You are on page 1of 10

J. Am. Ceram. Soc.

, 97 [1] 163–172 (2013)


DOI: 10.1111/jace.12657
© 2013 The American Ceramic Society

Journal
A Comprehensive Study of the Carbon Contamination in Tellurite Glasses and
Glass-Ceramics Sintered by Spark Plasma Sintering (SPS)
Anthony Bertrand,‡,§ Julie Carreaud,‡,§ Gaëlle Delaizir,‡,§,† Jean-René Duclere,‡,§ Maggy Colas,‡,§
Julie Cornette,‡,§ Marion Vandenhende,‡,§ Vincent Couderc,§,¶ and Philippe Thomas‡,§

Sciences des Procedes Ceramique et de Traitements de Surface (SPCTS), UMR 7315 CNRS, Centre Europeen de la Ceramique,
Limoges, France
§
Labex Sigma-Lim, Universite de Limoges, Limoges, France

XLIM, UMR 7252 CNRS, Universite de Limoges, Limoges, France

Highly transparent tellurite glasses and glass-ceramics based them, one is of particular interest to us, that is, c-TeO2, as
on the 85TeO2–15WO3 composition (mol%) were produced by its crystal structure was determined in our laboratory.4
spark plasma sintering (SPS) from powders previously pre- The control of the partial crystallization remains a chal-
pared by the conventional melt-quenching technique and then lenge for the targeted applications especially for optical
grinded. We report a study based on the understanding of car- applications. Indeed, the crystals size must remain small
bon contamination that is commonly observed by this noncon- enough to avoid too much light scattering according to the
ventional sintering technique and which constitutes a drawback Rayleigh theory to keep transparent glass-ceramics. Another
for optical applications. First, the influence of the particle size approach consists in matching the refractive indices of
of the initial amorphous powders as well as an additional pres- both the crystals and the glassy matrix which also leads to
sureless sintering step prior to SPS experiments and the use of transparent glass-ceramics5 and ceramics.6
a physical carbon diffusion barrier have been investigated to There are basically different types of kinetics crystal growth
reduce the carbon contamination in glass bulks. Second, once such as: crystals which nucleate at the surface and grow in the
reducing the carbon contamination, glass-ceramics were bulk, surface nucleated crystals which grow two dimensionally
obtained by varying the SPS conditions. The noncentrosym- at the surface, internal crystals which grow in the bulk, etc.7
metric c-TeO2 phase crystallized within the bulk volume while The kinetics of crystals growth for one given glass composi-
maintaining good optical transparency and led to the genera- tion can be predicted by either isothermal or nonisothermal
tion of second harmonic. This approach paves the way to fur- treatments thanks to, respectively, the well-known Mehl–Av-
ther applications in the domain of linear and nonlinear optics. rami equation and the Ozawa and Kissinger equations.8–10
The kinetics is dependent upon several parameters such as the
glass composition, the presence of nucleating agent, the qual-
I. Introduction
ity of the surface, etc. If the surface crystallization is a major

T glasses are very attractive materials for optical


ELLURITE
applications due to their good transmission in the mid-
infrared range (up to 6 lm). These glasses also exhibit high
advantage in increasing the mechanical strength of the base
glass, it dramatically affects the optical properties by scatter-
ing light preventing the glass from any optical-related applica-
linear and nonlinear refractive indices and they can easily be tions. However, starting from glassy powder instead of bulk
doped with rare-earth ions for laser applications.1,2 The first could overcome this limitation. Indeed, grained glass samples
transparent tellurite glass-ceramics were developed by Koma- devitrify more easily than bulk glass samples due to the gener-
tsu et al. at the beginning of the 1990s, in the system ation of imperfections at the surface of each grain and upon a
TeO2–LiNbO33 in which they studied the second-order optical process of simultaneous amorphous state sintering and crys-
nonlinearity provided by the noncentrosymmetric crystalline tallization, glass-ceramic materials can be obtained. This pro-
phases. Indeed, due to their isotropic nature, glassy materials cess is known as sinter crystallization.11
do not generate any second harmonic (SH) signal, which is Recently, the nonconventional sintering technique, namely,
one of the basic phenomena in nonlinear optics (NLO). the spark plasma sintering (SPS) also known as field-assisted
Therefore, materials which combine the advantages of both sintering technique (FAST) appeared as a promising route
glass and crystalline compounds constitute some source of for obtaining glass-ceramics from powders.12 If metallic and
great attention, explaining why glass-ceramics research chalcogenide glasses have been studied so far,12,13 no study
remains a key focus of the scientific community interested in reports the SPS sintering of tellurite glasses to the best of our
NLO properties. knowledge. The main advantages of SPS over others sintering
Among highly nonlinear crystals, LiNbO3 remains the techniques are the fast heating rates (up to 600°C/min) due
most attractive as it is extremely efficient to generate SH and to Joule heating and the online control of parameters (sinter-
matches well the refractive index value of tellurite glasses. ing through displacement of punches, pressure, temperature,
Other nonlinear phases were finally crystallized, or inserted, vacuum, etc.) allowing to obtain glass-ceramics with desired
in various tellurite glasses (e.g., BaTiO3, c-TeO2). Among crystalline fraction within few minutes while several hours
are required for a conventional heat treatment in a furnace.
The main drawback of SPS remains the graphite pollution in
J. Ballato—contributing editor the samples which is commonly observed. Several groups
have reported on significant carbon contamination in samples
such as in MgAl2O4 transparent ceramics,14 in soda lime15
or GeS2 chalcogenide glasses16 having undergone a SPS den-
Manuscript No. 33438. Received June 27, 2013; approved September 17, 2013.

sification treatment. This contamination was attributed to the
Author to whom correspondence should be addressed. e-mail: gaelle.delaizir@
unilim.fr
proximity of the graphite dies that conduct the high electric

