Vous êtes sur la page 1sur 12

Journal of Non-Crystalline Solids 244 (1999) 211222

FTIR and XPS studies of P2O5Na2OCuO glasses


P.Y. Shih
b

a,b,*

, S.W. Yung b, T.S. Chin

a
Department of Materials Science and Engineering, National Tsing Hua University, Hsinchu 30043, Taiwan, Republic of China
Department of Ceramic Engineering, National Lien Ho College of Technology and Commerce, 1 Lein-Kun, Kung-Jing Li, Miaoli 36012,
Taiwan, Republic of China

Received 4 May 1998; received in revised form 9 November 1998

Abstract
The short range structure of ternary sodiumcopperphosphate [(4070)P2 O5 (060)Na2 O(050)CuO (mol%)]
samples were examined by Fourier-transform infrared (FTIR) spectroscopy and X-ray photoelectron spectroscopy
(XPS). The XPS O 1s spectrum was deconvoluted into three components, symmetric bridging oxygen (POP), nonbridging oxygen (
and P@O) and asymmetric bridging oxygen (POCu). In the ultraphosphate and metabonds while keeping the same fraction of PO
phosphate glasses, the POCu bonds are formed replacing
P bonds as Na2 O is replaced by CuO. Whereas, in the polyphosphate glasses, the formation of POCu will replace
both POP and
bonds. The results reveal the formation of POCu bonds in these glasses, and an increase
of crosslink density. Thus, chemical durability and Tg of the glasses are increased. 1999 Elsevier Science B.V. All
rights reserved.

1. Introduction
Phosphate glasses usually melt at temperatures
less than 1000C [16], have thermal expansion
coecients (a) in the range of 90 to 250 107 /C
[46], have glass transition temperatures (Tg ) less
than below 350C and softening temperature (Td )
below 400C [4,5]. Thus, they are of increasing
interest for many applications, e.g. glass to metal
seals, thick lm paste, the molding of optical elements, low temperature enamels for metals [24].
Their relatively poor chemical durability makes
them generally unsuitable for practical applications [79]. It was suggested [915] that the addition of one or more of SnO, PbO, ZnO, Al2 O3 and

*
Corresponding author. Tel.: +886-37 355028; fax: +886-37
324047; e-mail: byshih@mail.lctc.edu.tw

Fe2 O3 , results in the formation of SnOP, PbO


P, ZnOP, POAl and POFe bonds, and leads
to improvement in the chemical durability of the
modied phosphate glasses. Some glass compositions have a greater durability than others.
By doping proper cations or by nitridation,
crosslink density of the glass network can be increased which improves chemical durability
[7,8,1618]. The improved chemical durability is
often accompanied by a decrease in the thermal
expansion coecient and an increase in the glass
softening temperature [710]. Of all the additions
made to phosphate glasses to improve their durability, PbO is the only one which reduces the dissolution rate and softening temperature at the
same time [4,5,19].
In recent years many investigators have reported on the electrical properties of copper
phosphate binary glasses [2023]. Salim et al. [23],

0022-3093/99/$ see front matter 1999 Elsevier Science B.V. All rights reserved.
PII: S 0 0 2 2 - 3 0 9 3 ( 9 9 ) 0 0 0 1 1 - 3

212

P.Y. Shih et al. / Journal of Non-Crystalline Solids 244 (1999) 211222

used X-ray photoelectron spectroscopy (XPS) and


Fourier-transform infrared (FTIR) spectrophotometer to study the structure of binary copper
phosphate glasses. It was suggested that POCu
bond may exist in the glass network. Nevertheless,
no direct evidence was oered to conrm the
presence of POCu bonds in their work [23]. The
electrical and optical properties of the coppersodium metaphosphate glass have been studied [24].
A systematic study of the properties of the ternary
sodiumcopperphosphate glass system has been
reported [25]. However, no systematic studies on
the structure/property relation of these ternary
glasses have been published.
In our previous investigations, properties of
sodiumcopperphosphate glasses containing 40
80 mol% P2 O5 , 060 mol% Na2 O and 050 mol%
CuO were prepared and properties measured [25].
The compositional dependence of thermal and
corrosion properties (glass transition temperature,
Tg ; glass softening temperature, Td ; coecient of
thermal expansion, a; and chemical durability) was
explored, as shown in Table 1. The substitution of

