You are on page 1of 6

Journal of Alloys and Compounds 541 (2012) 144–149

Contents lists available at SciVerse ScienceDirect

Journal of Alloys and Compounds


journal homepage: www.elsevier.com/locate/jalcom

Infrared normal spectral emissivity of Ti–6Al–4V alloy in the 500–1150 K


temperature range
L. González-Fernández a,b, E. Risueño c, R.B. Pérez-Sáez a,d,⇑, M.J. Tello a,d
a
Departamento de Física de la Materia Condensada, Facultad de Ciencia y Tecnología, Universidad del País Vasco, Barrio Sarriena s/n, 48940 Leioa, Bizkaia, Spain
b
Industria de Turbo Propulsores, S.A., Planta de Zamudio, Edificio 300, 48170 Zamudio, Bizkaia, Spain
c
CIC Energigune, Parque Tecnológico, Albert Einstein 48, 01510 Miñano, Álava, Spain.
d
Instituto de Síntesis y Estudio de Materiales, Universidad del País Vasco, Apdo. 644,48080 Bilbao, Spain.

a r t i c l e i n f o a b s t r a c t

Article history: Thermal radiative emissivity is related to the optical and electrical properties of materials, and it is a key
Received 23 March 2012 parameter required in a large number of industrial applications. In the case of Ti–6Al–4V, spectral emis-
Received in revised form 21 June 2012 sivity experimental data are not available for the range of temperatures between 400 and 1200 K, where
Accepted 22 June 2012
almost all industrial applications take place. The experimental results in this paper show that the normal
Available online 29 June 2012
spectral emissivity decreases with wavelength from a value of about 0.35 at 2.5 lm to about 0.10 at
22 lm. At the same time, the spectral emissivity shows a slight linear increase with temperature between
Keywords:
500 and 1150 K, with approximately the same slope for all wavelengths. Additionally, the influence of the
Titanium alloy
Infrared emissivity
samples thermal history on the emissivity is studied. A strong decrease in the emissivity values appears
Electrical resistivity due to the effect of surface stress relaxation processes. This means that the radiative properties of this
Surface stresses alloy strongly depend on the surface stress state. A thermal treatment to relieve the surface stress should
be carried out to achieve a steady state of the radiative properties. In addition, a good qualitative
agreement is found between the temperature dependence of the electrical resistivity obtained using con-
ventional measurements and the one obtained from the emissivity experimental results by using the
Hagen–Rubens equation.
Ó 2012 Elsevier B.V. All rights reserved.

1. Introduction other thermophysical properties at high temperature, for example,


specific heat [11]. Emissivity is related to the optical and electrical
The Ti–6Al–4V alloy is widely used in engineering and biomed- properties of the materials [12,13], and it is a key parameter
ical applications because of its great combination of high strength, required in heat-transfer calculations for several industrial applica-
low density and excellent corrosion resistance. These properties tions, for example, in aerospace and aircraft components [1]. It is
have made it a world standard in many applications ranging from important to mention that emissivity is a surface property that de-
aerospace and aircraft components to marine, power generation, pends on temperature, wavelength, direction, surface morphology,
high performance automotive parts, sports and medicine [1–3]. contamination, aging and oxidation state [14,15]. Therefore, in or-
However, there are some limitations related to its low wear perfor- der to compare results, the experimental data available in the liter-
mance in abrasive and adhesive wearing conditions. A lot of work ature should be taken with caution.
was made in order to improve its tribological characteristics, see In spite of the importance of the emissivity data for a large
for example Refs. [4–7] and references therein. Recently, several number of Ti–6Al–4V alloy applications, relatively few studies
studies have focused on the development of new Ti alloys that have been reported on its normal or hemispherical emissivity. It
could potentially be used for biomedical applications [8,9]. is difficult to find emissivity experimental data that can be used
Knowledge of the complete set of thermophysical properties of in applied research or in engineering design related to thermal
the Ti–6Al–4V alloy is necessary for its industrial applications. radiation. The literature data are scarce, without surface control,
Among the thermophysical properties, thermal radiative emissiv- at specific wavelengths or temperatures, or for temperatures
ity has special relevance. Emissivity is essential for temperature (T > 1200 K) out of the range of most industrial applications
measurements by remote sensing devices in, for example, manu- [16–21]. A total hemispherical emissivity of 0.47 and 0.39 at
facturing processes [10]. It is also useful in the measurements of T > 1200 K for rough and smooth surfaces respectively was
obtained by using different experimental techniques [16,17]. In
⇑ Corresponding author. Tel.: +34 94 601 2655; fax: +34 94 601 3500.
addition, using the radiance temperature, a value of 0.395 was
E-mail address: raul.perez@ehu.es (R.B. Pérez-Sáez).
measured at melting (1943 K) for the normal emissivity at