163
164 Journal of the American Ceramic Society—Bertrand et al. Vol. 97, No. 1

current and is also enhanced by plasma-assisted deposition have also been used for some experiments. For all the SPS
within the sample undergoing consolidation.14 runs, the temperature was measured with a thermocouple
The influence of pressure, heating rate, particle size of ini- positioned at the close vicinity of the sample, through the die.
tial powder, chemical composition of the die seems to play an The SPS parameters were defined as follows:
important role on the carbon contamination.17 Nevertheless,
1. Temperature: 370°C.
to obtain high light transmission, all the sources of light scat-
2. Pressure: 50 MPa.
tering have to be avoided (carbon, inclusions, porosity, etc.).
3. Dwell time: From 0 to 10 min.
It is well-known that the addition of WO3 to tellurite glass
4. Atmosphere: Ar or air.
provides numerous advantages such as the enhancement of
chemical durability and resistance against devitrifica- The density of the different glass and glass-ceramic bulks
tion.2,18,19 This last point is of particular interest since only was determined through the Archimedean principle using a
surface crystallization of high nonlinear c-TeO2 phase is Kern balance. This balance enables weighing solids in air
reported in this binary composition when using the conven- (mair) as well as in a solvent (in our case, absolute ethanol
tional melt-quenching (MQ) technique and subsequent heat methanol). If the density of the buoyancy medium is known
treatment in a furnace.20 Surface crystallization, while (q0) the density of the solid (q) is calculated as follows:
providing enhancement of the mechanical properties, limits
drastically the optical potential of this particular glass. mair
q¼ q
This study, focused on the 85TeO2–15WO3 glass composi- mair  methanol 0
tion, aims at studying the influence of the initial glass particle
size, the use of carbon diffusion barrier (alumina, tantalum The microstructure of a fresh polished fracture on the cross
foil, platinum environment), and the effect of an additional section of SPS glass and glass-ceramics was observed using a
pressureless sintering step prior to SPS experiments on the Cambridge electronic microscope. The nature of the crystal-
optical properties of the samples (glass and glass-ceramics) line phases was determined using a D8 Bruker equipment
obtained through the SPS technique. (Billerica, MA) operating in the 2h range 10°–60° with the
CuKa1 radiation.
II. Experimental Procedure The optical transmission measurements were carried out in
the 300–3300 nm range, at normal incidence, using a Varian
The glassy samples with composition 85TeO2–15WO3 (mol%) Cary 5000 spectrophotometer (Palo Alto, CA) operating in a
were synthesized in platinum crucible in air. Raw materials, dual beam configuration.
TeO2 (Alfa Aesar, London, UK, 99.99%) and WO3 (Sigma- The schematic representation of the setup employed to
Aldrich, St. Louis, MO, 99.9%), were first weighed in pre carry out the SHG experiments is shown in Fig. 1. The light
determined quantities (to obtain a total batch of 30 g) and source employed was an Infrared passively Q-switched Pow-
heated in a furnace up to 850°C for 2 h. The samples were erChip NanoLaser, emitting a radiation at k = 1064 nm. It
quenched between two stainless plates. Two populations of was operated at a repetition rate of 1 kHz and had a pulse
particles were then selected: the first one named coarse parti- width of 500 ps. We operated with an unfocused laser beam,
cles in the following was obtained by grinding and sieving the with a beam diameter at the laser output of 2 mm. The fun-
particles, keeping those ranging from 80 to 250 lm. The sec- damental (1064 nm) beam energy was controlled by rotating
ond population named fine particles was obtained by ball- a half-wave plate. The maximum output energy that could be
milling (Fritsch, Idar Oberstein, Germany, pulverisette 6). delivered at the sample was 50 lJ/pulse. The SH light inten-
The particle size distribution was checked by a laser scattering sity generated by the sample was detected by a photomulti-
analyzer (Malvern, Worcestershire, UK, Mastersizer 2000). plier tube, connected to an oscilloscope. Any residual
Differential scanning calorimetry (DSC) experiments were fundamental light was removed using a 10-nm band pass
performed using TA instruments (New Castle, DE) Q100 interference filter centered at 532 nm (F2). In addition, care-
equipment (with a heating rate of 10°C/min). To investigate ful attention was paid to thoroughly control that the detected
the phase crystallization sequence during the glass powder signal was actually SHG (Cf. log–log plot) and not due to
(fine and coarse particles) heat-treatment, temperature Raman any residual fluorescence.
spectroscopy was performed from room temperature up to We finally emphasize that the SHG intensities displayed in
580°C. All data were recorded using a T64000 Jobin-Yvon this study correspond only to the signal collected in reflection
spectrophotometer at the 514 nm wavelength (Edison, NJ).
Heating of the samples from room temperature to 580°C
was achieved using a TS1500 Linkam device (Surrey, UK)
with the same heating rate as the DSC measurements, that is, H P
10°C/min. Raman spectra were recorded in backscattering S
LASER
mode using a long working distance objective (950), and 1064
1064
532

acquisition time was 10 s with a step of 100°C between room


temperature and 300°C and each 10°C between 300°C and L
580°C.
The image (400 lm 9 400 lm, 5 lm step) is reconstructed
from the integrated intensity of main c-TeO2 band (between L
658 and 695 cm1) obtained from the 6400 spectra. The col- F2
Oscilloscope
532

