You are on page 1of 17

J Mater Sci (2019) 54:13724–13739

Materials
M A T E R I A L S Ffor
O R Llife
I F E Ssciences
CIENCES

Anodization growth of TiO2 nanotubes on Ti–35Nb–


7Zr–5Ta alloy: effects of anodization time, strain
hardening, and crystallographic texture
Leonardo Fanton1,* , Alessandra Cremasco1 , Mariana G. Mello1 , and Rubens Caram1

1
University of Campinas (UNICAMP), Rua Mendeleyev, 200, Campinas, SP 13083-860, Brazil

Received: 17 June 2019 ABSTRACT


Accepted: 22 July 2019 Titanium and its alloys are the most suitable metallic materials available for the
Published online: fabrication of medical implants. Their biocompatibility can be improved by the
31 July 2019 growth of TiO2 nanotubes on their surface by a simple anodization process. This
work involved an investigation into the anodization behavior of Ti–35Nb–7Zr–
Ó Springer Science+Business 5Ta (TNZT) alloy, focusing on the effect of processing conditions (anodization
Media, LLC, part of Springer time and type of electrolyte), previous strain hardening, and crystallographic
Nature 2019 texture of the substrate. Studies about the growth of TiO2 nanotubes on b-type
titanium alloys, as the TNZT alloy, are rare in the literature. The TNZT alloy
proved to be an excellent substrate for the growth of TiO2 nanotubes, resulting
in threefold longer nanotubes than those obtained on a commercially pure (CP)
Ti substrate. Moreover, TiO2 nanowires grew after 6 h of anodization in an
organic electrolyte, which could not be achieved using the CP-Ti substrate.
Samples with different crystallographic textures displayed similar nanotube
morphology and only slight differences in grain length, indicating that grain
orientation played only a minor role in the growth kinetics. Lastly, the crys-
tallization of nanotubes at 450 °C did not alter their morphology, but caused
complete detachment of the TiO2 nanotubes at 700 °C.

Introduction temperature, pure titanium (Ti) has a hexagonal


close-packed (hcp) structure (a-phase). When heated
Titanium and its alloys have enhanced corrosion above 882 °C, it transforms into body-centered cubic
resistance and high specific strength. Albeit relatively (bcc) allotropic form (b-phase) [1]. The b-phase can be
expensive, they are highly attractive for the aerospace retained at room temperature in a stable or
sector, where weight reduction is decisive, and for metastable form by adding b-stabilizing alloying
the fabrication of medical implants, due to their good elements such as Mo, Nb, Zr, and Ta. Ti alloys based
biocompatibility and corrosion resistance. At room on the b-phase stand out for their relatively low

Address correspondence to E-mail: leofanton@gmail.com

https://doi.org/10.1007/s10853-019-03870-5

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci (2019) 54:13724–13739 13725

elastic modulus, which is important to reduce the aqueous and organic electrolytes is given in a review
stress-shielding effect [2], and their enhanced bio- by Regonini et al. [17].
corrosion resistance [3]. Ti–35Nb–7Zr–5Ta (TNZT) TiO2 nanotubes are formed during anodization due
alloy is a commercially available metastable b-Ti to a combination of simultaneous processes. Metallic
alloy designed for use in the manufacture of implants Ti dissolves continuously, forming Ti4? ions that
with an elastic modulus of 60 GPa [3, 4] and a yield react with the O2- ions present in water to form an
strength of 530 MPa (solution-treated condition) [3]. oxide (TiO2) layer. The oxide layer continues to grow
Despite persistent efforts to render metallic by means of field-assisted ion transport (O2- and Ti4?
implants more compatible with the human body, ions) through the growing oxide [7]. Oxide growth is
implant rejection is still an unresolved issue. Implant self-limited because the electric field is reduced as the
surface properties play a major role in the implant’s thickness of the oxide layer increases. The use of a
biocompatibility; thus, it is understandable that fluoride-containing electrolyte is necessary because it
numerous surface modifications have been proposed dissolves the TiO2 layer, creating a porous structure.
to address this issue. Rougher surfaces enhance the The competition between the dissolution of oxide,
growth and attachment of bone cells. Therefore, which occurs preferentially at the pore/tube base,
several treatments to increase the surface roughness and its field-assisted growth enables the synthesis of
have been investigated, such as anodic oxidation, TiO2 nanotubes. It has also been reported [18] that
grit-blasting, calcium phosphate coatings, plasma TiO2 nanowires (sometimes called nanofibers) can be
spraying, chemical modifications [5], laser surface formed on top of nanotubes. Although this phe-
melting [6], and others. nomenon is not yet fully understood, Lim and Choi
Highly ordered nanotube or nanopore oxide [19] proposed the ‘‘bamboo-splitting’’ model,
structures can be grown on the surface of some whereby nanotubes are vertically split by the longi-
metals such as Al, Ti, Zr, Nb, W, Ta, and Hf [7]. In the tudinal flow of ions in the channel of nanotubes.
case of Ti, the growth of titanium dioxide (titania, Nanowires have several morphological characteris-
TiO2) nanotube arrays is attracting the attention of tics that are advantageous for biological applications,
researchers due to their many potential applications, favoring cell activities. They are highly porous, have
as in gas sensors, solar cells, photocatalysis, biological high surface-to-volume ratios, and are similar to the
applications, etc. [8]. The presence of TiO2 nanotubes natural extracellular matrix (hydroxyapatite and
on the surface of Ti alloys can increase biocompati- collagen fibers). TiO2 nanotubes/nanowires formed
bility and reduce inflammatory processes [9, 10]. The during anodization are generally amorphous,
most common and efficient way to grow TiO2 nan- although they can be crystalline under certain
otubes is by electrochemical synthesis (anodization), anodizing conditions [7]. Amorphous TiO2 nanotubes
in which an electric potential is applied between two can be crystallized by means of a subsequent heat
electrodes separated by a fluoride-containing elec- treatment, usually from 300 to 700 °C, and different
trolyte. This is a simple process that can be applied allotropic phases can be formed, the most common
even to complex implant geometries [11]. The work- one being anatase and rutile [20]. As for implant
ing electrode (anode) corresponds to the Ti substrate applications, several studies [21] indicate that the
on which the TiO2 nanotubes will grow, and the anatase form is more effective for the formation of
counter electrode (cathode) is usually made of Pt, apatite, which is important for the osseointegration
although different materials have been used, includ- process. Anatase is also better for photocatalytic
ing stainless steel [12], Ni [13], and others [14]. Two applications [22].
main types of electrolytes are used to synthesize TiO2 It is well known that nanotube morphology
nanotubes, i.e., aqueous and organic. Aqueous elec- depends on several factors, such as the electrolyte,
trolytes were the first ones to be studied, and they the voltage applied to the system, alloy composition,
generally consist of hydrofluoric acid (HF) [15] or temperature, and anodization time [23–26]. In fact,
fluoride salts [16] dissolved in water. Many organic the condition of the substrate can also affect the
media can be used as organic electrolytes, including growth of TiO2 nanotubes. Several studies have
glycerol and ethylene glycol, with the addition of shown that mechanically polished or electropolished
fluoride salts, commonly NH4F, and small amounts surfaces generate more organized nanotubes, since a
of water. A detailed description of the evolution of smoother surface ensures a uniform electric field