0925-8388/$ - see front matter Ó 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jallcom.2012.06.117
L. González-Fernández et al. / Journal of Alloys and Compounds 541 (2012) 144–149 145

653 nm [17]. On the other hand, Betz et al. [18] found that the nor- blackbody follow equivalent optical paths to reach the detector. Inside the FT–IR
spectrometer, a diaphragm adjusts the sample area viewed by the detector. The
mal emissivity at 665 nm shows a nearly linear temperature
sample holder allows directional measurements and the sample chamber permits
dependence between 1200 and 1700 K. An extrapolation of these to measure at controlled atmosphere (vacuum, gas or open atmosphere). The
experimental values to the melting point yields a value of 0.369, sample is heated by using a resistive wire. The heating system ensures good tem-
which is close to those obtained using the radiance temperature perature homogeneity in the measured area of the sample. The sample temperature
in Ref. [17]. By means of the relation between the true temperature is measured by two K-type thermocouples spot-welded to the sample surface out of
the area viewed by the detector, in this case, a 4 mm diameter circle.
of the sample and the radiance temperature, a value of 0.769 was
The emissivity is obtained using the blacksur method [26] and the modified
found for the normal emissivity in the 1215–1225 K temperature two-temperature calibration method [25]. This election is supported by the analysis
range [19]. In this case, the radiance temperature was obtained of the accuracy of the methods for the direct emissivity measurements, the emissiv-
by a modified high temperature pyrometry for a single wavelength. ity error due to the calibration process, and the short and long-term temporal
stability of the calibration processes. The combined standard uncertainty of this de-
The effect of the concentration of aluminum and vanadium on the
vice was previously analyzed for all the uncertainty sources [27]. For the measure-
emissivity of titanium alloys was studied in Ref. [21] between 1100 ments presented in this paper the maximum combined standard uncertainty varies
and 1700 K by using electronic heating under conditions of oilless between 1 and 10% depending on wavelength and temperature, its average value
vacuum. Finally, it must be noticed that a fast laser polarimeter being around 4%.
was used to determine the emissivity of liquid Ti–6Al–4V at All measurements were made following the same procedure. Once the sample
had been introduced into the sample chamber, a moderate vacuum was made
684.5 nm between the melting temperature and 2920 K [20].
and, before the sample was heated up to the measurement temperatures, a slightly
As previously commented, taking into account the accuracy of reducing atmosphere (N2 + 5% H2) was introduced into the chamber in order to
the measurement methods and the experimental setups used, as prevent the oxidation of the sample.
well as the surface sample characterization and the environment
control, there are serious difficulties to compare the few emissivity
experimental data available. The lack of control of these parame- 3. Results and discussion
ters limits the usefulness of some of the available emissivity exper-
imental data. Therefore, precise emissivity measurements for this The first part of this paper is dedicated to study the influence of
alloy should be carried out in the temperature range of technolog- thermal history in the emissivity of the Ti–6Al–4V alloy. The sec-
ical interest for a broader spectral range. ond part focuses on presenting and analyzing the normal spectral
The main objective of this paper is to investigate the thermal emissivity experimental results. Likewise, a comparison with the
radiative emissivity of the Ti–6Al–4V alloy between 500 and few available experimental data in literature is carried out. In
1150 K in the medium infrared spectral range. The influence of addition, the Hagen–Rubens approach has allowed a comparison
both the thermal cycling and the surface roughness on the emissiv- between the electrical resistivity obtained using conventional
ity values is also studied. It should be emphasized that emissivity measurements and those obtained from the emissivity experimen-
experimental data are not available in the literature for this tem- tal. Each of the nine samples underwent several thermal cycles, up
perature range, where almost all the industrial applications of this to four in most cases. In each heating cycle, the emissivity was
alloy take place. The present study is done with an experimental measured at seven temperatures between 500 and 1150 K approx-
setup, which allows to obtain high accuracy emissivity data with imately every 100 K, with a 20 K/min typical heating rate and a sta-
a good sample surface control during all the experiments. bilization time of around 20 min at each measurement
temperature. Afterwards, the sample was free-cooled to room
temperature.
2. Experimental
As expected, very small differences in the experimental values
Nine samples of a commercial Ti–6Al–4V alloy have been used, with a disc were found between samples with similar roughness. As an exam-
shape of 55 mm of diameter and 3 mm of thickness. Before the experimental emis-
sivity measurements, the samples have been characterized by X-ray diffraction
(XRD) and scanning electron microscopy (SEM). The XRD patterns show the peaks
corresponding to a-Ti and b-Ti phases but with a small shift in the diffraction angles
due to the inclusion of aluminum and vanadium in the alloy structure [22,23]. The
samples have been polished with SiC paper up to 320 grit, except for samples 3 and
4, which have been polished up to 4000 grit to obtain smaller roughness values. Its
roughness has been measured in two perpendicular directions with a Mitutoyo
SJ201 roughness tester. Table 1 shows the average roughness values of each sample.
It should be noted that samples 3 and 4 have a smoother surface than the others.
A high accuracy infrared radiometer (HAIRL) [24] has been used to obtain the
normal spectral emissivity of the samples between 500 and 1150 K. The three main
modules of the experimental device are a FT–IR spectrometer with a DTGLaS
pyroelectric detector, a blackbody and a sample chamber. The FT–IR spectrometer
provides a rapid accurate detection and processing of the signal. The blackbody is
used to calibrate the radiometer [25]. The thermal radiations of both sample and