ors contrast represents the variation in the intensity fluctua-


tion associated with the stretching Te–O vibration band.
Spark plasma sintering performed with the Dr. Sinter 825
Syntex machine (Kawasaki, Japan) was used to sinter the PMT
glass powders to obtain amorphous bulks as well as glass-
ceramics. Different protocols were tested to avoid the graphite
contamination. First, the amorphous powder was introduced
in an 8-mm-diameter graphite die with or without carbon dif- Fig. 1. Sketch of the experimental setup used to detect SHG from
fusion barrier (alumina under and above the glass powder the SPS samples. The measurements are carried out only in reflection
and tantalum foil). Second, prior to the SPS experiment, a geometry. H, half-wave plate; L, lenses; P, polarizer; PMT,
pressureless sintering at Tg + 30°C for 1 h was performed. photomultiplier tube; F2, 10-nm band pass interference filter
WC (tungsten carbide) die as well as platinum environment centered at 532 nm.
January 2014 Highly Transparent Tellurite Glasses and Glass-Ceramics Prepared by Spark Plasma Sintering 165

geometry. The intensity of the SHG signal collected in trans- For the synthesis of glass-ceramics, two mechanisms have
mission geometry is indeed strictly equal to zero due to either been assumed depending on the glass composition. The first
some signal already weak in reflection geometry or to the one associates the densification of the glassy powder through
opacity of the sample resulting from its partial/total devitrifi- viscous sintering to a subsequent devitrification of the
cation. In our SHG measurements, we estimate the error matrix.21 The second possible mechanism implies the densifi-
bars to be around 10%. cation of the matrix through the growth of neck between
glassy particles and concomitantly a gradual crystallization
of the latter. For one given glass composition, the amor-
III. Results and Discussion
phous phase has a considerably lower viscosity than the cor-
(1) 85TeO2–15WO3 Amorphous Powder responding crystalline phase, so that the sintering of
Figure 2(a) shows the particle size distribution of both fine polycrystalline material is inevitably more difficult than its
and coarse 85TeO2–15WO3 amorphous particles. The mean amorphous counterpart. This suggests that the first mecha-
particle diameter (D50) is, respectively, 6 and 167 lm for the nism described above, that is, achieving full density prior to
fine and coarse particles. A broadening of the size distribu- any significant crystallization, is likely at play. However, the
tion is observed in the case of fine particles indicating the glass-ceramics obtained in the TeO2–WO3 system from the
presence of different modes. At the opposite, the size distri- conventional MQ technique and subsequent heat treatment
bution of coarse particles is almost single mode. between Tg and Tc (crystallization temperature) of the amor-
The sintering of glassy particles is generally assumed to phous bulks leads to surface instead of bulk crystallization.20
occur through a viscous flow mechanism just above the glass Therefore, the resulting sintering mechanisms might be more
transition temperature, Tg.21 Therefore, it is believed that the complex.
initial size of the amorphous powder does not influence the The morphology of both fine and coarse particles is repre-
final properties of the samples. In contrary, the initial grain sented in Figs. 2(b) and (c). In the latter case, the particles
size of crystallized particles is of great importance for the display sharp angular edges as shown by the SEM photo-
targeted applications in the process of ceramics sintering graphs.
(alumina, spinel, etc.) where the sintering mechanisms (solid– Figure 3 shows the DSC thermograms recorded for the
solid, solid–liquid, etc.) are different from the ones using 85TeO2–15WO3 fine and coarse particles, as well as that cor-
amorphous powders. responding to the glass bulk of the same composition. The

(a) 9

8 85TeO2-15WO3 fine particle


7 85TeO2-15WO3 coarse particle

5
% volume

–1
1 10 100 1000
Size (μm)

(b) (c)

Fig. 2. (a) Particle size distribution of the initial fine and coarse 85TeO2–15WO3 glass particles. SEM photographs of initial 85TeO2–15WO3
glass particles, (b) fine, (c) coarse.
166 Journal of the American Ceramic Society—Bertrand et al. Vol. 97, No. 1

glass transition temperature, Tg, is constant for the three ranging from 420°C to 435°C, suggesting surface crystalliza-
experiments and is equal to 335°C. The increase in the mean tion as the dominant mechanism in the investigated composi-
particle size from 6 lm (fine) to 167 lm (coarse) causes a tion.22 The broadening of the crystallization peaks for the
shift of the crystallization peaks toward higher temperatures bulk as well as the study of Cß elikbilek et al.20 also confirms
this observation.
To identify the nature of the different crystalline phases
appearing upon heating, Raman spectroscopy data were col-
lected as a function of temperature applying the same heat-
ing rate as the DSC runs, that is, 10°C/min. The Raman
(a) shifts evidenced for these crystalline phases associated with
the different crystallization peaks, respectively, corresponds
to c-TeO2, a-TeO2, and WO3 for the coarse particles and
c-TeO2, WO3, and a-TeO2 for the fine particles (Fig. 4).23

(b)
(2) SPS of the 85TeO2–15WO3 Amorphous Powder
The SPS parameters (pressure, temperature, vertical displace-
ment of punches) during the sintering process are shown in
Fig. 5. The sintering temperature was chosen accordingly to
lie between the glass transition and the first crystallization
temperatures. The sintering process starts in the temperature
(c) range 310°C–325°C and ends around 350°C or 370°C
depending on the pressureless sintering treatment made or
not prior to the SPS experiments. The dwell temperature of
370°C was therefore chosen. Different experiments were car-
Fig. 3. DSC curves of coarse (a), fine (b) 85TeO2–15WO3 glass ried out at constant pressure (50 MPa) and temperature
particles, and (c) glass bulk. (370°C) but different dwell times were applied (ranging from

(a) Heat flow (a.u.)


300

1) TeO2γ 2) 0.35

TeO2 α
350

0.30
WO3
0.25
400
Temperature (°C)

Intensity
0.20
450

(1)
(2) 0.15
(3)
500

0.10

0.05
550

40
0.00
200 400 600 800 1 000
Wavenumber (cm–1)

(b) Heat flow (a.u.)