CuO for Na2 O increases Tg , Td , chemical durability, and decreases a of sodiumcopperphosphate glasses.
We suggested [25] that copper may exist in
glass-forming oxide melts as cuprous (Cu ) and
cupric (Cu2 ) ions [22,23,26,27]. Besides, the ratio
[Cu2 ]/[Cutotal ] ([Cutotal ] is the total copper in the
glasses, that is [Cu ]+[Cu2 ]) of binary copper
phosphate glasses may aect the property and
structure of the glasses [21,22,28]. The [Cu2 ]/
[Cutotal ] ratio in a glass is known to depend on the
composition and thermal history (melting temperature, melting time, atmosphere, quenching
rate, and annealing conditions). We assume that
these aect the properties of sodiumcopper
phosphate glasses, and we are interested in
studying the relationship among thermal history,
[Cu2 ]/[Cutotal ] ratio, property and structure of the
ternary sodiumcopperphosphate glasses.
FTIR and XPS are important tools to provide
short range structural information about phosphate glasses [14,15,23,2830]. Particularly, it is
possible for O 1s XPS spectrum to distinguish the

Table 1
Properties of the P2 O5 Na2 OCuO glasses
Tg (3C)

Td (5C)

Thermal expansion
coecient
(5 107 /C)

Dissolution rate
(g/cm2 min)
(10%)

60P2 O5 (40x)Na2 OxCuO


0
245
10
322
20
353
30
364
40
377

212
252
272
291
330

231
281
302
322
352

243
228
182
168
148

1.5
2.2
7.8
6.6
5.9

104
106
107
107
107

50P2 O5 (50x)Na2 OxCuO


0
243
10
271
20
309
30
339
40
363
50
395

258
278
294
327
358
420

276
297
316
353
383
432

254
232
191
178
148
99

2.6
3.4
9.1
9.0
3.9
1.9

103
104
105
106
106
106

40P2 O5 (60x)Na2 OxCuO


0
256
10
289
20
318
30
357
40
391

, 361
298g1 , 374g2
324,
347,
372,

, 383
325d1 , 400d2
344,
370,
393,

258
243
240
202
173

4.7
3.3
6.8
3.6
2.5

103
106
106
106
107

Glass sample
CuO (x) (mol%)

Hardness VHN
(5 kg/mm2 )

The superscripts g1, g2, d1 and d2 represent Tg1 , Tg2 , Td1 and Td2 respectively.

P.Y. Shih et al. / Journal of Non-Crystalline Solids 244 (1999) 211222

bridging oxygen from the non-bridging oxygen.


Hence, it is useful for analyzing the structural
changes in such glasses. In the present study, we
systematically examine the FTIR and XPS O 1s
spectra of the ternary sodiumcopperphosphate
glass system.

2. Experimental procedures
2.1. Glass preparation
Sodiumcopperphosphate samples were prepared from mixtures of reagent grade sodium
dihydrogen phosphate (NaH2 PO4 ), sodium carbonate (Na2 CO3 ), copper oxide (CuO) and ammonium dihydrogen phosphate (NH4 H2 PO4 ). The
batch compositions of the samples are listed in
Table 1. The mixed powders were heated at 300C
for 1 h in alumina crucibles to evaporate ammonia
and water in the batch. Then batches were melted
in air for 30 min at 7001100C, depending on
glass composition. The liquids were cast onto a
stainless steel plate and annealed for 1 h at about
10C above Tg , as determined by dierential
thermal analyses (from 220C to 420C, depending on the composition [25]). The glass samples were kept in a desiccator to prevent possible
reactions with moisture.
Powder X-ray diraction (XRD) was used to
examine all the as-quenched melts to conrm the
amorphous state of the samples. The properties of
these glasses, such as Tg , Td , a, and chemical durability were measured and listed in Table 1. The
analyzing methods of the properties and the uncertainty of each measurement are described and
discussed in Ref. [25].
2.2. FTIR analyses
FTIR spectra were recorded with a FTIR
spectrometer (Jasco 300E) in the frequency range
4001400 cm1 at room temperature. Pellets were
prepared for FTIR measurements by mixing and
grinding a small quantity of glass powder with
spectroscopic grade dry KBr powder and then
compressing the mixtures to form pellets for