Table 1
Summary of the samples surface roughness: roughness average (Ra), average
maximum height (Rz), and maximum height of the profile (Rt).

Sample Ra (lm) Rz (lm) Rt (lm)


1 0.46 3.70 4.77
2 0.31 2.62 3.33
3 0.06 0.50 0.58
4 0.05 0.39 0.52
5 0.31 2.44 3.47
6 0.26 2.06 2.46
7 0.27 2.07 2.54
8 0.24 1.86 2.25 Fig. 1. Normal spectral emissivity (e) of the Ti–6Al–4V alloy as a function of
9 0.22 1.65 1.91 wavelength (k) for three temperatures: (a) sample 1 and (b) sample 2. Dashed and
solid lines refer to the first and fourth heating cycles respectively.
146 L. González-Fernández et al. / Journal of Alloys and Compounds 541 (2012) 144–149

ple, Fig. 1 shows the normal spectral emissivity between 2.5 and
22 lm for samples 1 (Fig. 1a) and 2 (Fig. 1b). The emissivity spectra
at three temperatures are shown for the first (dashed lines) and
fourth (solid lines) thermal cycles. The emissivity decreases with
the wavelength according to the electromagnetic theory for metal-
lic materials [12,13].

3.1. Thermal history

A careful analysis of Fig. 1 shows a decrease of the emissivity


values in all the spectral range between the first and fourth
thermal cycle, i.e. the emissivity obtained at the same temperature
is, in general, lower in the fourth cycle than in the first one. For the
sake of a better visualization of this apparently anomalous behav-
ior, the normal emissivity at 4, 12 and 20 lm as a function of the
heating cycle number is plotted in Fig. 2 for low and medium
temperatures. The significant decrease of the emissivity values
from the first to the second thermal cycle is clearly shown. How-
ever, a substantial change in the emissivity values in the following
heating cycles cannot be observed. A similar behavior was previ-
ously noticed in other metallic materials as Armco iron [15]. In or-
der to check the evolution of the emissivity during the first thermal
cycle, the normal spectral emissivity for several temperatures is
plotted in Fig. 3. It is remarkable that, between 546 and 1127 K,
the emissivity goes through a maximum at 858 K. A similar
qualitative behavior was also found in measurements of electrical Fig. 3. Normal spectral emissivity of the sample 2 obtained during the first heating
resistivity for the same temperature range [20]. This feature was cycle, as a function of wavelength (a), and as a function of temperature (b).
also observed during a flash heating as a function of time [17] or
enthalpy [28]. Since the alloy studied is composed mainly of Ti, it
is also interesting to mention that the electrical resistivity of Ti cycle, because during thermal oxidation the oxide layer becomes
exhibits the same feature just before the alpha-to-beta transforma- porous, and this porosity develops a stratified structure, which
tion [29]. produces an increase in the surface roughness [22,30–32]. On the
Taking into account that thermal emissivity is a surface prop- other hand, it is important to mention that the XRD patterns show
erty, the decrease in the value of the normal spectral emissivity be- no change in the phase composition of the samples (a + b) before
tween the first and the second cycle should necessarily be related and after the thermal cycles.
to a change in the sample surface. A possible cause for the alter- The physical surface change that can explain this apparently
ation of the sample surface could be an oxidation process due to anomalous behavior is the presence of surface stresses generated
a poor atmosphere control. However, an analysis of the surface during the machining and grinding of the samples. Assuming this
after four heating cycles by means of XRD and SEM showed no fact, the first heating cycle acts as an annealing process that re-
signs of oxidation. Furthermore, the oxidation usually produces lieves the surface stresses, explaining the lower emissivity values
an emissivity increase, which is the opposite of what occurs in after the first cycle. In this case, if a new machining and/or grinding
our experimental results. Another possible change of the sample process is performed on the sample, the emissivity value should re-
surface could be a roughness decrease during the first heating cover the preheating value. To check this point, sample 9 was grin-
cycle. However, roughness measurements carried out before and ded after four thermal cycles, and then the emissivity was
after the first cycle showed very small changes, which are within measured for four cycles more. The obtained emissivity results
the experimental uncertainty. In any case, these changes are not were similar to those in Fig. 2. This result confirms that the ob-
enough to explain the decrease of emissivity. In addition, this served behavior of the emissivity is due to the relief of the sample
result confirms again the absence of oxidation during the heating surface stresses produced during the grinding.
On the other hand, it is known that annealing usually involves
thermally activated processes. An activation energy must be over-
come in order for these processes to progress. This may statistically
occur at any temperature, but when the temperature increases, so
does the system energy, and the rate of the process is higher. As it
can be seen in Fig. 3b, the emissivity increases for the first four
temperatures of the first heating cycle and then it begins to de-
crease. This seems to indicate that the relieving process is so slow
for temperatures below 858 K that the surface stresses are not re-
moved, but above this temperature the relieving process starts to
take place in a noticeable manner. To check this fact, the emissivity
of sample 7 at a constant temperature around 870 K was measured
for more than one hour during the first heating cycle. In Fig. 4, the
evolution of the emissivity during the measurement is plotted for
two wavelengths. A clear decreasing tendency is observed in both
Fig. 2. Evolution of the normal emissivity of the sample 2 with the thermal cycle
cases. However, the decrease rate is not very high, at least not as
number at three wavelengths (k = 4, 12 and 20 lm) for low and medium
temperatures. high as it would be expected on seeing the variation between
L. González-Fernández et al. / Journal of Alloys and Compounds 541 (2012) 144–149 147