300

0.30
1) TeO2γ 2)
TeO2 α
350

0.25
WO3
400

0.20
Temperature (°C)

Intensity

(1) 0.15
450

(2) 0.10
500

(3)
0.05
550

40
0.00
200 400 600 800 1 000
Wavenumber (cm–1)

Fig. 4. Temperature-dependent Raman spectroscopy data recorded for coarse (a) and fine (b) 85TeO2–15WO3 glass particles showing the
appearance sequence of the different crystalline phases. The Raman data are compared to the DSC thermograms of the corresponding powder.
January 2014 Highly Transparent Tellurite Glasses and Glass-Ceramics Prepared by Spark Plasma Sintering 167

0 to 10 min) to obtain either a glass bulk material or glass- environment, alumina, tantalum foil, or both. The use of a
ceramics with different crystalline fractions. The amplitude of carbon diffusion barrier listed previously has then a very
the displacement of punches related to the sintering is limited positive effect on the carbon contamination as the bulks turn
in the case of the bulk prepared by pressureless sintering from dark or brownish color to lighter color. However, even
treatment prior to SPS. At the opposite, the shrinkage with the use of a carbon diffusion barrier, many black spots
observed is more important for the “free” powder as also are still visible in the glass bulks. These black spots can be
shown in Fig. 5. The density of SPS amorphous bulk is related to some carbon contamination as no periodic lattice
5.83 g/cm3 which corresponds to more than 99% of the theo- exists in a glass enabling the creation of lattice point defects.
retical density (the reference is the density of a glass obtained SEM observation and Energy-Dispersive X-ray Spectroscopy
by melt quenching). It is noteworthy that the density of the analysis of samples displayed in Figs. 6(d) and (e) lead to the
pressureless sintering pellet prior to SPS is more than 95%. confirmation of the presence of graphite particles [Figs. 7(a)].
Figure 6 shows the photographs of the different glass and Figure 7(b) shows the Raman signatures of both samples
glass-ceramic bulks obtained by SPS from the coarse and fine corresponding to Figs. 6(e) and (f). In particular, for the
particles as well as from the pressureless pellets heat-treated sample depicted in Fig. 6(e), the Raman spectrum has been
at Tg + 30°C for 1 h prior to SPS. Figure 6(a) corresponds recorded after focusing into a black spot, revealing clearly
to the reference glass bulk obtained by the conventional MQ the carbon contamination. Indeed, the Raman spectrum evi-
technique. Figures 6(b)–(e), respectively, show the photo- dences the so-called D and G bands, respectively, at 1380
graphs of glass bulks obtained from fine particles (b and c) and 1590 cm1 reflecting undoubtedly the presence of carbon
and coarse particles (d and e) with or without the presence in materials.24 The origin of these bands will be explained
of some carbon diffusion barrier (carbon comes from the die later in the study.
and/or the papyexâ; Mersen, Paris, France) such as platinum The sintering of fine particles leads to a darker homoge-
neous color than the sintering of coarse particles which
exhibits brownish color with dark spots [Figs. 6(b) and (d)].
This is not fully understood at this point, but we may attri-
bute this phenomenon to the higher number of pores that is
higher in the packing of fine particles which would act as
more carbon diffusing centers through the bulk leading to
the black color in comparison with coarse particles as
explained later in the text (CO gas contained in the pores).
The carbon diffusion kinetic rate is as well higher in the case
of fine particles.
The combination of the pressureless sintering at Tg + 30°C
followed by a SPS treatment permits to eliminate efficiently the
carbon contamination as described by Fig. 6(f). Indeed, the
yellow color is predominant in the glass bulk and the quantity
of black spots is strongly reduced. It is noteworthy that a
single pressureless heat treatment at Tg + 30°C is not sufficient
to get the glass transparency as shown in the insert of Fig. 6,
therefore proving the importance of the SPS treatment.
Fig. 5. Shrinkage rate and load applied as a function of The source of graphite precipitation is believed to be the
temperature during the SPS experiments. atmosphere within the SPS apparatus as well as its graphite
tools that serve as heating element (spacer, die, etc.). This
source can explain the G band in the Raman spectra in
Fig. 7(b).
(a) During heating and densification (even under vacuum or
Ar atmosphere), when an open pore system is present, the
gaseous phase may consist of CO and CO2 which results
from the interaction between carbon and residual oxygen.25
(b) (c) (d) (e) As the pores close up, each pore is filled by this gas, which
undergoes a composition change as a function of tempera-
ture and pressure (due to decreasing pore volume). The
process, which takes place within the closed pores, may be
described by the exothermic reaction25:

2COðgÞ ¼ CO2 ðgÞ þ CðsÞ (1)

with the corresponding equilibrium constant K:

PCO2 xCO2
(f) (g) (h) K¼ ¼ (2)
ðPCO Þ2 x2CO  Ptot

P
Fig. 6. Photographs of (a) melt-quenching (MQ) glass (reference), (ðPtot ¼ PCO þ PCO2 þ Pargon or Pair Þ; xCO ¼ PPCO
tot
; xCO2 ¼ PCO 2
tot
;
(b) and (c) SPS glasses prepared without any pressureless sintering xCO and xCO2 are respectively the molar fraction of CO and
step prior to SPS from fine particles, respectively, without (b) and CO2 species) The forward reaction of Eq. (1) is favored at
with (c) a carbon diffusion barrier, (d) and (e) SPS glasses elaborated low temperatures, that is, below 700°C. Then, during viscous
without any pressureless sintering step from coarse particles, sintering, the diffusion of gas (especially CO) out of the pore
respectively, without (d) and with (e) a carbon diffusion barrier, (f)
(g), and (h) SPS glass and glass-ceramics prepared with a
into the glass is rapid relative to the surface-tension-induced
pressureless sintering step prior to SPS (from coarse particles). The viscous flow of material.26 Therefore, with increasing densifi-
inset highlights the strong impact of the SPS step for the cation, the total pressure Ptot within the pores decreases as
transparency. the volume of the pores decreases due to the gas dissolution
168 Journal of the American Ceramic Society—Bertrand et al. Vol. 97, No. 1

(a)

Te
Glassy
matrix
C particle
Te

c w Te
o Te
Te
5 μm

0 1 2 3 4 5
KeV

(b) SPS glass (with pressure-less sintering prior to SPS)


SPS glass (without pressure-less sintering prior to SPS)
Average Raman intensity (a.u.)