213

measurements. All measurements were at 2 cm1


resolution.
2.3. XPS analyses
A spectrometer (V.G. Scientic EscaLab 210)
with 300 W Mg Ka (1253 eV) X-ray as the excitation radiation was used for the XPS measurements. The spectra were collected in a xed
retarding ratio mode with a bandpass energy of
about 10 eV. Disc samples were polished on one
side to optical atness using kerosene and emery
papers, and were rinsed in ethanol for several
minutes prior to XPS measurements. XPS spectra
of O 1s core level were recorded. Drift of the
electron binding energy due to a surface charging
eect was calibrated by utilizing the C 1s peaks of
the contamination from the pumping oil
(EB 284.6 eV). Experimental uncertainty of XPS
binding energy was about 0.1 eV. Each O 1s
spectrum was tted with Gaussian functions (two
peaks for binary sodiumphosphate samples, three
peaks for ternary sodiumcopperphosphates) by
means of a least-squares tting process. Peak positions, amplitudes and the full width at half
maximum amplitude were varied to attain the best
t to each experimental spectrum. More than one
sample of each glass was analyzed in this manner
and the quantitative oxygen bonding analysis
(based on relative peak areas) were reproducible to
5%.
3. Results
The glass state was conrmed by the absence of
XRD peaks. All glasses listed in Table 1 are
transparent. The appearance is blue due to the
presence of copper ions in the ternary system, and
turns darker with increasing CuO content. In the
glass system with 40 mol% P2 O5 , the melts devitried when CuO content exceeded 40 mol%. Fig. 1
is the XRD pattern of a 40P2 O5 10Na2 O50CuO
composition. The main crystalline phase is crystalline Cu2 P2 O7 , according to the JCPDS card, le
No. 21880.
Analyses of the physical and chemical properties of P2 O5 Na2 OCuO samples taken from the

214

P.Y. Shih et al. / Journal of Non-Crystalline Solids 244 (1999) 211222

Fig. 1. X-ray powder diraction pattern of the quenched/annealed 40P2 O5 10Na2 O50CuO glasss. The diraction peaks
corresponding to crystalline Cu2 P2 O7 phase.

same specimens as these samples were presented in


our previous paper [25]. Properties of these samples are summarized in Table 1. The glass transition temperatures (Tg ) and dilatometric softening
temperatures (Td ) are a function of glass composition, and decrease with increasing P2 O5 content,
and increase with increasing CuO content. In the
polyphosphate system with 40 mol% P2 O5 , there
are two Tg s (Tg1 and Tg2 ) and Td s (Td1 and Td2 )
corresponding to two glass states, as depicted by
the dierential thermal analyses (DTA) of the
samples [25]. The thermal expansion coecient
and dissolution rate of P2 O5 Na2 OCuO samples
decrease with increasing CuO contents. These results indicate that the substitution of CuO for
Na2 O in the phosphate glasses will increase the
structural connectivity in the samples.
3.1. FTIR analyses
Fig. 2 shows FTIR spectra of 40P2 O5
(60x)Na2 OxCuO glasses. The glasses containing 40 mol% P2 O5 have an analogous FTIR
spectra. The band near 1270 cm1 is assigned to
asymmetric stretching mode of the two nonbridging oxygen atoms bonded to phosphorus atoms, the OPO or (PO2 )as units, in the phosphate
tetrahedra [3133]. The band near 1190 cm1 is
assigned to the PO2 symmetric stretching mode,
(PO2 )s [31,34], and their amplitudes decrease with
increasing CuO content. The absorption bands

Fig. 2. FTIR spectra of the 40P2 O5 (60x)Na2 OxCuO


glasses: (a) x 0, (b) x 10, (c) x 20, (d) x 30, (e)
x 40 mol%.

near 1100 and 1000 cm1 have been assigned to P


O groups (chain terminator) [29]. The PO absorption bands near 1100 cm1 shifts to lower
frequency as CuO replaces Na2 O. The band near
1030 cm1 , which is attributed to PO3 end groups
[35], tends to decrease with the substitution of

