After peak- oxygen. During titanium deposition, titanium atoms
fitting treatment, the dominant doublet peaks in the evaporated from the titanium source may have reacted
256 K. Cai et al. / Applied Surface Science 250 (2005) 252–267

Fig. 1. Typical high-resolution XPS-spectra of titanium sample binding state: (a) Ti 2P; (b) O 1s; and (c) C 1s (film thickness: 100 nm;
evaporation rate: 0.1 nm/s).

with oxygen atoms in the chamber and titanium– that the carbon concentration dropped dramatically
oxygen compounds may have been captured within from ca. 11 at.% down to zero at the initial etching
the titanium films. Since the vacuum was around cycles.
8  106 mbar, it is not surprising to find some oxygen
in the chamber. The second factor was the margin error 3.2. XRD
associated with XPS quantitative analysis, which is
about 5–10% of the measured value [32]. The third Fig. 3 shows the XRD patterns of titanium films
factor might be related to the well-known ‘‘knock-on’’ deposited at a fixed evaporation rate of 0.5 nm/s, but
effect [33] of ion etching, pushing the surface oxygen with different film thicknesses of 100 nm, 200 nm and
atoms deeper into the film at the initial sputtering stage. 500 nm, respectively. It displayed the (0 0 2), (1 0 2)
Simultaneously with the reduction in titanium con- and (1 0 3) reflection peak around 2u = 388, 538, and
centration, the oxygen level rose and a silicon signal 70.68, respectively, which were characteristic peaks of
appeared in the spectrum when the titanium–glass a-titanium [31,32]. The crystal structure of titanium is
substrate interface reached. It was important to note hexagonal with a0 = 0.295 nm and c0 = 0.4686 nm
K. Cai et al. / Applied Surface Science 250 (2005) 252–267 257

[34]. The (1 0 3) peak is the typical reflection in


hexagonal a-titanium. No evidence for TiO2, a
common titanium oxide on the surface of titanium
was found in the XRD data. The native TiO2 layer is
typically about 5 nm [27], which was too thin to yield
sufficient signals detectable by XRD technique used in
this study.
The distinct XRD peaks mentioned above were
only found on samples with a film thickness of 200 nm
and 500 nm, respectively. A small peak contributed to
(0 0 2) and no peak attributed to (1 0 2) reflections
were observed on the titanium film with a thickness of
100 nm. The main reason is that the titanium sample
with a film thickness of 100 nm did not yield sufficient
XRD signals to identify these peaks. However, (1 0 3)
reflection pattern was clearly identified in all titanium
films with different film thicknesses. Different relative
Fig. 2. XPS depth profile of titanium sample surface (film thickness: change behaviours of reflection patterns [(0 0 2),
100 nm; evaporation rate: 0.1 nm/s). (1 0 2) and (1 0 3)] indicated the texture existed in
titanium films presented in this study.

3.3. Contact angle measurement

Fig. 4 shows the contact angle data determined


using double-distilled water of titanium films prepared
with various film thicknesses and evaporation rates.
Within the margin of error there was no difference
of contact angle except for the titanium sample,
which was deposited with parameters of 500 nm film
thickness and evaporation rate of 0.5 nm/s. It had the

Fig. 3. XRD patterns of the titanium films (fixed evaporation


rate = 0.5 nm/s) as a function of film thicknesses: (a) 500 nm; (b) Fig. 4. The contact angles of titanium films prepared using different
200 nm; and (c) 100 nm. evaporation rates and film thicknesses (n = 4, mean  S.D.).
258 K. Cai et al. / Applied Surface Science 250 (2005) 252–267