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


13726 J Mater Sci (2019) 54:13724–13739

distribution over the substrate surface [27, 28]. Sub- Experimental procedure
strate grain orientation may also influence TiO2
nanotube growth, according to Leonardi et al. [29], The Ti–35Nb–7Zr–5Ta (TNZT) alloy was melted in an
who anodized a CP-Ti sample using an organic electric arc furnace under a high-purity (99.999%)
electrolyte solution composed of glycerol and 0.25 argon atmosphere. The resulting ingot was encapsu-
wt% NH4F and applied a voltage of 20 V to the sys- lated in a quartz glass tube filled with argon gas,
tem for 6 h. They found that the strong influence of homogenized in a resistance furnace at 950 °C for
grain orientation along the length of the nanotubes 24 h, and water-quenched. The sample was then
was inversely proportional to the surface atom den- cold-rolled (room temperature) using multiple passes
sity. Crawford and Chawla [30] reached a similar to prevent adiabatic heating, reducing its thickness
conclusion. Hence, in principle, the crystallographic by up to 75%. It was then cut in two parts: One of
texture (preferred grain orientation) of the substrate them was kept in the deformed (strain-hardened)
may play an important role in the overall morphol- state, while the other was recrystallized for 30 min at
ogy of TiO2 nanotubes. 800 °C in a furnace in an argon atmosphere and
This study evaluates the electrochemical anodiza- water-quenched. These samples are hereinafter
tion growth of self-organized TiO2 nanotubes on a referred to as ‘‘strain-hardened’’ and ‘‘recrystallized’’
biomedical TNZT alloy, highlighting the effects of samples.
anodization time, strain hardening, and crystallo- The composition of alloying elements in the
graphic texture, using aqueous and organic elec- deformed and recrystallized substrates was exam-
trolytes. Although TNZT alloy is a commercially ined by X-ray fluorescence spectroscopy (XRF)
available b-type Ti alloy designed for the fabrication
of medical implants, no reports were found on its use
as a substrate for growing TiO2 nanotubes.

Figure 1 Schematic diagram of the experimental apparatus used Figure 2 XRD patterns (2h scan) and micrographs (VLM) of:
for anodization. a strain-hardened and b recrystallized samples.

Table 1 Chemical
composition of the TNZT Ti Nb Zr Ta N (9 10-3) O (9 10-3)
samples (in wt%)
54.0 ± 0.1 33.8 ± 0.1 7.0 ± 0.1 5.2 ± 0.1 5.99 ± 0.35 182 ± 5

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci (2019) 54:13724–13739 13727

Figure 3 FESEM micrographs showing the top view of the TiO2 TNZT, and i–l CP-Ti substrates. The side view (length) of the
nanotubes after 4 min, 8 min, 30 min, and 1 h of anodization in an nanotubes is depicted in the upper right-hand corner of each
aqueous HF (0.3% vol.) electrolyte at a maximum voltage of 20 V, image.
grown on the a–d strain-hardened TNZT, e–h recrystallized

(Shimadzu EDX 7000 spectrometer), while nitrogen


and oxygen gas contents were measured in a LECO
TC400 nitrogen/oxygen analyzer. The samples were
prepared for evaluation by visible-light microscopy
(VLM) by sanding them with SiC abrasive paper up
to 1500 grit, polishing them with diamond paste
(6 lm) and colloidal silica (0.05 lm), and then etching
them with Kroll reagent (6 ml HNO3, 3 ml HF, and
91 ml H2O) to reveal the grain boundaries. Phase
identification was carried out by X-ray diffraction
(XRD) (PANalytical X’Pert Pro).
The crystallographic texture of the strain-hardened
and recrystallized samples was determined by XRD
pole figure measurements (PANalytical X’Pert Pro).
To obtain a quantifiable description of the crystallo-
graphic texture, the orientation distribution functions
(ODFs) were calculated from the experimental pole
Figure 4 Side view of TiO2 nanotubes grown on a strain- figures, using MTEX MATLABÒ toolbox software.
hardened TNZT substrate, after 8 min of anodization in the The method used by MTEX to calculate the ODFs is
aqueous electrolyte.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