P.Y. Shih et al. / Journal of Non-Crystalline Solids 244 (1999) 211222

CuO for Na2 O. The absorption bands near 900


and 775 cm1 are assigned to the asymmetric and
symmetric stretching modes of the POP linkages, respectively. The asymmetric stretching band of
POP near 900 cm1 shifts to lower frequency
initially as Na2 O is replaced by CuO, then shifts to
a higher frequency when the CuO content exceeds
20 mol%. Besides, its amplitude decreases with
increasing CuO content, especially as CuO content
exceeds 20 mol%. The shift and the intensity
change of POP band are correlated with the
changes in properties in this glass system, as shown
in Table 1. It has been indicated that Tg1 (belonging to the low Tg glass phase) will appear in
the glass as Na2 O is replaced by CuO, and Tg2
(belonging to the high Tg glass phase) disappears
in the glass containing 20 mol% CuO [25]. The
bands around 530 cm1 are attributed to deformation modes of PO (PO3
4 ) groups [23,34,35].
In addition, there is no evidence for the existence of bands due to POCu linkage. We assume
that the stretching band corresponding to POCu
linkage might be near 1100 cm1 similar to the P
OSn linkage in SnPOF glasses proposed by
Day et al. [36], or it may be buried in the dominant
PO stretching band which has a width sucient
to bury the band. The overlapping of absorption
bands make their assignment dicult.
FTIR absorption spectra of 50P2 O5
(50x)Na2 OxCuO glasses in the frequency range
between 400 and 1400 cm1 are shown in Fig. 3.
There is no dierence in the line shapes of the
spectra among samples containing 50 mol% P2 O5 ,
however, the amplitudes of the bands vary with
composition.
Compared with Fig. 2, the most noticeable
change in the FTIR spectra is that the amplitude
of the (PO2 )as stretching band near 1270 cm1 increases as the P2 O5 content increases from 40 to 50
mol%. These bands shift to lower frequency and
their amplitudes decrease as Na2 O is replaced by
CuO. The band at 1160 cm1 is assigned to the
PO2 symmetric stretching mode, (PO2 )s [29,32].
The band at 1160 cm1 for the (PO2 )s is not well
resolved in Fig. 3 except in the spectrum of
Fig. 3(e). The absorption band for PO groups
near 1100 cm1 obviously shifts to lower frequency
as CuO replaces Na2 O. The absorption band near

215

Fig. 3. FTIR spectra of the 50P2 O5 (50x)Na2 OxCuO


glasses: (a) x 10, (b) x 20, (c) x 30, (d) x 40, (e)
x 50 mol%.

900 cm1 is assigned to the asymmetric stretching


mode of the POP linkages, while two modes
around 725 and 785 cm1 are attributed to the
symmetric stretching of POP groups [37,38].
Both bands for POP asymmetric and symmetric
stretching shift to higher frequency with the substitution of CuO for Na2 O.
FTIR spectra of ultraphosphate glasses with 60
and 70 mol% P2 O5 are shown in Figs. 4 and 5,
respectively. The absorption band near 1320 cm1
is attributed to the stretching mode of P@O double
bonds, which are only present in the glasses with a
P2 O5 content > 50 mol% [33,39]. It is observed in
all these spectra. The POP asymmetric stretching band near 910920 cm1 shifts to higher

216

P.Y. Shih et al. / Journal of Non-Crystalline Solids 244 (1999) 211222

Fig. 5. FTIR spectra of the 70P2 O5 (30x)Na2 OxCuO


glasses: (a) x 0, (b) x 10, (c) x 20, (d) x 30 mol%.

Fig. 4. FTIR spectra of the 60P2 O5 (40x)Na2 OxCuO


glasses: (a) x 0, (b) x 10, (c) x 20, (d) x 30, (e)
x 40 mol%.

frequency as Na2 O is replaced by CuO. Compared


with Fig. 3, as the P2 O5 content increases, the P
OP asymmetric stretching band near 910 cm1
shifts to higher frequency and the bands around
12601360 cm1 are broadened especially for
glasses with CuO content > 30 mol%.
We could not determine the structural role of
CuO directly by the presence of cation-oxygen

bonds. However, FTIR spectra still oer some


indirect structural information. In silver metaphosphate glass, the PO stretching band near
1100 cm1 occurs at a wavenumber 30 cm1 less
than the frequency of the corresponding band in
sodium metaphosphate glass [29]. Such a shift was
attributed to the existence of covalent bonds between silver ions and the non-bridging oxygens
[29]. Similarly, we suggest that the shift of absorption bands for OPO (PO ) stretching to
lower frequencies with the substitution of CuO for
Na2 O in P2 O5 Na2 OCuO glasses, see Fig. 3, are
due to arise from an increase in the number of
cations in the network with increasing CuO content.
3.2. XPS analyses
The XPS O 1s spectra of the samples are shown
in Figs. 68. There are compositionally dependent

P.Y. Shih et al. / Journal of Non-Crystalline Solids 244 (1999) 211222

Fig. 6. XPS spectra of O 1s binding energy of the 40P2 O5


(60x)Na2 OCuO glasses.