lowest contact angle (77.38  2.48) among all at a minimum, then the overall change of surface free
titanium samples. energy must vanish and this gives Young’s law for the
Contact angle measurements are important to equilibrium contact angle (Eq. (1)). Roughness of
characterize the wetting proprieties of biomedical surface modifies this argument because the rough
materials. Previous studies [35,36] revealed that the surface has a larger area f DA than horizontal projec-
hydrophilic and/or hydrophobic properties, which tion of the area [39], where f is the roughness factor
were determined by contact angle measurements, of and is greater than 1. Where perfect smoothness is an
biomaterials are probably closely related to protein acceptable assumption, as at liquid–liquid or liquid–
adsorption behaviour of the biomaterials. The contact gas interfaces, actual surface and geometric surface
angle of a material was determined by the surface are identical, but at the surface of any real solid the
chemistry (elements and functional groups) and actual surface will be greater than the geometric sur-
surface roughness of the material surface [37,38]. It face because of surface roughness. The surface ratio
has been reported that totally hydrophobic materials here will be termed the ‘‘roughness factor’’, and
can be converted to hydrophilic ones just by rough- designed by f [41]:
ening their surface morphologies [39]. In the present
study, XPS characterization was performed on the actual surface
f ¼ roughness factor ¼ (2)
titanium films deposited with different parameters geometric surface
(film thickness and evaporation rate). Within the
While the surface roughness alters the surface free
margin of error of the XPS measurements, the samples
energy at the solid interface, it does not change the
showed small differences in the chemical composition
liquid–vapor contribution.
of the film surfaces. Differences in contact angle, if
The overall effect of topography is to result in an
any, may therefore originate from differences in
equilibrium contact angle on a rough surface, ure , given
surface roughness of the sample.
by Wenzel’s equation [41]:
The wetting behaviour of a surface is mainly
determined by intrinsic factors of its surface chemistry cos ure ¼ f cos se (3)
and extrinsic factors of surface roughness. The
hydrophobicity and wetting behaviour of a liquid on Wenzel’s equation can be used to predict the influence
a solid surface depends on the balance of forces at the of roughness on the contact angle. If the contact angle
contact line arsing from the three interfacial tensions, on a smooth surface is lower than 908, roughness
gSL, gSV and gLV, occurring at the solid–liquid, solid– (rougher surface) will reduce the observed contact
vapor, and liquid–vapor interfaces, respectively. angle, but if it is higher than 908, roughness will
The relationship of the forces, respectively, surface/ further increase the observed contact angle. The flat
interfacial tensions at the three phase points can be titanium thin films presented contact angles around
described by Young’s equation: 808 in the present study, lower than 908. Thus, it was
g SV  g SL not surprising to find that the roughest titanium film
cos use ¼ (1) (confirmed by AFM measurement) was having the
g LV
lowest contact angle of 77.38  2.48. However, since
where use is the equilibrium contact angle. Another the surface roughness difference among various tita-
view of balance forces is to consider the interfacial nium films was not so much, much difference could
tensions as surface energies per unit area of the inter- not be expected among different groups.
face. In this approach, the effect of increasing the area However, the data showed that high film thickness
covered by contact line by a small amount, DA, is to and high evaporation rate had little influence on the
replace the solid–vapor interface by a solid–liquid contact angles of titanium films. The contact angle
interface, thus changing the surface free energy by values in the present study (around 808) were higher
(gSL  gSV)DA. In addition, the movement of the than those of previous studies [42–44]. The major
liquid creates additional liquid surface area of reason for this was the larger surface roughness of
DA cos u, resulting in a surface free energy increase samples used in those studies. Since the differences in
of gLV DA cos u [40]. If the surface free energy is to be contact angles between the samples in this study were
K. Cai et al. / Applied Surface Science 250 (2005) 252–267 259