13728 J Mater Sci (2019) 54:13724–13739

explained by Hielscher and Schaeben [31]. According Results and discussion


to common practice for rolled samples, the axes of the
external reference frame were defined as the rolling Substrate composition and microstructure
direction (RD), normal direction (ND), and transverse
direction (TD). Table 1 describes the measured chemical composi-
TiO2 nanotubes were grown by anodization on the tion of the TNZT samples. The content of alloying
rolling surface (orthogonal to ND) of the strain- elements Ti, Nb, Zr, and Ta is consistent with the
hardened and recrystallized TNZT samples. For nominal composition, and the levels of oxygen and
comparison, samples of commercially pure (CP) Ti nitrogen contamination can be considered low,
(99.6% pure, according to the supplier) were similarly according to standard specifications for Ti alloy
anodized. Before anodization, the samples were castings (ASTM B367-13-2017).
sanded with abrasive paper (up to 1500 grit) and The XRD patterns of the strain-hardened (Fig. 2a)
chemically polished by immersing them for 10 s in a and recrystallized (Fig. 2b) TNZT samples presented
solution containing equal proportions of nitric acid
and hydrofluoric acid. Chemical polishing was
employed to remove the surface layer deformed by
sanding. An electrolytic cell made of polypropylene
with interior dimensions of 50 mm diameter and
50 mm height was used. The sample to be treated
was placed in contact with the electrolyte through a
round side window, 8 mm in diameter, located at the
midpoint of the cell. A Pt foil with dimensions of
about 40 9 40 mm2 immersed in the electrolyte was
used as the cathode. The process was carried out
using two different electrolytes: (1) an aqueous
solution with 0.3% (vol.%) HF and (2) an organic
(ethylene glycol) solution with 0.5% NH4F and 10%
deionized water (vol.%). The electrolyte was stirred
continuously during the process. Direct current (DC)
was applied to the system by a power supply (BK
Precision 9201), first increasing the voltage from 0 to
20 V at a rate of 2 V min-1 and then keeping this
voltage constant until the end of the process. Samples
were prepared with different anodization times,
depending on the electrolyte employed. When the
aqueous electrolyte was used, the process was stop-
ped after 4 min, 8 min, 30 min, and 1 h, and in the
case of the organic electrolyte, the anodization times
were 8 min, 30 min, 1 h, and 6 h, always counting
from the beginning of the initial ramp. The anodiza-
tion process is schematized in Fig. 1. The morphology
of the nanotubes was examined by field emission
scanning electron microscopy (FESEM, FEI Quanta
650). Nanotube length was measured by scratching
the anodized surface with a scalpel, causing the
nanotubes to break and occasionally fall on their side.
Figure 5 a Top view and b side view of the nanotubes grown on
recrystallized TNZT substrate after 1 h of anodization in the
aqueous electrolyte, showing two groups of nanotubes with
different diameters and lengths.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci (2019) 54:13724–13739 13729

only reflections of the cubic b-phase, as was expected hardened and recrystallized TNZT substrates pre-
for this alloy. The VLM micrograph of the strain- sented similar length, diameter, and general appear-
hardened sample (inset in Fig. 2a) shows a typical ance, confirming that the strain-hardened condition
microstructure obtained after cold-rolling, with did not significantly affect nanotube growth. The
highly elongated grains, while the recrystallized nanotubes grown on TNZT alloy were much longer,
sample (inset in Fig. 2b) shows well-defined equiax- reaching approximately 1500 nm after 1 h of
ial grains, indicating complete recrystallization. The anodization, compared to a length of only 490 nm on
strain-hardened sample showed a measured hard- the CP-Ti substrate. The diameter of nanotubes
ness of 246 ± 6 HV, which decreased to 198 ± 6 HV grown on the CP-Ti substrate was larger at the
after recrystallization. beginning of anodization (4 and 8 min), but became
similar on the two alloys after longer anodization
TiO2 nanotube morphology times (30 min and 1 h), reaching about 90 nm. After
30 min and 60 min of anodization, the morphology of
Figure 3 depicts the general morphology of TiO2 the nanotubes was similar on all the alloys, indicating
nanotubes grown in aqueous electrolyte after 4 min, that the maximum length and diameter had already
8 min, 30 min, and 1 h of anodization on strain- been attained in the first 30 min of anodization.
hardened (a–d) and recrystallized (e–h) TNZT sub- TiO2 nanotubes in the early stages of growth of up
strates, and on the CP-Ti substrate (i–l) used as ref- to 8 min of anodization appear to be less organized,
erence. A comparison of equal anodization times and their diameter is about three times smaller than
indicates that the nanotubes grown on strain- the nanotubes obtained after longer anodization

Figure 6 FESEM micrographs showing the top view of TiO2 on the a–d strain-hardened TNZT, e–h recrystallized TNZT, and
nanotubes after 8 min, 30 min, 1 h, and 6 h of anodization in an i–l CP-Ti substrates. The side view (length) of the nanotubes is
organic electrolyte (ethylene glycol ? 0.5 vol.% NH4F ? 10 depicted in the upper right-hand corner of each image.
vol.% deionized water) at a maximum voltage of 20 V, grown