Fig. 7. XPS spectra of O 1s binding energy of the 50P2 O5


(50x)Na2 OCuO glasses.

changes in the O 1s spectra of the P2 O5 Na2 O


CuO samples.
Sodium metaphosphate has the resonance chain
structure as suggested by Gresch et al. [40], who
proposed that the p bond of the P@O group is
distributed across the new non-bridging oxygen
with the addition of alkali ions to P2 O5 glass.
Thus, both the double bond, P@O, and nonbridging PO    Na , are not resolved by XPS
. Recent
[31,34] and were represented as
FTIR and Raman data revealed that P@O bonds

217

Fig. 8. XPS spectra of O 1s binding energy of the 60P2 O5


(40x)Na2 OCuO glasses.

are not present in metaphosphate glasses [31,39].


These observations are consistent with the resonance model.
Figs. 68 show the O 1s spectra of the samples
with 40, 50 and 60 mol% P2 O5 , respectively. These
gures indicate that the high binding energy feature broadens and the binding energy maximum
occurs at larger energy with increasing CuO content. Fig. 9 shows the decomposed O 1s binding
energy of the 50P2 O5 (50x)Na2 OCuO glasses.
Previous studies have established that the high
binding energy peak in the spectrum of the NaPO3
based sample is due to bridging oxygen (POP),
and the low binding energy peak is due to non) [15,40]. By deconvolution
bridging oxygen (
into their respective components, the O 1s spectrum from 50P2 O5 50Na2 O samples (Fig. 9(a))
yields an experimental bridging-to-non-bridging
oxygen ratio (BO/NBO) of 0.496, in excellent
agreement with the theoretical ratio of 0.500 for
NaPO3 . With increasing CuO content, the high
binding energy feature broadens and the binding
energy maximum occurs at larger energy. These
spectral changes are due to the substitution of
CuO for Na2 O and can be explained by the development of a third peak in the decomposed
spectra for oxygen bonded between P and Cu
(Fig. 9). The electronegativity of Cu is between
that of P and Na, resulting in smaller binding

218

P.Y. Shih et al. / Journal of Non-Crystalline Solids 244 (1999) 211222

Fig. 10. The eect of glass composition on the fractions of


bridging (POP), non-bridging and asymmetric (POCu)
oxygens measured from the integrated peak areas of decomposed O 1s spectra of 40P2 O5 (60x)Na2 OxCuO glasses. The
lines are drawn using the least square tting method.

Fig. 9. Decomposed O 1s binding energy of the 50P2 O5


(50x)Na2 OCuO glasses: (a) x 0, (b) x 20, (c) x 40, (d)
x 50 mol%.

energy of the O 1s in the POCu than that of PO


bonds.
P bonds but larger than that of
In metaphosphate sample, the largest binding
energy, 533.2 eV, is referenced to the bridging
oxygen (POP), and the smallest binding energy,
531.5 eV, is referenced to non-bridging oxygen
). These results are consistent with the work
(
of Onyirluka [42], who assigned the bridging oxygen and non-bridging oxygen of zinc metaphosphate glass to 533.1 and 531.6 eV, respectively.
Besides, the third peak at a binding energy of
532.2 eV is due to POCu bonds which has been

suggested in binary copper phosphate glasses by


Salim et al. [23], but without experimental evidence.
Figs. 1012 show the eect of glass composition
on the bridging (POP), non-bridging (P@O,
) and asymmetric bridging (POCu) oxygen contents of decomposed O 1s spectra of the
samples with 40, 50 and 60 mol% P2 O5 , respectively. The linear lines in these gures are plotted
using the least square tting method. In the samples with 40 mol% P2 O5 , the fraction of POCu
and POP bonds
bonds increases and both
decrease as Na2 O is replaced by CuO. We observe
that the rate of decrease of the fraction of POP
bonds increases when the CuO content exceeds 20
mol%. This observation agrees well with the FTIR
results. This structural change corresponds to the
phase change (high Tg phase disappears when CuO
content exceeds 20 mol%). In the samples with 50
and 60 mol% P2 O5 , the fraction of POCu bonds
decreases, whereas, the
increases and that of
fraction of POP is almost unchanged as Na2 O
is replaced by CuO. Hence, the aect of copper

P.Y. Shih et al. / Journal of Non-Crystalline Solids 244 (1999) 211222

219

Fig. 11. The eect of glass composition on the fraction of


bridging (POP), non-bridging and asymmetric (POCu)
oxygens measured from the integrated Peak areas of decomposed O 1s spectra of 50P2 O5 (50x)Na2 OxCuO glasses. The
lines are drawn using the least square tting method.