the studied spectral range. Fig. 5 shows the emissivity spectra at sev-
eral temperatures for a smooth sample (sample 4) and a rougher one
(sample 2). The small differences between the experimental emis-
sivity values of samples 2 and 4 are in agreement with the surface
roughness dependence of the normal spectral emissivity. The emis-
sivity increases slightly with temperature and decreases with wave-
length. The temperature dependence of the emissivity is linear with
a small slope, as shown in Fig. 6 for sample 2 (fourth heating cycle).
The same slope is approximately observed for long wavelengths
(>10 lm). Deviations from linear behavior are within the experi-
mental uncertainty.
In order to compare our results with those in the literature, it
has to be taken into account that, as it has been previously men-
Fig. 4. Time evolution of the normal spectral emissivity at 870 K for 5 and 10 lm tioned, the emissivity values found in the literature are scarce, usu-
(sample 7, first heating cycle).
ally limited to a very high temperature range, at short wavelengths
and without sample surface characterization or control. Consider-
858 and 949 K shown if the Fig. 3b for the same wavelengths. ing the small measured temperature dependence of the emissivity
Nevertheless, it has to be taken into account that the rate of the (Fig. 6), the temperature difference between results is not so
thermally activated processes exponentially increases with important. However, our results must be extrapolated to short
temperature, and that to carry out each measurement the sample wavelengths to be compared to data obtained at around 660 nm
temperature is stabilized for around 20 min to ensure the thermal [17,18,20,21] and between 0.8 and 1.1 lm [19]. The emissivity
stability. Therefore, each measurement involves around 20 min of values of 0.395 [17] and 0.369 [18] at a temperature near the melt-
relieving process at that temperature. Thus, the emissivity change ing point, and between 0.23 and 0.30 in the 1100–1700 tempera-
shown in Fig. 3b between 858 and 949 K cannot be compared to ture range [21] are smaller than the results obtained
the one observed in Fig. 4. The accumulated relief at each temper- extrapolating the data of Fig. 5. These differences can be partially
ature is the cause for having almost completely relaxed all the explained because of the possible presence of an emissivity
surface stresses when the sample temperature reaches 1127 K. X-point, reported for titanium at around 1.4 lm [33], which pro-
Above this temperature, the emissivity values are very close to duces opposite temperature trends at visible and IR. Thus, a nega-
those obtained during the second cycle. tive temperature coefficient will be observed in the visible,
whereas a positive temperature coefficient is observed in the IR.
3.2. Normal emissivity and electrical resistivity On the other hand, our extrapolated data (0.50-0-55) are in rela-
tively good agreement with the values given in Ref. [20], which
The normal spectral emissivity of the nine samples specified in vary between 0.56 and 0.60. This comparison must be made with
Table 1 was measured between 500 and 1150 K for several thermal some caution, considering that the laser polarimeter used to obtain
cycles. When the first cycle, which acts as an annealing process, was emissivity in Ref. [17] was not designed for measurements in the
not taken into account, the results were qualitatively equal for all the presence of surface roughness. A normal emissivity of 0.769
samples. The observed differences were due to variations in surface around 1 lm between 1215 and 1225 K is provided by Ref. [19].
roughness. In all cases, the emissivity values obtained for the second This value is much higher than those shown in Fig. 5. These
heating cycle and the following ones varied between 0.35 and 0.10 in differences can be due to a different state of the sample surface
(roughness, oxidation, etc.), together with the uncertainty associ-
ated to the measurement method based on pyrometry.
Classic electromagnetic theory allows to determine the normal
spectral emissivity from the resistivity values by means of the two-
term approximation to the Hagen–Rubens relation [12]:
rffiffiffiffi
r r
eðkÞ ¼ 36:5  e  464  e ð1Þ
k k