relatively small, major differences in surface rough- the present study, the effects of film thickness and
ness between the samples were not expected. To evaporation rate on surface roughness were investi-
confirm this, AFM measurements of the titanium gated in a matrix (3  3) mode, i.e. using three
samples were carried out. different film thicknesses and three various evapora-
tion rates. All surfaces exhibited grain features.
3.4. AFM characterization Fig. 6a–c, d–f and g–i showed the influence of film
thickness in each group on film surface morphology at
While XPS allows an overview of surface fixed evaporation rates of 0.05 nm/s, 0.1 nm/s and
chemistry as well as element distribution along depth 0.5 nm/s, respectively. The morphology was relatively
of titanium films prepared with different parameters uniform for films with 100 nm film thickness as shown
(film thickness and evaporation rate) on a comparable in Fig. 6a, d and g compared to films with thicknesses
big area, AFM has its strength on a microscopic view of 200 nm and 500 nm (where some larger grains can
of titanium surface on a qualitative and quantitative be seen) in each group. The results suggested that
basis. Surface morphology, morphological evolution surface morphology evolutions of these films were
and grain growth behaviours of titanium films were dependent on their film thickness. This finding is in
resolved in the present study. agreement with the results from a previous study [32].
The surface morphologies of bare glass substrate With increasing film thickness, some small voids were
and the titanium thin films had been investigated in present and enlarge.
this study with AFM. Fig. 5 presents bare glass As shown in Fig. 6, the line profiles of titanium
substrate and a representative group of titanium film films displayed a trend: with increasing film thickness
surface morphologies as a function of evaporation rate and/or evaporation rate, the surface became rougher.
with the fixed film thickness of 200 nm. The studied The line profiles of titanium films provided information
surfaces are equal to 1 mm  1 mm in area. about the surface roughness along the lines at the
Bare glass substrate displayed quite smooth section. To obtain more information about the mean
surface morphology with little defects compared with grain size, particle analysis was performed in this study.
those of titanium films. Its RMS roughness was Fig. 7 shows mean grain sizes (n = 4, at least 50
0.8 nm  0.2 nm in a 50 mm  50 mm measuring area. measurements were performed) of titanium film
Samples deposited with the lowest evaporation rate of surfaces prepared with various film thicknesses and
0.05 nm/s (Fig. 5a), showed small spherical grains with evaporation rates. The grains varied in a broad range
an average diameter of 29.3 nm  2.9 nm (n = 50), from 16 nm to 108 nm depending on their deposition
more or less coalescent and easily distinguishable, parameters. It is interesting to note that in the present
forming a relatively smooth surface. At a higher study titanium films with different sizes of grains at
evaporation rate of 0.1 nm/s, the increased diameter their surfaces but having the same chemistry and film
(38.4 nm  4.3 nm; n = 50) of spherical grains lead to a bulk structure can be obtained in a controllable way.
slightly rougher surface morphology (Fig. 6b). Increas- With increasing film thickness and evaporation rate,
ing the evaporation rate to 0.5 nm/s (Fig. 5c), the strong the mean diameters of grains increased (see Fig. 7). It
coalescence of neighbouring grains resulted in larger was well known that both surface chemistry [45–47]
grains with an average diameter of 64.1 nm  7.3 nm and surface topography [48–50] would affect protein
(n = 50) and rough surface morphology. The RMS adsorption and cell adhesion behaviours. Since most
surface roughness measured over 50 mm  50 mm topographical variations were accompanied with
area correspondingly increased from 2.2 nm  0.2 nm chemical heterogeneities [51], dividing topographical
nm (0.05 nm/s), to 2.6 nm  0.2 nm (0.1 nm/s) and from chemical effects was quite difficult in most cases.
4.9 nm  0.2 nm (0.5 nm/s). This indicated that the The titanium films in the present study may be
evaporation rate had a distinct effect on the surface regarded as a potentially important model surface for
morphology. such kinds of investigations, especially on a nano-
Fig. 6 shows the evolution of surface morphology scopic scale. Galli et al. [16] demonstrated that
accompanying line profiles of titanium films as a proteins could ‘‘sense’’ the topography of surfaces at
function of film thicknesses and evaporation rates. In the nanometer scale.
260 K. Cai et al. / Applied Surface Science 250 (2005) 252–267

Fig. 5. AFM images of the glass substrate and titanium films (thickness = 200 nm) deposited with different evaporation rates: (a) substrate; (b)
0.05 nm/s; (c) 0.1 nm/s; and (d) 0.5 nm/s.

Fig. 8 shows quantitative AFM data of titanium each group was observed: with increasing film
film RMS surface roughness measured in 50 mm  thickness surface roughness increased. It showed
50 mm areas as functions of evaporation rate and film good correlation with the variations of grain sizes for
thickness, respectively. In three groups with different titanium films as shown Fig. 7. This result reflected
fixed evaporation rates (Fig. 8a), a similar effect that the grain size variations caused the surface
of film thickness on film surface roughness in roughness change on titanium films prepared with
K. Cai et al. / Applied Surface Science 250 (2005) 252–267 261