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


13730 J Mater Sci (2019) 54:13724–13739

times. To exemplify, the nanotubes produced after nanotubes with larger diameters (about 85 nm),
8 min of anodization on strain-hardened TNZT sub- spaced apart from each other, surrounded by a sec-
strate (Fig. 3) have an apparent diameter of only ond group of nanotubes whose diameter is approxi-
about 22 nm at their top opening. However, a side mately half that of the first group. The narrower
view of the nanotubes, as in the example shown in nanotubes also appear to be slightly shorter, which
Fig. 4, reveals that their base is thicker, with a can be confirmed by a side view (Fig. 5b). TiO2 nan-
diameter of approximately 65 nm. A nanoporous otubes ordered on two different scales have already
layer of TiO2 is initially formed at the beginning of been reported by Ozkan et al. [32, 33], but with a very
the anodization process. After some time, nanotubes different morphology. Their SEM images show that
begin to grow and their diameter increases as the the longer nanotubes of the first group were about
process continues. This initial porous structure is still 2 lm in length, while the shorter nanotubes in the
present during the first 8 min of anodization as a second group were only about 0.1 lm in length and
remnant layer on top of the nanotubes, as seen in were morphologically thinner. They also found that
Fig. 4, resulting in nanotubes of poor quality. This the shorter nanotubes split after a short growth,
initial nanopore/nanotube layer is consumed during forming a stack of small nanotubes. The size of this
the process, and the diameter at the top of nanotubes stack of nanotubes was very limited, since it was
is considerably larger after longer anodization times. easily dissolved by etching after some time during
After 30 min and 1 h of anodization, the TiO2 anodization. However, none of the characteristics
nanotubes grown on TNZT substrates (both strain- reported by Ozkan et al. was observed in our study.
hardened and recrystallized) (Fig. 3c, d, g, h) can be Figure 6 shows the general morphology of the
divided into two groups of different sizes. The nan- nanotubes grown in the organic electrolyte after
otubes grown on the CP-Ti substrate did not exhibit 8 min, 30 min, 1 h, and 6 h of anodization on strain-
this behavior. Figure 5a clearly shows a group of hardened TNZT (a–d), recrystallized TNZT (e–h),

Figure 7 Cross sections of TiO2 nanotubes from their base (near showing different perspectives of the initial growth process, and on
the substrate) to their tip (top opening), grown in the organic d–f the CP-Ti substrate, after different anodization times.
electrolyte on a–c the TNZT substrate after 30 min of anodization,

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci (2019) 54:13724–13739 13731

and CP-Ti (i–l) substrates. The morphologies of the the nanotubes grown in organic electrolyte was sig-
nanotubes grown on strain-hardened and recrystal- nificantly smaller than those of the nanotubes grown
lized TNZT substrates were similar. The maximum in aqueous electrolyte. The maximum nanotube
length of the nanotubes grown on the TNZT alloy diameter reached on the TNZT alloy was 39 nm (after
was about 5000 nm, which was reached during the 1 h), and on the CP-Ti substrate, it was about 68 nm
first hour of anodization. However, after 6 h of (after 6 h).
anodization, TiO2 nanowire structures were found to As explained in the ‘‘Experimental procedure’’
grow on top of the nanotubes. This growth, which section, the surface of TiO2 nanotubes was scratched
will be discussed later, was not considered in the with a scalpel in order to separate the nanotubes from
measurement of nanotube length. After 1 h of the substrate, enabling them to be viewed from the
anodization, the nanotubes grown on the CP-Ti side. In the case of nanotubes grown in organic
substrate reached a length of about 1200 nm, which electrolyte, this procedure also caused some nan-
more than doubled after 6 h of anodization, reaching otubes to break in several sections along their length,
about 3300 nm. However, the surface of the nan- enabling the nanotube growth process to be exam-
otubes after 6 h of anodization showed an irregular ined in greater detail. Figure 7 shows some examples
aspect, called ‘‘nanograss’’ [34], which is characteristic of nanotubes grown on the CP-Ti and TNZT sub-
of longer anodization times, leading to excessive strates. In each case, note that the nanotubes are
etching in the electrolyte. In general, the diameter of considerably thicker close to their base, which can be

Figure 8 a Length and


b diameter of nanotubes
grown in aqueous (0.3% vol.
HF) and organic (ethylene
glycol ? 0.5 vol.%
NH4F ? 10 vol.% deionized
water) electrolytes after
different anodization times at a
maximum voltage of 20 V.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


13732 J Mater Sci (2019) 54:13724–13739

explained by the fact that the upper portion of the As a reference, Table 2 shows the percentage of
nanotubes is older and was therefore etched by the grains whose {100}, {110}, and {111} planes are ori-
electrolyte solution for a longer time. ented parallel to the rolling surface, considering a
Similarly to the anodization process in aqueous misorientation tolerance angle of 15°.
electrolyte (Fig. 4), a residual porous layer was visi- Considering a0 as the lattice parameter of the bcc
ble on the nanotubes in the initial stages of anodiza- unity cell, the relative atomic densities of the {100},
tion. In the case of nanotubes grown on TNZT {110}, and {111} planes are, respectively, 1.00, 1.41,
substrate, this layer was still present after 30 min of and 0.58 atoms per a20. Given that atomic planes of
anodization (Fig. 7a, b, c), but on those grown on CP- lower density favor TiO2 nanotube growth [29, 30],
Ti substrate it was only visible after 8 min (Fig. 7d), the crystallographic texture of the recrystallized
having disappeared after longer anodization times TNZT sample should, in principle, result in a higher
(Fig. 7e, f). The porous layer took longer to consume average length of TiO2 nanotubes. However, the
in both aqueous and organic electrolytes when length of both strain-hardened and recrystallized
anodization was performed on TNZT substrate. The nanotube samples is similar. In addition, only minor
nanotubes grown on TNZT substrate appeared to be changes in length from grain to grain are visible in
more resistant to etching during anodization, which
may also explain why they were able to grow longer.
The graph shown in Fig. 8 depicts the length and
diameter of TiO2 nanotubes grown on CP-Ti and
TNZT substrates after different anodization times in
aqueous and organic electrolytes.