where k is the wavelength given in micrometer and re is the electri-


cal resistivity in ohms centimeter. This relation is valid for smooth

Fig. 5. Normal emissivity spectra for several temperatures, (a) for a smooth sample
(sample 4) during the second heating cycle, and (b) for a rougher sample (sample 2) Fig. 6. Temperature dependence of the normal spectral emissivity of sample 2
during the fourth heating cycle. (fourth heating cycle) at 5, 10, 15 and 20 lm.
148 L. González-Fernández et al. / Journal of Alloys and Compounds 541 (2012) 144–149

measured in some electrical resistivity data [17,20,28]. However,


for the second and the following thermal cycles the calculated
resistivity is nearly linear with the temperature. As shown in
Fig. 8, this nearly linear behavior is also observed in other direct
measurements of resistivity [34,35]. Therefore, the anomalous
behavior found in the electrical resistivity [17,20,28] could be asso-
ciated with the surface stress relieving. Instead, this seems to be
more linked to the (a + b) to b transformation, as the authors
stated, because the anomalous behavior in the electrical resistivity
was reproduced in all heating cycles. However, in these cases, it is
possible that the samples were not exposed to a high temperature
long enough to produce the stress relieving and, as a consequence,
the feature could be observed in the following heating cycles. Any-
Fig. 7. Normal emissivity spectra for several temperatures, obtained experimen- way, the anomalous behavior shown in Fig. 8 for the calculated
tally for the sample 2 during the fourth heating cycle (solid lines) and calculated by resistivity is observed only during the first heating cycle and is
means of Hagen–Rubens relation with the electrical resistivity of Ref. [35] (dashed
clearly due to the stress relieving. In order to observe the possible
lines).
change due to the (a + b) to b transformation in our measurements,
we would need to extend our measurement capabilities to higher
temperatures. In the case of titanium, a similar feature is reported
for the emissivity and the electrical resistivity [36]. The emissivity
change was only observed in the first heating and was associated
to the sublimation of the oxides and the degassing of the surface,
whereas the resistivity change was repeatable and linked to the
alpha–beta transition of titanium.
On the other hand, it must be remarked that the qualitative
agreement of Eq. (1) with the experimental data implies that the
primary contribution to the thermal radiation phenomena in this
alloy is the intraband transition of electrons [14]. Nevertheless,
the electrical resistivity values obtained from emissivity data are
slightly higher than those of direct measurements. The same
behavior was found for Armco iron [15]. If we take into account
Fig. 8. Electrical resistivity versus temperature. Curves (1) and (2) show values the approximations included in the Hagen–Rubens relation, this
calculated from emissivity measurements for first (sample 2) and second (sample 4) small difference can be related to the sample surface roughness.
heating cycle respectively. Curves (3) and (4) show electrical resistivity data More work in this direction is being carried out in our laboratory
measured in Refs. [34,35] respectively. with pure metals.