Fig. 6. AFM images (1 mm  1 mm) and accompanying line profiles of titanium films deposited with different evaporation rates and film
thicknesses (z-axis range: 30 nm). (a) Evaporation rate: 0.05 nm/s; film thickness: 100 nm; (b) evaporation rate: 0.05 nm/s; film thickness:
200 nm; (c) evaporation rate: 0.05 nm/s; film thickness: 500 nm; (d) evaporation rate: 0.1 nm/s; film thickness: 100 nm; (e) evaporation rate:
0.1 nm/s; film thickness: 200 nm; (f) evaporation rate: 0.1 nm/s; film thickness: 500 nm; (g) evaporation rate: 0.5 nm/s; film thickness: 100 nm;
(h) evaporation rate: 0.5 nm/s; film thickness: 200 nm; (i) evaporation rate: 0.5 nm/s; film thickness: 500 nm.
262 K. Cai et al. / Applied Surface Science 250 (2005) 252–267

Fig. 6. (Continued ).

different parameters (film thickness and evaporation when the evaporation rate was increased in each
rate). Fig. 8b shows the influence of the evaporation group.
rate on surface roughness when the film thickness Upon coalescence of the surface nuclei to form a
is fixed. It showed an increased surface roughness continuous film, the nucleation process of film
K. Cai et al. / Applied Surface Science 250 (2005) 252–267 263

Fig. 6. (Continued ).

deposition is complete, and the step for development the ratio of the substrate temperature Ts to the melting
of bulk film structure starts. The film structure depends point of the film Tm, Th = Ts/Tm (in K), known as
among other factors on the amount of thermal motion ‘‘homogeneous’’ or ‘‘reduced’’ temperature (Th).
taking place during film formation, which scales with Three so-called structural zones (Z1, Z2 and Z3)
264 K. Cai et al. / Applied Surface Science 250 (2005) 252–267

Fig. 7. Mean grain sizes of titanium films prepared using different


evaporation rates and film thicknesses.

plus a transition zone were identified in the


evaporative deposition study that included both metals
and ceramics [52], and those zones have been
observed in a wide variety of film materials deposited
by all of the vapor-phase processes. In the first zone
(Z1), which is occurring at Ts/Tm < 0.3, the grain
growth is controlled by the low atom mobility and the
grain size increases with the growth temperature. The
second zone (Z2), which occurs at Ts/Tm > 0.3,
displays crystallites formed by higher surface mobi-
lity. The crystallite size increases with increasing
growth temperature. A transition to Z3 occurs in
certain instances at Ts/Tm > 0.5, which is high enough
Fig. 8. The relationship between surface roughness and (a) film
for considerable bulk annealing of the film to take
thickness; (b) evaporation rate of samples used in this study (n = 4,
place during deposition. Z3 is characterized by more mean  S.D.).
isotropic or equiaxial structure, and again, the grain
size increases with growth temperature. The zone ZT (evaporation rate is 0.5 nm/s) increased from 100 nm
is found between Z1 and Z2 where small grains are not to 200 nm and 500 nm, some small voids were present
well defined which is usually associated with energy- and enlarge. Their RMS surface roughnesses corre-
enhanced processes. It is clear that the temperature is spondingly increased from 3.5  0.3 nm, to 4.9 
the main determinant of film structure. 0.2 nm and 6.7  0.2 nm, which were consistent with
In the present study, titanium films were deposited the previous research results from Messier and Yehoda
at room temperature (around 22 8C). The melting [54]. The temperature was not kept constant during
point for titanium is 1660 8C. Thus, the ratio of Ts/Tm film deposition in this study. It changed little in the
(0.153) was so low that surface diffusion was chamber during the e-beam evaporation process.
negligible, belonging to Z1 structure [53]. Z1 consists However, according to the zone model mentioned
of columns typically tens of nm in diameter separated above, slight temperature changes, even several ten
by voids of a few nm across. These characteristics can centigrades, will not lead to transition from one zone
be seen in Fig. 6i in this study. In Z1 growth, voids are to another. Since in the present study the homogeneous
a key feature to considerably increase the rate of temperature (Th) was only 0.153, no change in zone
roughening. As shown in Fig. 7, when film thickness transition is expected. It also means slight temperature
K. Cai et al. / Applied Surface Science 250 (2005) 252–267 265