Effect of crystallographic texture

An XRD crystallographic texture analysis was made


of the strain-hardened and recrystallized TNZT
samples. The ODF calculated from the experimental
pole figures provides a complete description of the
texture, but the most important information to be
extracted from this result is the planes that are pref-
erentially exposed on the substrate surface, since
their density may influence the growth of TiO2 nan-
otubes to some degree [29, 30]. The most practical
way to display this information is through inverse
pole figures (IPFs). Figure 9 illustrates the IPFs of the
strain-hardened and recrystallized substrates, show- Figure 9 Inverse pole figures (IPFs) of the a strain-hardened and
ing the crystals planes that are most frequently par- b recrystallized samples, showing the crystal planes that are
allel to the rolling direction (orthogonal to the ND). preferentially parallel to the rolling surface (orthogonal to ND).
The intensity of the IPFs is indicated in multiples of The intensity of the IPFs is indicated in multiples of uniform
uniform density, i.e., in the number of times a certain density (MUD).
plane is more frequently aligned to the rolling surface
compared to a non-textured (uniform orientation Table 2 Fraction of grains with their {hkl} plane oriented parallel
density) material. The IPF of the strain-hardened to the substrate surface, considering a misorientation tolerance
sample shows a strong intensity on the {100} plane angle of 15°
(Fig. 9a), which means that the {100} plane of most of
Crystal plane Strain-hardened TNZT Recrystallized TNZT
the grains is parallel to the rolling surface. The
{hkl} (%) (%)
intensities of the IPFs of the recrystallized sample
(Fig. 9b) are significantly weaker, indicating a more {100} 52.4 13.6
varied distribution of orientations, with a maximum {110} 2.1 9.8
intensity on the {111} plane. {111} 18.0 22.6

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci (2019) 54:13724–13739 13733

the recrystallized TNZT and CP-Ti samples. After contradictory finding. Those authors used electrolytic
4 min of anodization in the aqueous electrolyte, polishing to remove the layer deformed during the
nanotube growth on the CP-Ti sample was found to preceding mechanical polishing, while we employed
be inhomogeneous. As can be seen in the low-mag- chemical polishing in our study. The sample grown
nification SEM image in Fig. 10a, some grains are on recrystallized TNZT showed slight differences in
brighter than others. A closer look at Fig. 10b, c nanotube length between grains after 8 min of
reveals that the nanotubes with brighter grains have a anodization in organic solution, as illustrated in
well-defined porous/tubular structure, while the Fig. 10e. However, this behavior was not observed
darker grains appear to indicate an earlier stage of after longer periods of anodization in either the
the anodization process. After 1 h of anodization, aqueous or the organic electrolyte solutions. No
some grains, albeit only a few, still showed slight information was found in the literature about the
differences in length (Fig. 10d). A reasonable expla- effect of grain orientation on the growth of TiO2
nation for the differences in grain length is that their nanotubes on b-Ti alloys. The degree of anisotropy of
orientation affected the TiO2 nanotube growth rate, the cubic structure of b-Ti phase is lower than that of
since this was the only characteristic that varied from the hexagonal a-phase present in CP-Ti alloy, which
one grain to another. However, these variations in in principle should reduce the influence of grain
length were not as significant as those reported by orientation on nanotube growth. Moreover, the fact
Leonardi et al. [29]. Differences in the surface that the nanotubes grown on the recrystallized sam-
preparation employed in our study and in that of ple presented the same morphology as those grown
Leonardi et al. may have been the cause of this on the strain-hardened sample is another indication

Figure 10 SEM micrographs illustrating the difference in TiO2 magnification images of the same sample shown in (a).
nanotube morphology between different grains after anodization. d Surface of nanotubes grown on the CP-Ti sample after 1 h of
a Low-magnification image of the CP-Ti sample after 4 min of anodization in aqueous solution. e Surface of nanotubes on the
anodization in aqueous solution, showing grains with (light recrystallized TNZT sample after 8 min of anodization in organic
regions) and without (dark regions) nanotubes. b, c High- solution.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


13734 J Mater Sci (2019) 54:13724–13739

Figure 11 SEM micrographs of the nanowires formed on TNZT the same sample shown in (b). d, e Side views of the nanotubes
samples after anodization in the organic electrolyte. a Onset of covered with nanowires after 6 h of anodization. f Image of a
nanowire formation after 1 h of anodization, with the sample tilted region where the nanowires were broken, providing a view of the
at an angle of 45° to facilitate viewing. b Top-view image of underlying nanotubes.
nanowires after 6 h of anodization. c High-magnification image of

that grain orientation does not play a significant role nanowires is difficult because of their flexibility and
in the anodization process. hair-like appearance. However, it is reasonable to
assume that they are more than 5 lm long, and
TiO2 nanowires longer than the underlying nanotubes. Figure 11f
shows a region where the nanowires were acciden-
As already shown in Fig. 6, TiO2 nanowires were tally broken, providing a clear view of the nanotubes
found to grow on nanotubes during the anodization underneath them.
of TNZT substrates in the organic electrolyte solu- It has been already reported [19] that nanowires
tion. The formation of nanowires began after 1 h of can be grown through the anodization of CP-Ti in
anodization. Albeit not clearly visible in the top view ethylene glycol ? NH4F electrolytes. However, they
images (Fig. 6g, k), nanowire clusters were detected were only formed upon applying voltages above
upon tilting the sample at an angle of 45° (Fig. 11a). 80 V. The formation of TiO2 nanowires on TNZT
After 6 h of anodization, the nanowires were much substrates using a voltage of only 20 V indicates that
longer, well developed, and still grouped in clusters this alloy is an excellent substrate for nanowire
(Fig. 11b). Individual nanowires were visible under growth.
higher magnification (Fig. 11c), evidently growing
from the borders of the nanotubes underneath them. Crystallization of TiO2 nanotubes
This finding is consistent with the ‘‘bamboo-splitting’’
model proposed by Lim and Choi [19] to explain the Heat treatments were applied at 450 and 700 °C for
formation of TiO2 nanowires. A side view of the TiO2 1 h to evaluate the crystallization behavior of TiO2
nanostructures (Fig. 11d, e) reveals the nanotube and nanostructures. The samples chosen for this analysis
nanowire sizes. Measuring the actual length of were the ones anodized for 6 h in organic electrolyte,