G band

D band

900 1000 1100 1200 1300 1400 1500 1600 1700 1800 1900
Wavenumber (cm–1)
Fig. 7. (a) SEM image and EDXS analysis of a carbon particle embedded in the glassy matrix, (b) Raman spectra of glass bulks sintered by
SPS with and without pressureless sintering prior to SPS.

4
Intensity (arb. units)

(e)

2 (d)

(b)
(c)
(a)
(b)

(a)
Fig. 8. Raman spectra collected for glass bulks prepared (a) by 0
melt quenching (MQ), (b) by SPS combined with a pressureless 20 25 30 35 40
sintering step prior to SPS. 2θ CuKα1 (°)

Fig. 9. XRD patterns collected for various samples sintered by


SPS: (a) glass (370°C, 50 MPa, 0 min), (b)–(e) glass-ceramics
into the viscous flow. The precipitation of carbon at the free sintered by SPS (370°C, 50 MPa, from 1 to 10 min). The symbols ✱
surface of the pores may therefore take place at the interface and ○ correspond, respectively, to the presence of c-TeO2 and
gas–glass. This source of carbon could be the signature of a-TeO2 phases.
January 2014 Highly Transparent Tellurite Glasses and Glass-Ceramics Prepared by Spark Plasma Sintering 169

both the G bands and D bands in the Raman spectra in become more and more opaque [Figs. 6(g) and (h)].
Fig. 7(b). The disordered carbon characterized by the D Figure 6(h) even shows the aspect of a sample that was almost
bands could result from the combined presence of the viscous completely crystallized.
flow during the sintering and the variation in pressure inside For the following part of this study, we will focus on the
the closed porosity. glasses and glass-ceramics made by a pressureless sintering
A pressureless sintering prior to SPS at Tg + 30°C in a step prior to SPS followed by an SPS experiment from coarse
conventional furnace with no graphite environment allows particles which are of particular interest as they do not
first the densification of the sample (>95%) and reducing the present any detectable and visual carbon contamination.
amount of CO and CO2 trapped in the closed porosity due
to the nongraphite environment. It is noteworthy that graph-
ite pollution in refractory oxide ceramics or glasses can be (3) Characterizations of Glass and Glass-Ceramics
easily suppressed by a thermal treatment above 600°C.27 In Prepared by SPS
our case, the very low Tg of tellurite glasses prevents any The structure of the glasses made by melt quenching
treatment at such a temperature. [Fig. 6(a)] and SPS [Fig. 6(f)] have been compared by Raman
Increasing the dwell time at constant pressure and temper- spectroscopy and a perfect match between the two spectra is
ature during the SPS experiments leads to glass-ceramics that shown in Fig. 8. This information is crucial to prove that

(a) (b) (c)

100 µm 100 µm 100 µm


(d) (e)

100 µm 20 µm

(f)
-200
0.30
-150

0.25
-100

-50 0.20
Intensite (coups)

(1)
Y (µm)

0 0.15
(2)
50 0.10

100
0.05
50 µm
150
10 µm 0.00
200
–200 –100 0 100 200
X (µm)

1,4
0,3 (1) (2)
1,2
I (arbritrary unit)

0,25
I (arbritrary unit)

1
0,2
0,8
0,15 0,6
0,1 0,4
0,05 0,2
0
0 300 400 500 600 700 800 900
300 400 500 600 700 800 900
wavenumber (cm–1)
wavenumber (cm–1)

Fig. 10. (a, b, c, and d) EDXS-SEM pictures of glass and glass-ceramics prepared by SPS (370°C, 50 MPa, dwell time of, respectively, 0, 1, 3,
and 5 min), (e) Higher magnification of the picture displayed in (d), (f) Raman mapping of a glass-ceramic prepared by SPS (370°C, 50 MPa,
5 min) associated with the optical image and the corresponding spectra. (1) and (2) correspond to the labels denoting, respectively, the core or
the surface of the glass particles.
170 Journal of the American Ceramic Society—Bertrand et al. Vol. 97, No. 1