4. Conclusions
metallic surfaces and for wavelengths longer than 5 lm. This
equation is a good qualitative approach, and sometimes exhibits a
This paper presents the first experimental study on normal
good quantitative agreement [14,15] for wavelengths as short as
spectral emissivity of the Ti–6Al–4V alloy in the temperature range
5 lm.
between 500 and 1150 K. The emissivity decreases with wave-
Eq. (1) allows to use the available electrical resistivity data of
length, according to electromagnetic theory, and linearly increases
Ti–6Al–4V alloy [17,20,28,34,35] to obtain emissivity values.
with temperature with a small slope, which is independent of the
Unfortunately, the sample conditions are not specified for the
wavelength for k > 10 lm.
electrical resistivity measurements. In Fig. 7, the normal spectral
The Hagen–Rubens approach allows a comparison between
emissivity of the sample 2 and the spectra obtained using the Ha-
experimental electrical resistivity data and those obtained from
gen–Rubens relation and the electrical resistivity data of Ref. [35]
the emissivity measured for smooth surface samples. A good qual-
are plotted. It is remarkable that, although the Hagen–Rubens ap-
itative agreement between them is found.
proach is applicable for relatively long wavelengths (k >  5 lm),
An important evolution in the normal spectral emissivity values
the agreement is good in all the spectral range. Theoretically, the
between the first and second thermal cycle has been observed. The
comparison between the experimental emissivity and the one cal-
experimental results suggest that these changes in emissivity are
culated using the Hagen–Rubens relation should be done using the
related to the relieving during the first heating of the surface stres-
emissivity spectra from the smoothest sample. However, the best
ses generated during machining and grinding. In order to obtain
agreement has been obtained using a rougher sample (sample 2).
reproducible emissivity and electrical resistivity data, the
In a similar way, the electrical resistivity can be obtained by fit-
measurements must be carried out after several thermal cycles
ting the experimental emissivity curves to Eq. (1). Fig. 8 shows the
or after a suitable annealing process above 1200 K. After the stress
electrical resistivity calculated from emissivity spectra obtained
relieving process, there are no differences between measurements
during the first and second thermal cycle, together with the
carried out for different heating cycles.
literature data of Refs. [34,35], where the electrical resistivity is
measured for T > 1350 and T < 850 K respectively. The first cycle
data correspond to sample 2, and the second ones to sample 4. Acknowledgements
The evolution observed in the emissivity during the first heating
cycle can also be observed in the calculated electrical resistivity This work has been carried out with the financial support of the
values. This feature is observed in all the samples, but it is dis- SAIOTEK program (Project number S-PC08UN07) of the Basque
played for sample 2 because it is more noticeable for a rougher Government and the ‘‘Universidad-Empresa’’ program (Project
sample. It is important to notice that the same behavior is also number UE06/01) of the University of the Basque Country in
L. González-Fernández et al. / Journal of Alloys and Compounds 541 (2012) 144–149 149