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci (2019) 54:13724–13739 13735

Figure 13 XRD analysis of nanotubes grown on a CP-Ti


Figure 12 XRD analysis of nanotubes grown on a TNZT
substrate after 1 h of anodization in organic solution (ethylene
substrate after 1 h of anodization in organic solution, a as-
glycol ? 0.5 vol.% NH4F ? 10 vol.% deionized water). a As-
anodized and heat-treated at b 450 °C and c 700 °C for 1 h. The
anodized, b heat-treated at 450 °C for 1 h, and c heat-treated at
insets in (b) and (c) are SEM (backscatter detector) cross-sectional
700 °C for 1 h. The inset in (c) is a visible-light microscopy image
images of the region near the surface.
(real colors) of the nanotube surface.
which was the condition that produced the longer
Fig. 12c). This result was similar to that of the ano-
nanotubes. Figure 12 shows the XRD patterns of TiO2
dized specimens; hence, the TiO2 nanotube layer was
nanotubes/nanowires grown on the TNZT sample.
not related to a-phase precipitation. The cross sec-
The as-anodized nanotubes did not present peaks
tions of the substrates were investigated by SEM.
corresponding to any crystalline phase other than the
While the sample treated at 450 °C showed only
substrate (b-Ti), indicating their amorphous nature
equiaxed b-phase grains (inset in Fig. 12b), the sam-
(Fig. 12a). After a heat treatment at 450 °C, several
ple treated at 700 °C presented a-precipitates close to
anatase peaks and one rutile peak were detected
the surface, which was revealed by their darker shade
(Fig. 12b). When treated at 700 °C, the XRD patterns
(inset in Fig. 12c). An XRD analysis of the cross sec-
showed anatase and rutile peaks similar to those of
tion of the surface of the sample treated at 700 °C
the sample heat-treated at 450 °C. However, peaks of
(third XRD pattern in Fig. 12c) showed only b-peaks,
metallic a-Ti phase were also observed, which was
which is explained by the fact that a-precipitates are
not expected. To determine the source of the a-phase,
present only in a very thin layer near the surface. The
non-anodized TNZT samples were also heat-treated
precipitation of a-phase close to the surface of b-Ti
at 450 and 700 °C for 1 h and their surfaces were
alloys after heat treatments is a common phe-
analyzed by XRD. The sample treated at 450 °C
nomenon [35]. This precipitation is attributed to the
showed only b-peaks (second XRD pattern in
high affinity of Ti to absorb oxygen from the atmo-
Fig. 12b), while the sample treated at 700 °C showed
sphere, which is an element known to stabilize the a-
peaks of a- and b-phases (second XRD pattern in

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


13736 J Mater Sci (2019) 54:13724–13739

Figure 14 SEM micrographs


of TiO2 nanotubes/nanowires
grown on the CP-Ti and TNZT
substrates, after 6 h of
anodization in organic
electrolyte and heat-treated at
a, b 450 °C and c, d 700 °C
for 1 h. The insets in (b), (c),
and (d) show high-
magnification images of the
regions marked in a black
rectangle.

phase. Although the heat treatments were performed observed before the treatments (Fig. 6d, h). However,
in a protective argon atmosphere, the relatively high when heat-treated at 700 °C, the TiO2 nanotubes
temperature (700 °C) may have facilitated the grown on the CP-Ti sample cracked (Fig. 14c) and
absorption of some residual oxygen. became detached from the substrate, as can be seen
Figure 13 shows XRD patterns of nanotubes grown by the elevation of the nanotubes near the cracks
on the CP-Ti substrate. As expected, only peaks of the (inset in Fig. 14c). This effect was even worse on the
substrate (a-Ti) were visible before the heat treatment TNZT substrate, where the nanotubes became com-
(Fig. 13a). After the heat treatment at 450 °C, crys- pletely detached from the substrate, forming small
tallization occurred and anatase peaks became visible ‘‘blocks’’ of nanotubes that were visible all over the
(Fig. 13b), but no rutile phase was detected. When surface (Fig. 14d). No nanotubes remained on the
treated at 700 °C, three titanium oxide allotropes substrate surface, indicating that the nanotubes broke
were detected: anatase, rutile, and Ti2O3 (Fig. 13c). at their base, as illustrated by several detached blocks
The anatase peaks were more intense than those of that were turned upside down (inset in Fig. 14d).
the sample treated at 450 °C, indicating that the Losertová et al. [37] also reported the collapse of TiO2
amount of this phase was higher. The nanotube and nanotubes after heat treatments, showing that it is an
substrate surfaces acquired a strong blue coloration important issue to be taken into consideration when
(see inset in Fig. 13c), which is the characteristic color choosing the crystallization treatment temperature.
of Ti2O3 phase [36].
The morphology of the TiO2 nanostructures was
examined after applying crystallization heat treat- Conclusions
ments. After the heat treatment at 450 °C, the nan-
otubes grown on the CP-Ti substrate (Fig. 14a) and (a) The TNZT substrate proved to be suitable for
the nanotubes/nanowires grown on the TNZT sub- growing TiO2 nanotubes, enabling the
strate (Fig. 14b) showed a similar appearance to those