SPS experiments do not modify the structure of this tellurite these samples, the effect of light scattering produced by the
glass. The Raman shifts evidenced in these spectra are sum- c-TeO2 crystals is mainly due to their 2D dendritic shape
marized in Fig. 8. Further details about the structure of this [Fig. 10(e)].
glass composition can be found in.28 It is noteworthy that the broad absorption band around
XRD measurements were carried out first to establish the 3300 nm is attributed to the presence of OH groups intro-
crystalline/amorphous nature of the SPS samples. The XRD duced by the water content of the glass.
pattern depicted in Fig. 9(a) is obviously typical of an amor- As testified by both XRD and Raman spectroscopy data
phous sample, whereas Figs. 9(b)–(e) represent XRD pat- [Figs. 9 and 10(f)], the crystals located at the surface of
terns characteristic of glass-ceramics. As mentioned above, amorphous grains and disseminated within the whole volume
the latter XRD pattern [Fig. 9(e)] even corresponds to a SPS of the glass-ceramics correspond to the c-TeO2 phase. The
sample that was almost completely crystallized. The well- latter crystallizes into an orthorhombic unit cell with the
defined Bragg peaks showing up correspond to the crystalli- P212121 space group and is, as a consequence, noncentrosym-
zation of the c-TeO2 phase (marked by ✩). Additional peaks metric. Therefore, for SPS glass-ceramics, it is expected to
appearing for longer dwell time (marked by ○) corresponds detect some SHG signal.
to the a-TeO2 phase. Four SPS glass-ceramics samples were tested: their corre-
As previously mentioned, surface crystallization occurs sponding XRD patterns are represented in Figs. 9(b)–(e),
upon heating the glass bulks obtained by the conventional respectively, and the pictures of two of them are displayed in
MQ technique between Tg and Tc, thus preventing this com- Figs. 6(g) and (h). Thus, from the XRD data, it is obvious
position to be used in any potential optical applications.20 In that the crystalline fraction progressively increases.
our present study, crystals are found to be disseminated Figure 12(a) displays the variation in the SHG amplitude
within the whole volume of the glass-ceramics prepared by as a function of the approximate crystalline volume fraction
SPS from amorphous powders. To evaluate which sintering estimated from the processing of EDXS–SEM images and
mechanism is likely at play and to study the microstructure, described before. Considering that the shape (according to
we observed by SEM analysis some various freshly polished SEM observations) and size (based on similar full-width at
glass-ceramics with different crystalline fraction obtained by half maximum values) of crystals remain more or less con-
varying the dwell time during the SPS experiments. stant, the SHG amplitude should then be proportional to the
Figure 10(a) corresponds to the amorphous bulk of Fig. 6(f) volume fraction of crystals. The experimental observation
and serves as a reference. Figures 10(b)–(e) reveal a chemical clearly confirms this statement: a linear fit [Fig. 12(a)] is pro-
contrast at the grain boundaries characteristics of the surface vided as a guide for the eyes to highlight this tendency. Only
devitrification phenomenon leading to the coast-and-island the first three points were fitted, as they correspond to glass-
microstructure.29,30 Such phenomenon can be confirmed ceramic samples containing uniquely c-TeO2 crystals. The
based on Raman mapping data displayed in Fig. 10(f) which last point corresponds to the sample which is almost com-
unambiguously demonstrate the localization of crystals (c- pletely crystallized (i.e., such point was placed at an abscissa
TeO2) at the grain boundaries, whereas the grains remain equal to 100%), and where a-TeO2 and c-TeO2 crystals are
amorphous. Moreover, SEM photographs [Figs. 10(b)–(d)] both present (see Fig. 9). The SHG of the base glass is equal
perfectly corroborate the XRD data (Fig. 9) by showing to zero and is not represented. As the SHG efficiency is dif-
that, as the crystalline fraction increases when longer dwell ferent between a-TeO2 and c-TeO2 crystals, it is unfortu-
time is applied during SPS, the number of crystals increases nately impossible to include this last data in the fitting
at the amorphous particle boundaries. As a first estimation, procedure.
only the relative part the crystalline domain was determined The threshold for detecting SHG is difficult to define accu-
using image analysis software on these EDXS–SEM pictures. rately, but can be, however, estimated. Indeed, the linear fit
The percentage obtained should therefore be regarded as an to the data should go through the origin of the graph (as for
upper bound of the actual crystalline fraction due to the den-
dritic shape of crystals that are embedded in the glass.
Within this rough approximation, the results indicate an
increase in the crystalline fraction from 3% [Fig. 10(b)], 5%
[Fig. 10(c)] to 40% [Fig. 10(d)] as the dwell time increases
up to 5 min under the same temperature and pressure
conditions.
The data were collected for three samples elaborated by
SPS and they were compared to the optical transmission
curve measured for one reference glass sample elaborated by
the conventional MQ technique. First, we started comparing
the optical transmission curve measured for a MQ glass with
that measured for a typical glass made by SPS. Based on the
Raman spectroscopy data commented in Fig. 8, which indi-
cate that the structure of (MQ) and SPS glass samples is
identical, the reduction in the optical transmission in the
short wavelengths noticed for a SPS glass sample is attrib-
uted to light scattered by residual porosity, small inclusions/
crystals or residual pollution. However, despite this reduc-
tion, the optical transmission value appears already high (up
to 70% transmission) for these first tellurite glasses prepared
by SPS.
The formation of c-TeO2 crystals at the surface of amor-
phous grains in the very first steps of the sintering process
partially hinders further densification of the samples due to
Fig. 11. Visible and near-IR optical transmission curves of (a) glass
viscosity considerations, explaining the scattering of light evi- bulk obtained by the conventional melt-quenching (MQ) technique
denced in the optical transmission spectra represented in (reference), (b) glass bulk obtained by pressureless sintering step
Fig. 11. prior to SPS + SPS (370°C, 50 MPa, 0 min), and (c, d) two glass-
For glass-ceramics SPS samples, the optical transmission ceramics obtained by pressureless sintering step prior to SPS + SPS
drops with the increasing crystalline volume fractions. For (370°C, 50 MPa and two different dwell times).
January 2014 Highly Transparent Tellurite Glasses and Glass-Ceramics Prepared by Spark Plasma Sintering 171

(a) with K′ being another constant including the previous con-


stants. Thus, theoretically, if the observed physical phenom-
enon is actually SHG, the curve representing the logarithm
of the SH power versus the logarithm of the fundamental
power must be a straight line with a slope strictly equal to
two.
For the sample displaying the highest signal, Fig. 12(b)
indicates that the log–log curve can be clearly fitted by a
straight line (the goodness of fit parameter r2 = 0.9989,
whereas a perfect fit gives r2 = 1) with a slope equal to 2.03.
The experimental value matches extremely well that of the
theory, demonstrating by this way that the detected signal
actually corresponds to SHG and is not due to any residual
fluorescence.