collaboration with ‘‘Industria de Turbo Propulsores S.A.’’. L. Gon- [15] L. del Campo, R.B. Pérez-Sáez, M.J. Tello, X. Esquisabel, I. Fernández, Int. J.
Thermophys. 27 (2006) 1160–1172.
zález-Fernández acknowledges the Basque Government and Indu-
[16] H. de L’Estoile, L. Rosental, Advisory Group for Aeronautical Research and
stria de Turbo Populsores S.A. for their support through a Ph.D. Development, Paris, France, AGARD-211 N63-21549, (1958).
fellowship. We thank the reviewer for his/her thorough report [17] A. Cezairliyan, J.L. McClure, R. Taylor, J. Res. Nat. Bur. Stand. A, Phys. Chem. 81
and highly appreciate the insightful and constructive comments (1977) 251–256.
[18] H.T. Betz, O.H. Olson, B.D. Schurin, J.C. Morris, WADC-TR-56-222 (Part 2), 1957,
and suggestions. 1-184 AD202493.
[19] P. Coppa, A. Consorti, Measurement 38 (2005) 124–131.
References [20] M. Boivineau, C. Cagran, D. Doytier, V. Eyraud, M.-H. Nadal, B. Wilthan, G.
Pottlacher, Int. J. Thermophys. 27 (2006) 507–529.
[21] B.A. Shur, V.E. Peletskii, High Temp. 42 (2004) 414–420.
[1] R. Boyer, G. Welch, E.W. Collings, Materials Properties Handbook - Titanium [22] S. Kumar, T.S.N.S. Narayanan, S.G.S. Raman, S.K. Seshadri, Mater. Chem. Phys.
Alloys, ASM International, ASM International, Ohio, 1994.
119 (2010) 337–346.
[2] M.R. Winstone, Titanium Alloys at Elevated Temperature: Structural [23] H. Güleryüz, H. Cimenoğlu, Biomaterials 25 (2004) 3325–3333.
Development and Service Behavior, IOM Comunications, London, 2001. [24] L. del Campo, R.B. Pérez-Sáez, X. Esquisabel, I. Fernández, M.J. Tello, Rev. Sci.
[3] P.A. Dearnley, Proc. Inst. Mech. Eng. Part H–J, Eng. Med. 213 (1999) 107–135. Instrum. 77 (2006) 113111.
[4] C. Boettcher, T. Bell, H. Dong, Metall. Mater. Trans. A 33A (2002) 1201–1211. [25] L. González-Fernández, R.B. Pérez-Sáez, L. del Campo, M.J. Tello, Appl. Opt. 49
[5] J. Szewczenko, W. Walke, K. Nowinska, J. Marciniak, Mater. Werkst. 41 (2010)
(2010) 2728–2735.