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci (2019) 54:13724–13739 13737

formation of threefold longer nanotubes than References


those grown on the CP-Ti substrate.
(b) The nanotubes grown on the recrystallized and [1] Lutjering G, Willians JC (2007) Titanium (engineering
strain-hardened TNZT substrates showed no materials and processes). Springer, Berlin
significant morphological difference. Moreover, [2] Ridzwan MIZ, Shuib S, Hassan AY, Shokri AA, Mohammad
the crystallographic texture also did not play an Ibrahim MN (2007) Problem of stress shielding and
important role. improvement to the hip implant designs: a review. J Med Sci
(c) TiO2 nanotubes grown on the TNZT samples 7:460–467
presented two groups of different scales, with [3] Qazi JI, Rack HJ (2005) Metastable beta titanium alloys for
characteristics that have never before been orthopedic applications. Adv Eng Mater 7:993–998. https://d
reported in the literature. oi.org/10.1002/adem.200500060
(d) Upon anodizing the TNZT sample in organic [4] Kopova I, Stráský J, Harcuba P, Landa M, Janeček M,
electrolyte for 6 h, long nanowires grew on top Bačákova L (2016) Newly developed Ti–Nb–Zr–Ta–Si–Fe
of the nanotubes, in conditions that would not biomedical beta titanium alloys with increased strength and
be possible using a CP-Ti substrate. enhanced biocompatibility. Mater Sci Eng C 60:230–238. h
(e) Crystallization of the nanotubes grown on the ttps://doi.org/10.1016/j.msec.2015.11.043
TNZT substrate was successful after 1 h of heat [5] Patil PS, Bhongade ML (2016) Dental implant surface
treatment at 450 °C, forming anatase and rutile modifications: a review. J Dent Med Sci 15:132–141. http
phases with no noticeable change in the nan- s://doi.org/10.9790/0853-151003132141
otube morphology. However, crystallization at [6] Balla VK, Soderlind J, Bose S, Bandyopadhyay A (2014)
700 °C for 1 h caused detachment of the Microstructure, mechanical and wear properties of laser
nanotubes from the substrate and a-phase surface melted Ti6Al4 V alloy. J Mech Behav Biomed Mater
precipitation. 32:335–344. https://doi.org/10.1016/j.jmbbm.2013.12.001
[7] Macak JM, Tsuchiya H, Ghicov A, Yasuda K, Hahn R,
Bauer S, Schmuki P (2007) TiO2 nanotubes: self-organized
electrochemical formation, properties and applications. Curr
Acknowledgements Opin Solid State Mater Sci 11:3–18. https://doi.org/10.1016/
j.cossms.2007.08.004
The authors gratefully acknowledge LNNano/
[8] Lee M, Kim T, Bae C, Shin H, Kim J (2010) Fabrication and
CNPEM (National Nanotechnology Laboratory of
applications of metal-oxide nanotubes. JOM 62:44–49
the National Center for Research on Energy and
[9] Duvvuru MK, Han W, Chowdhury PR, Vahabzadeh S, Sci-
Materials) for providing access to its SEM facilities.
ammarella F, Elsawa SF (2019) Bone marrow stromal cells
This work was supported by the Brazilian research
interaction with titanium; Effects of composition and surface
funding agency FAPESP (São Paulo Research Foun-
modification. PLoS ONE 14:e0216087. https://doi.org/10.1
dation) [Grant #2017/16715-0].
371/journal.pone.0216087
[10] Barjaktarević DR, Djokić VR, Damnjanović ID, Rakin MP
(2018) Nanotubular oxide layers formed on the Ti-based
Authors’ contributions
implants surfaces-application and possible damages: a
LF is the main author, performed the experiments, review. Metall Mater Eng. https://doi.org/10.30544/401
and took the lead in writing the paper. MGM and AC [11] Li T, Gulati K, Wang N, Zhang Z, Ivanovski S (2018)
contributed to the implementation of the research Bridging the gap: optimized fabrication of robust titania
and to the interpretation of the results. RC designed nanostructures on complex implant geometries towards
and supervised the study and revised the clinical translation. J Colloid Interface Sci 529:452–463. h
manuscript. ttps://doi.org/10.1016/j.jcis.2018.06.004
[12] Contri Campanelli L, Pereira Sergio Carvalho, da Silva P,
Compliance with ethical standards Camarinho Oliveira NT, Bolfarini C (2017) Effect of the
modification by titanium dioxide nanotubes with different
Conflict of interest All authors declare that they structures on the fatigue response of Ti grade 2. Mater Res
have no conflict of interest. 20:120–124. https://doi.org/10.1590/1980-5373-mr-2016-
0681