IV. Conclusion
Glasses and glass-ceramics based on the 85TeO2–15WO3
composition have been prepared through the SPS technique.
Several parameters have been studied to limit the carbon
(b) contamination that is common with this nonconventional
technique and which is detrimental for optical applications.
The influence of the initial amorphous particles size as well
as the effect of a pressureless sintering step prior to SPS has
been investigated. The carbon contamination is higher in the
case of the fine particles leading to dark glass bulks. The use
of a diffusion barrier such as platinum, tantalum foil or alu-
mina has a limited effect on the carbon pollution. The best
results were obtained by a pressureless sintering step around
Tg prior to SPS experiment. Yellow glasses with relatively
high optical transparency properties were obtained which
enlighten the positive influence of such a pressureless sinter-
ing step. The carbon contamination is mainly due to graphite
environment (die, papyexâ) which diffuses into the material
and which also lead to residual CO or CO2 inside the SPS
chamber even operating under argon atmosphere. By increas-
ing the dwell time at constant pressure and temperature dur-
Fig. 12. (a) Evolution of the SHG amplitude detected for the ing the SPS experiments, the noncentrosymmetric c-TeO2
various samples tested within this study, as a function of the volume phase crystallized through the matrix which leads to the gen-
fraction of crystals. The linear fit of the first three points serves only eration of a SH signal.
as a guide for the eyes. The dashed line corresponds to the Further experiments will be now carried out to improve
“continuity” of the linear fit. The last point is not fitted and the the optical performances by reducing the proportion of scat-
SHG of the base glass equal to zero is not represented, (b) log tered light induced by residual porosity or graphite particles.
(second harmonic (SH) intensity)–log (pump laser intensity) plot.
The straight line corresponds to the linear fit of the data.
Acknowledgments
a volume fraction of crystals equal to zero, there is abso- Authors would like to thank Dr. Claude Godart and Dr. Judith Monnier from
lutely no SHG). It is almost, but not exactly, what we ICMPE, Thiais (France) as well as Dr. Alain Largeteau from ICMCB,
observe. Thus, the slight shift evidenced corresponds to the Bordeaux (France) for some SPS experiments. This work is supported by the
Labex Sigma-Lim (no. ANR-10-LABX-0074-01).
SHG amplitude detected at the ordinate at the origin. Such
value can be directly related to a volume fraction approxi-
mately equivalent to 1.5%, which constitutes the volume References
fraction of crystals beyond which it will become possible to 1
D. Milanese, H. Gebavi, J. Lousteau, M. Ferraris, A. Sch€ ulzgen, L. Li, N.
detect SHG. It is finally important to undermine that the Peyghambarian, S. Taccheo, and F. Auzel, “Tm3+ and Yb3+ Co-Doped Tel-
SHG signal measured for a glass sample (Figs. 6(f)–10(a)] is lurite Glasses for Short Cavity Optical Fiber Lasers: Fabrication and Optical
strictly equal to zero. Characterization,” J. Non-Cryst. Solids, 356 [44–49] 2378–83 (2010).
2
As the SHG phenomenon is a second-order nonlinear pro- R. A. H. El-Mallawany, Tellurite Glasses Handbook, Physical Properties
and Data. 2nd edition, CRC press, Boca Raton, FL, 2012.
cess, the dependence of the SH energy must be strictly a 3
T. Komatsu, H. Tawarayama, H. Mohri, and K. Matusita, “Properties and
quadratic function of the fundamental energy.31 In,32 the Crystallization Behaviors of TeO2–LiNbO3 Glasses,” J. Non-Cryst., 135 [2–3]
authors have expressed the SH power, P (2x), as follows: 105–13 (1991).
4
G. Vrillet, C. Lasbrugnas, P. Thomas, and O. Masson, “Efficient Second
Harmonic Generation in c-TeO2 Phase,” J. Mater. Sci., 41, 305–7 (2006).
K:l2c :P2 ðxÞ 5
T. Berthier, V. M. Fokin, and E. D. Zanotto, “New Large Grain, Highly
Pð2xÞ ¼ (3) Crystalline, Transparent Glass-Ceramics,” J. Non-Cryst. Solids, 354, 1721–30
A (2008).
6
M. Allix, S. Alahrache, F. Fayon, M. Suchomel, F. Porcher, T. Cardinal,
where lc is the coherence length, A the area beam, K a mate- and G. Matzen, “Highly Transparent BaAl4O7 Polycrystalline Ceramic
rial-dependent constant proportional to d2(2x) (d being the Obtained by Full Crystallization from Glass,” Adv. Mater., 2–4, 5570–5
(2012).
second-order NLO susceptibility), and P(x) the fundamental 7
E. D. Zanotto, “Experimental Studies of Surface Nucleation and Crystalli-
power. If one applies the logarithmic function to Eq. (3), it zation of Glass”; pp. 65–74 in Ceramic Transactions, Nucleation and Crystalli-
turns out that: zation in Liquids and Glasses, Vol. 30, Edited by M. C. Weinberg. American