changes will not result in changes of titanium film known columnar growth structure [55]. When the
structure and topography. depositing atoms obliquely reaching the substrate
A Z1 structure generally consists of columns surface, self-shadowing increases and results in an
having poor crystallinity (many defects) or is increased formation of Z1 structure. As an example in
amorphous. In thicker films, there are some wider this study, one group with fixed 500 nm film thickness
voids between the columns expected [53]. Our result, but with various evaporation rates was chosen to
as shown in Fig. 6i clearly demonstrated the void investigate how evaporation rate affected its film
formation (see arrows) as a function of film thickness. topography. With the lowest evaporation rate of
The cones terminate in domes at the surface, and the 0.05 nm/s (Fig. 6g), the coalescence of free titanium
size of the domes increases with film thickness [53]. atoms is not so strong on the surface, forming a
Concerning the observation regarding grain size smoother surface with a homogenous sphere structure.
growth as function of film thickness (Fig. 7), our An increased evaporation rate leads to higher vapor
results also presented a strong analogy with the zone pressure, thus a higher concentration of titanium
model as mentioned above. A previous study revealed atoms. In this case, the nucleation and fast grain
the microstructure of titanium films via transmission growth dominated the surface diffusion of titanium
electron microscopy (TEM). Those titanium films atoms, excessive grain growth then occurred, resulting
were prepared using the electron-beam evaporation in the development of a coarse grain sized morphology
method at ambient temperature, which was the same (Fig. 6i). Thus, a higher surface roughness could be
condition as the present study. TEM results showed a expected.
columnar structure within the titanium films [55].
It is well known that most metal adatoms are highly
mobile at room temperature. The immobility of the 4. Conclusions
surfaces results from the fact that the detachment of
metal surface atoms from steps and kinks is kinetically Flat thin titanium films were prepared by electron-
hindered at these temperatures [56]. It is well known beam evaporation on a borofloat glass substrate. The
that when the depositing titanium atoms reach the surface properties (chemical composition, surface
substrate surface, two destabilizing factors, namely roughness and hydrophobicity/hydrophilicity) as
statistical roughening and self-shadowing are always functions of film thickness and evaporation rate were
at work. These two factors will affect the micro- investigated.
structure and morphology development of the films. Qualitative and quantitative XPS characterizations
Statistical roughening arises from the statistical were performed on titanium films deposited by
fluctuation in the vapor arrival flux. A model has different parameters. Depth profile XPS displayed
been developed to describe this phenomenon [53], the distribution trend for Ti, O as well as C along the
which demonstrated that the roughness would be film layer. Carbon contaminant was found only
increased as the film thickness increases. Wang and located on the top layer of titanium films. Quantitative
Zhang [55] confirmed that surface roughness of films XPS data showed that there was no difference of their
increased with layer thickness. The results of the chemical composition within error range of XPS
present study were consistent with those findings as measurements among groups with different deposition
shown in Fig. 7a. parameters.
Self-shadowing is the second factor that destabi- XRD results showed that there was no difference in
lizes surface smoothness during film deposition, and it crystal structure among different titanium films except
is the cause of the characteristic voided columnar for the 100 nm film. The intensity difference among
structure. It has been suggested [53] that depositing different samples results from the XRD signals from
atoms generally can not perch on top of each other, but various titanium films with different film thickness.
rather settle sideways into the nearest ‘‘cradle’’ Contact angle data showed that there were little
position in which they establish relaxed bond lengths differences among samples deposited with different
to their nearest neighbours. This is known as ‘‘ballistic parameters. Since there were small differences in the
aggregation’’, leading to a rough surface with well- chemical composition, any differences in the contact
266 K. Cai et al. / Applied Surface Science 250 (2005) 252–267