360–371. [26] R.B. Pérez-Sáez, L. del Campo, M.J. Tello, Int. J. Thermophys. 29 (2008) 1141–
[6] S.L. R da Silva, L.O. Kerber, L. Amaral, C.A. dos Santos, Surf. Coat. Technol. 116 1155.
(1999) 342–346. [27] L. del Campo, R.B. Pérez-Sáez, L. González-Fernández, M.J. Tello, J. Appl. Phys.
[7] V. Sreedhar, J. Das, R. Mitra, S.K. Roy, J. Alloys Compd. 519 (2012) 106–111. 107 (2010) 113510.
[8] S.B. Gabriel, J.V.P. Panaino, I.D. Santos, L.S. Araujo, P.R. Mei, L.H. de Almeida,
[28] E. Kaschnitz, P. Reiter, J.L. McClure, Int. J. Thermophys. 23 (2002) 267–275.
C.A. Nunes, J. Alloys Compd. In Press. (2011) http://dx.doi.org/10.1016/ [29] B. Wilthan, C. Cagran, G. Pottlacher, Int. J. Thermophys. 26 (2005) 1017–1029.
j.jallcom.2011.11.035. [30] H. Dong, A. Bloyce, P.H. Morton, T. Bell, Surf. Eng. 13 (1997) 402–406.
[9] C. Zhao, X. Zhang, P. Cao, J. Alloys Compd. 509 (2011) 8235–8238. [31] G. Bertrand, K. Jarraya, J.M. Chaix, Oxid. Met. 21 (1983) 1–19.
[10] J. Pujana, L. del Campo, R.B. Pérez-Sáez, M.J. Tello, I. Gallego, P.J. Arrazola, Meas. [32] F. Borgioli, E. Galvanetto, F. Iozzelli, G. Pradelli, Mater. Lett. 59 (2005) 2159–
Sci. Technol. 18 (2007) 3409–3416.
2162.
[11] J.J.Z. Li, Study of liquid metals by electrostatic levitation, Ph. D. Thesis, [33] P. Herve, A. Sadou, Infrared Phys. Technol. 51 (2008) 249–255.
California Institute of Technology (USA), 2009.
[34] D. Basak, R.A. Overfelt, D. Wang, Int. J. Thermophys. 24 (2003) 1721–1733.
[12] R. Siegel, J. Howell, Thermal Radiation Heat Transfer, fourth ed., Taylor & [35] M.W. Mote, R.B. Hooper, P.D. Frost, TML Rep. 92 (1958).
Francis, Washington, 2002. [36] G.A. Zhorov, High Temp. 5 (1967) 881-888 – translated from Teplofizika
[13] M.F. Modest, Radiative Heat Transfer, second ed., Academic Press, San Diego, Vysokikh Temperatur 5 (1967) 987-994.
2003.
[14] G. Teodorescu, Radiative emissivity of metals and oxidized metals at high
temperature, Ph. D. Thesis, Auburn University, Alabama (USA), (2007).

You might also like