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


13738 J Mater Sci (2019) 54:13724–13739

[13] Low IM, Yam FK, Pang WK (2012) In-situ diffraction 759:122–128. https://doi.org/10.1016/j.jelechem.2015.11.
studies on the crystallization and crystal growth in anodized 002
TiO2 nanofibres. Mater Lett 87:150–152. https://doi.org/10. [25] Sulka GD, Kapusta-Kołodziej J, Brzózka A, Jaskuła M
1016/j.matlet.2012.08.001 (2013) Anodic growth of TiO2 nanopore arrays at various
[14] Allam NK, Grimes CA (2008) Effect of cathode material on temperatures. Electrochim Acta 104:526–535. https://doi.or
the morphology and photoelectrochemical properties of g/10.1016/j.electacta.2012.12.121
vertically oriented TiO2 nanotube arrays. Sol Energy Mater [26] Butail G, Ganesan PG, Raddiar M, Teki R, Ravishankar N,
Sol Cells 92:1468–1475. https://doi.org/10.1016/j.solmat.20 Duquette DJ, Ramanath G (2011) Kinetics of titania nan-
08.06.007 otube formation by anodization of titanium films. Thin Solid
[15] Gong D, Grimes CA, Varghese OK, Hu W, Singh RS, Chen Films 519:1821–1824. https://doi.org/10.1016/j.tsf.2010.10.
Z, Dickey EC (2001) Titanium oxide nanotube arrays pre- 004
pared by anodic oxidation. J Mater Res 16:3331–3334. h [27] Apolinário A, Sousa CT, Ventura J, Costa JD, Leitão DC,
ttps://doi.org/10.1557/JMR.2001.0457 Moreira JM, Sousa JB, Andrade L, Mendes AM, Araújo JP
[16] Raja KS, Misra M, Paramguru K (2005) Formation of self- (2014) The role of the Ti surface roughness in the self-
ordered nano-tubular structure of anodic oxide layer on ordering of TiO2 nanotubes: a detailed study of the growth
titanium. Electrochim Acta 51:154–165. https://doi.org/10. mechanism. J Mater Chem A 2:9067–9078. https://doi.org/
1016/J.ELECTACTA.2005.04.011 10.1039/C4TA00871E
[17] Regonini D, Bowen CR, Jaroenworaluck A, Stevens R [28] Shin Y, Lee S (2008) Self-organized regular arrays of anodic
(2013) A review of growth mechanism, structure and crys- TiO2 nanotubes. Nano Lett 8:3171–3173. https://doi.org/10.
tallinity of anodized TiO2 nanotubes. Mater Sci Eng R Rep 1021/nl801422w
74:377–406. https://doi.org/10.1016/j.mser.2013.10.001 [29] Leonardi S, Russo V, Li Bassi A, Di Fonzo F, Murray TM,
[18] Tan AW, Pingguan-Murphy B, Ahmad R, Akbar SA (2013) Efstathiadis H, Agnoli A, Kunze-Liebhäuser J (2015) TiO2
Advances in fabrication of TiO2 nanofiber/nanowire arrays nanotubes: interdependence of substrate grain orientation
toward the cellular response in biomedical implantations: a and growth rate. ACS Appl Mater Interfaces 7:1662–1668. h
review. J Mater Sci 48:8337–8353. https://doi.org/10.1007/ ttps://doi.org/10.1021/am507181p
s10853-013-7659-0 [30] Crawford GA, Chawla N (2009) Tailoring TiO2 nanotube
[19] Lim JH, Choi J (2007) Titanium oxide nanowires originating growth during anodic oxidation by crystallographic orien-
from anodically grown nanotubes: the bamboo-splitting tation of Ti. Scr Mater 60:874–877. https://doi.org/10.1016/j.
model. Small 3:1504–1507. https://doi.org/10.1002/smll. scriptamat.2009.01.043
200700114 [31] Hielscher R, Schaeben H (2008) A novel pole figure inver-
[20] Ismail S, Lockman Z, Ahmad ZA (2012) Crystallization of sion method: specification of the MTEX algorithm. J Appl
TiO2 nanotubes arrays grown by anodization of Ti in organic Crystallogr 41:1024–1037. https://doi.org/10.1107/
electrolyte. Adv Mater Res 620:412–417. https://doi.org/10. S0021889808030112
4028/www.scientific.net/AMR.620.412 [32] Ozkan S, Mazare A, Schmuki P (2018) Critical parameters
[21] Uchida M, Kim H-M, Kokubo T, Fujibayashi S, Nakamura T and factors in the formation of spaced TiO2 nanotubes by
(2003) Structural dependence of apatite formation on titania self-organizing anodization. Electrochim Acta 268:435–447.
gels in a simulated body fluid. J Biomed Mater Res https://doi.org/10.1016/j.electacta.2018.02.120
64A:164–170. https://doi.org/10.1002/jbm.a.10414 [33] Ozkan S, Nguyen NT, Mazare A, Hahn R, Cerri I, Schmuki
[22] Luttrell T, Halpegamage S, Tao J, Kramer A, Sutter E, P (2017) Fast growth of TiO2 nanotube arrays with con-
Batzill M (2015) Why is anatase a better photocatalyst than trolled tube spacing based on a self-ordering process at two
rutile? Model studies on epitaxial TiO2 films. Sci Rep different scales. Electrochem Commun 77:98–102. https://d
4:4043. https://doi.org/10.1038/srep04043 oi.org/10.1016/j.elecom.2017.03.007
[23] Regonini D, Satka A, Jaroenworaluck A, Allsopp DWE, [34] Berger S, Hahn R, Roy P, Schmuki P (2010) Self-organized
Bowen CR, Stevens R (2012) Factors influencing surface TiO2 nanotubes: factors affecting their morphology and
morphology of anodized TiO2 nanotubes. Electrochim Acta properties. Phys status solidi 247:2424–2435. https://doi.org/
74:244–253. https://doi.org/10.1016/j.electacta.2012.04.076 10.1002/pssb.201046373
[24] Sopha H, Hromadko L, Nechvilova K, Macak JM (2015) [35] Conradie F, Treurnicht N, Sacks N (2014) Alpha case
Effect of electrolyte age and potential changes on the mor- characterization of hot rolled titanium. Adv Mater Res
phology of TiO2 nanotubes. J Electroanal Chem 1019:311–317. https://doi.org/10.4028/www.scientific.net/A
MR.1019.311

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci (2019) 54:13724–13739 13739

[36] Peng W, Zeng W, Zhang Y, Shi C, Quan B, Wu J (2013) The based substrate before and after thermal treatment. J Nanosci
effect of colored titanium oxides on the color change on the Nanotechnol 19:2989–2996. https://doi.org/10.1166/jnn.201
surface of Ti–5Al–5Mo–5V–1Cr–1Fe alloy. J Mater Eng 9.15859
Perform 22:2588–2593. https://doi.org/10.1007/s11665-013-
0573-4 Publisher’s Note Springer Nature remains neutral with
[37] Losertová M, Štefek O, Galajda M, Konečná K, Martynkova regard to jurisdictional claims in published maps and
GS, Barabaszová KČ (2019) Microstructure and electro- institutional affiliations.
chemical behavior of TiO2 nanotubes coated on titanium-

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


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:

1. 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;
2. 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;
3. falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
4. use bots or other automated methods to access the content or redirect messages
5. override any security feature or exclusionary protocol; or
6. 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

You might also like