angle should relate to variance in their surface [12] D. Mardare, C. Baban, R. Gavrila, M. Modreanu, G.I. Rusu,
Surf. Sci. 507 (2002) 468.
roughness.
[13] K.H. Ahn, Y.B. Park, D.W. Park, Surf. Coat. Tech. 171 (2003)
AFM images showed the relationship between 198.
surface morphology and deposition parameters. [14] E. Vasilescu, P. Drob, M.V. Popa, M. Anghel, A.S. Lopez, I.M.
Quantitative AFM analysis showed a linear relation Rosca, Mater. Corros. 51 (2000) 413.
trend between surface roughness and evaporation rate. [15] D.G. Caster, B.D. Ratner, Surf. Sci. 500 (2002) 28.
As film thickness increased, the film roughness [16] C. Galli, M.C. Coen, R. Hauert, V.L. Katanaev, P. Gröning, L.
Schlapbach, Coll. Surf. B 26 (2002) 255.
increased although no strict linear correlation existed. [17] C. Galli, M.C. Coen, R. Hauert, V.L. Katanaev, M.P. Wymann,
The grain growth within the film layer depended on P. Gröning, L. Schlapbach, Surf. Sci. 474 (2001) L180.
the film deposition conditions. The films consisted of [18] J.R. Capadona, D.M. Collard, A.J. Garcia, Langmuir 19 (2003)
larger grains correlated with a higher RMS roughness. 1847.
[19] A. Krajewski, A. Piancastelli, R. Malavolti, Biomaterials 19
(1998) 637.
[20] F.A. Denis, P. Hanarp, D.S. Sutherland, J. Gold, C. Mustin,
P.G. Rouxhet, Y.F. Dufrene, Langmuir 18 (2002) 819.
Acknowledgements [21] T.P. Niesen, M.R. De Guire, J. Bill, F. Aldinger, M. Ruhle,
A. Fischer, F.C. Jentoft, R. Schlogl, J. Mater. Res. 14 (1999)
We gratefully acknowledge the financial support 2464.
from the Thüringer Ministerium für Wissenschaft, [22] F. Pinzart, P. Ascarelli, E. Cappelli, R. Giorgi, Langmuir 18
Forschung und Kunst (TMWFK) through the Projekt (2002) 5457.
[23] J. Lausma, B. Kasemo, Appl. Surf. Sci. 44 (1990) 133.
‘‘Thüringer Schwerpunkt Grenzflächentechnologien’’ [24] I. Bertoti, M. Mohai, J.L. Sullivan, S.O. Saied, Appl. Surf. Sci.
(B478-02001) and the partial financial support of the 84 (1995) 357.
BMBF for this work within the project ‘‘Innovations [25] C. Viornery, Y. Chevolot, D. Leonard, B.O. Aronsson, P.
und Gründerlabor für neue Werkstoffe (Biomateria- Pechy, H.J. Mathieu, P. Descouts, M. Gratzel, Langmuir 18
lien) und Verfahren (IGWV) an der Friedrich-Schiller- (2002) 2582.
[26] F. Zhang, Z. Zheng, Y. Chen, X. Liu, A. Chen, Z. Jiang, J.
Universität Jena’’ Förderkennzeichen: 03GL0026. Biomed. Mater. Res. 42 (1998) 128.
[27] C. Sittig, M. Textor, N.D. Spencer, M. Wieland, P.H. Vallotton,
J. Mater. Sci. Mater. Med. 10 (1999) 35.
[28] M. Jobin, M. Taborelli, P. Descouts, J. Appl. Phys. 77 (1995)
References 5149.
[29] S.J. Xiao, M. Textor, N.D. Spencer, M. Wieland, B. Keller, H.
[1] S.A. Brown, J.E. Lemons, Medical Applications of Titanium Sigrist, J. Mater. Sci. Mater. Med. 8 (1997) 867.
and its alloys: The Material and Biological Issues, American [30] M. Winkelmann, J. Gold, R. Hauert, B. Kasemo, N.D. Spencer,
Society for Testing and Materials, Philadelphia, PA, 1996, pp. D.M. Brunette, M. Textor, Biomaterials 24 (2003) 1133.
3–18. [31] H.K. Jang, S.W. Whangbo, H.B. Kim, J. Vac. Sci. Technol. A
[2] M. Long, H.J. Rack, Biomaterials 19 (1998) 1621. 18 (2002) 917.
[3] C. Massaro, P. Rotolo, F. De Riccardis, Milella, A. Napoli, M. [32] C.C. Ting, S.Y. Chen, D.M. Liu, J. Appl. Phys. 88 (2000)
Wieland, M. Textor, N.D. Spencer, D.M.E. Brunette, J. Mater. 4628.
Sci. Mater. Med. 13 (2002) 535. [33] P. Cacciafesta, K.R. Hallam, C.A. Oyedepo, A.D.L. Humphris,
[4] S. Poulin, E. Ndzbet, A. Debrie, O. Savadogo, E. Sacher, Appl. M.J. Miles, K.D. Jandt, Chem. Mater. 14 (2002) 777.
Surf. Sci. 143 (1999) 238. [34] W.B. Pearson, A Handbook of Lattice Spacings and Structures
[5] P. Cacciafesta, A.D.L. Humphris, K.D. Jandt, M.J. Miles, of Metals and Alloys, Pergamon Press, London, UK, 1994.
Langmuir 16 (2000) 8167. [35] F.Q. Nie, Z.K. Xu, P. Ye, J. Wu, P. Seta, Polymer 45 (2004) 399.
[6] K.D. Jandt, Surf. Sci. 491 (2001) 303. [36] A. Azioune, M.M. Chehimi, B. Miksa, T. Basinska, S. Slom-
[7] N.B. Holland, R.E. Marchant, J. Biomed. Mater. Res. 51 kowski, Langmuir 18 (2002) 1150.
(2000) 307. [37] M.C.L. Martins, B.D. Ratner, M.A. Barbosa, J. Biomed. Mater.
[8] R.E. Marchant, M.D. Barb, J.R. Shainoff, S.J. Eppell, D.L. Res. 67A (2003) 158.
Wilson, C.A. Siedlecki, Thromb Haemostasis 77 (1997) 1048. [38] C.S. Zhao, X.D. Liu, M. Nomizu, N. Nishi, Biomaterials 24
[9] P. Cacciafesta, K.R. Hallem, A.C. Watkinson, G.C. Allen, M.J. (2003) 3747.
Miles, K.D. Jandt, Surf. Sci. 491 (2001) 405. [39] J.F. Oliver, C. Huh, S.G. Mason, Coll. Surf. 1 (1980) 79.
[10] K. Okimura, T. Nakamura, J. Vac. Sci. Technol. A 21 (2003) [40] G. McHale, N.J. Shirtcliffe, S. Aqil, C.C. Perry, M.I. Newton,
988. Phys. Rev. Lett. 93 (2004) 036102.
[11] L. Sirghi, Y. Hatanaka, Surf. Sci. 530 (2003) L323. [41] R.N. Wenzel, Ind. Eng. Chem. 28 (1936) 988.
K. Cai et al. / Applied Surface Science 250 (2005) 252–267 267