Ceramic Society, Westerville, OH, 1992.
8
M. Avrami, “Kinetics of Phase Change. II. Transformation-Time Relations
log Pð2xÞ ¼ logðK0 Þ þ 2: log PðxÞ (4) for Random Distribution of Nuclei,” J. Chem. Phys., 8 [2] 212–24 (1940).
172 Journal of the American Ceramic Society—Bertrand et al. Vol. 97, No. 1
9 22
T. Ozawa, “Kinetics of Non-Isothermal Crystallization,” Polymer, 12, V. M. F. Marques, D. U. Tulyaganov, S. Agathopoulos, and J. M. F.
150–8 (1971). Ferreira, “Low Temperature Production of Glass Ceramics in the Anorthite-
10
H. E. Kissinger, “Variation of Peak Temperature with Heating Rate in Diopside System via Sintering and Crystallization of Glass Powder
Differential Thermal Analysis,” J. Res. Natl. Bur. Stand., 57 [4] 217–21 (1956). Compacts,” Ceram. Int., 34 [3] 1145–52 (2008).
11 23
W. Sack, Beitrage zur Angewandten Glasforschung, Edited by E. Schott. A. P. Mirgorodsky, T. M. Mejean, J. C. Champarnaud Mejean, P.
pp. 111, Universitat Mainz, Mainz, Germany, 1959. Thomas, and B. Frit, “Dynamics and Structure of TeO2 Polymorphs: Model
12
M. Hubert, G. Delaizir, J. Monnier, C. Godart, H. L. Ma, X. H. Zhang, Treatment of Paratellurite and Tellurite; Raman Scattering Evidence for New
and L. Calvez, “An Innovative Approach to Develop Highly Performant Chal- Gamma- and Delta-Phases,” J. of Phys. Chem. Sol, 61 [4] 501–9 (2000).
24
cogenide Glasses and Glass-Ceramics Transparent in the Infrared Range,” Y. Wang, D. C. Alsmeyer, and R. L. McCreery, “Raman Spectroscopy of
Opt. Express, 19 [23] 23513–22 (2011). Carbon Materials: Structural Basis of Observed Spectra,” Chem. Mater., 2,
13
S. Nowak, L. Perriere, L. Dembinski, S. Tusseau-Nenez, and Y. Cham- 557–63 (1990).
25
pion, “Approach of the Spark Plasma Sintering Mechanism in Zr57Cu20Al10 S. Meir, S. Kalabukhov, N. Froumin, M. P. Dariel, and N. Frage, “Syn-
Ni8Ti5 Metallic Glass,” J. Alloy. Compd., 509 [3] 1011–9 (2011). thesis and Densification of Transparent Magnesium Aluminate Spinel by SPS
14
G. Bernard-Granger, N. Benameur, C. Guizard, and M. Nygren, “Influ- Processing,” J. Am. Ceram. Soc., 92 [2] 358–64 (2009).
26
ence of Graphite Contamination on the Optical Properties of Transparent D. H. Eggler, B. O. Mysen, T. C. Hoering, and J. R. Holloway, “The Sol-
Spinel Obtained by Spark Plasma Sintering,” Scripta Mater., 60, 164–7 (2009). ubility of Carbon Monoxide in Silicate Melts at High Pressures and its Effect
15
L. Ramond, G. Bernard-Granger, A. Addad, and C. Guizard, “Sintering on Silicate Phase Relations,” Earth Planet. Sci. Lett., 43 [2] 321–30 (1979).
27
of Soda-Lime Glass Microspheres Using Spark Plasma Sintering,” J. Am. T. G. Mayerhofer, Z. Shen, E. Leonova, M. Eden, A. Kriltz, and J. Popp,
Ceram. Soc., 94 [9] 2926–32 (2011). “Consolidated Silica Glass from Nanoparticles,” J. Sol. State Chem., 181 [9]
16
B. Xue, L. Calvez, V. Nazabal, X. H. Zhang, G. Delaizir, J. Monnier, G. 2442–7 (2008).
28
Martinelli, and Y. Quiquempois, “Mechanical milling and SPS used to obtain A. Mirgorodski, M. Colas, M. Smirnov, T. Merle Mejean, R. El Mallaw-
GeS2-βGeS2 infrared glass-ceramics,” J. Non-Cryst. Sol., 377, 240–4 (2013). any, and P. Thomas, “Structural Peculiarities and Raman Spectra of TeO2/
17
C. Wang and Z. Zhao, “Transparent MgAl2O4 Ceramic Produced by WO3-Based Glasses: A Fresh Look at the Problem,” J. Sol. State Chem., 190,
Spark Plasma Sintering,” Scripta Mater., 61, 193–6 (2009). 45–51 (2012).
18 29
M. L. Ovecoglu, G. Ozen, S. Cenk, “Microstructural Characterization W. H€ oland, M. Schweiger, M. Frank, and V. Rheinberger, “A Compar-
Crystallization Behaviour of (1x)TeO2–XWO3 (x = 0.15, 0.25 and 0.3% mol) ison of the Microstructure and Properties of the IPS Empress 2 and the
Glasses,” J. Eur. Ceram. Soc., 26, 1149–58 (2006). IPS Empress Glass-Ceramics,” J. Biomed. Mater. Res., 53 [4] 297–303
19
S. Blanchandin, P. Marchet, P. Thomas, J. C. Champarnaud-Mesjard, and (2000).
30
B. Frit, “New Investigations Within the TeO2 WO3 System: Phase Equilibrium A. Karamanov, G. Taglieri, and M. Pelino, “Iron-Rich Sintered Glass-
Diagram and Glass Crystallization,” J. Mater. Sci., 34, 4285–92 (1999). Ceramics from Industrial Wastes,” J. Am. Ceram. Soc., 82 [11] 3012–6 (1999).
20
M. C ß elikbilek, A. E. Ersundu, N. Solak, and S. Aydin, “Crystallization 31
P. N. Butcher and D. Cotter, The Elements of Nonlinear Optics.
Kinetics of the Tungsten–Tellurite Glasses,” J. Non-Cryst. Solids, 357 [1] Cambridge university press, Cambridge, UK, 1990.
32
88–95 (2011). J. P. Dougherty and S. K. Kurtz, “A Second Harmonic Analyzer for the
21
M. N. Rahman, Sintering of Ceramics. p. 47, CRC Press, London, 2008. Detection of Non-Centrosymmetry,” J. Appl. Cryst., 9, 145–58 (1976). h

You might also like