[42] D.E. MacDonald, N. Deo, B. Markovic, M. Stranick, P. [49] A. Higushi, S. Tamiya, T. Tsubomura, A. Katoh, C.S. Akaike,
Somasundaran, Biomaterials 23 (2002) 1269. M. Hara, J. Biomater. Sci. Polym. E 11 (2000) 149.
[43] Q. Liu, J. Ding, F.K. Mante, S.L. Wunder, G.R. Baran, [50] Y.D. Dufrene, T.G. Marchat, P.G. Rouxhet, Langmuir 15
Biomaterials 23 (2002) 3103. (1999) 2871.
[44] D.E. MacDonald, B. Markovic, M. Allen, P. Somasundaran, [51] A. Curts, C. Wilkinson, Biomaterials 18 (1997) 1573.
A.L. Boskey, J. Biomed. Mater. Res. 41 (1998) 120. [52] B.A. Movchan, A.V. Demchishin, Fiz. Metal. Metalloved 28
[45] G.B. Sigal, M. Mrksich, G.M. Whitesides, J. Am. Chem. Soc. (1969) 653.
210 (1998) 3464. [53] D.L. Smith, Thin-Film Deposition: Principle and Practice,
[46] S. Kidoaki, T. Matsuda, Langmuir 15 (1999) 7639. McGraw-Hill, New York, 1995 , p. 159.
[47] K.B. MaClary, T. Ugarova, D.W. Grainger, J. Biomed. Mater. [54] R. Messier, J.E. Yehoda, J. Appl. Phys. 58 (1985) 3739.
Res. 50 (2000) 428. [55] Y.L. Wang, K.Y. Zhang, Surf. Coat. Tech. 140 (2001) 155.
[48] B.S. Wojciak, A. Curis, W. Monhagan, K. Macdonald, C. [56] G. Palasantzas, J.T.M. de Hosson, Physica C 330 (2000)
Wilkinson, Cell Res. 223 (1996) 426. 99.

You might also like