Fig. 12. The eect of glass composition on the fractions of


bridging (POP), non-bridging (P@O, ) and asymmetric (PO
Cu) oxygens measured from the integrated peak areas of decomposed O 1s spectra of 60P2 O5 (40x)Na2 OxCuO glasses.
The lines are drawn using the least square tting method.

cation in the sodium polyphosphate glasses diers


from that in sodium metaphosphate and sodium
ultraphosphate glasses.
Compare Figs. 1012, the relative amount of
bridging oxygen (POP) obviously decreases with
decreasing P2 O5 content. The decrease is indicative
of the smaller chain length with decreasing P2 O5
content owing to the depolymerization of the
phosphate network.

and is attributed to depolymerization of the


phosphate network [29].
The (PO2 )s and PO3 bands decrease in amplitude with increasing CuO content in the sample.
This is because PO2 is the stretching mode of two
non-bridging oxygen atoms bonded to phosphorus
atoms, and PO3 is the end group of the phosphate
chain. The fraction of non-bridging oxygen decreases and crosslinking increases due to the formation of POCu bonds in the glass network as
CuO replaces Na2 O.
When a modier cation was incorporated into
chain-like phosphate glasses, it will shorten the
chain by disrupting the POP bond and form
ionic cross-bonding between the broken chains.
The decreasing amplitude of the POP absorption band indicates that the copper ions may act as
a glass modier as do the alkali ions. However,
when Na2 O is replaced by CuO the number of
modier ions does not increase in this glass. This
suggestion is also contrary to the progressive decrease of amplitude of the PO3 end group and the

4. Discussion
4.1. Samples with less than 50 mol% P2 O5
A comparison of the spectra in Figs. 2 and 3,
the amplitude changes of the (PO2 )as stretching
mode is most remarkable. The amplitude of the
(PO2 )as stretching bands decreases as the P2 O5
content decreases from 50 to 40 mol%. This decrease may be duo to a decrease of the length of
phosphate chains with decreasing P2 O5 content,

220

P.Y. Shih et al. / Journal of Non-Crystalline Solids 244 (1999) 211222

properties of the glasses shown in Table 1.


Therefore, this proposition is not consistent with
the data. Thus, the role of copper ion is similar to
that of the aluminum cation in the sodium phosphate glasses [15]. We suggest that the formation
of POCu will replace both POP and PO
bonds and lead to a reduction in the amplitude of
PO2 , POP and PO3 bands in the FTIR spectra as
CuO replaces Na2 O in sodium polyphosphate
glasses.
The XPS results are in agreement with the
FTIR results. The magnitude of the high energy
component peak (associated with POP oxygen)
in the O 1s spectra follows a trend similar to the
900 cm1 absorption band (also associated with P
OP oxygen) in the FTIR spectra. From Fig. 10, It
is evident that the fraction of POCu bonds inand POP decrease
creases, whereas, both
as Na2 O is replaced by CuO. The replacement of
POP bonds with POCu bonds has an eect on
glass properties and is similar to the eects when
Al2 O3 or Fe2 O3 is added to the phosphate glass,
namely, increased glass transition temperature and
chemical durability [12,15].
Brow et al. [12] reported the eect of Fe valence
on the structure of iron phosphate glasses. They
suggested that the number of POP bonds can be
decreased by increasing the average Fe valence
because more [O]/[Fe] is incorporated into the
structure for equivalent concentrations of Fe3
than for Fe2 [12]. Since the formation of POCu
bonds will decrease the number of POP bonds in
our samples, we suggest that this decrease is due to
the increase of [Cu2 ]/[Cutotle ] ratio in the present
samples with 40 mol% P2 O5 . The crystal structures
of Cu(I)PO3 and Cu(II)2 P2 O7 have the same [Cu]/
[P] ratio, but with dierent bonding states. We can
illustrate an eect of Cu valence on the structure of
the polyphosphate glass that is simillar to that of
Fe ions. The metaphosphate crystal has large
amount of POP bonds, whereas the polyphosphate crystal has only one POP for every
six POCu bonds. Thus, an increase of [Cu2 ]
leads to a more depolymerized phosphate structure. Since the composition of polyphosphate
glasses favors the polyphosphate bonding structure, the result of XRD (Fig. 1) is consistent with
this suggestion. The metaphosphate and ultra-

phosphate glasses favor the chain-like structure.


The depolymerization of POP bonds is observed
in polyphosphate glasses and not in metaphosphate or ultraphosphate glasses.
4.2. Samples with 50 mol% P2 O5
When modier cations are added to phosphate
glasses, the P@O of phosphate groups is unaected
and depolymerization takes place through the
breaking of POP linkages only. The eect of a
network former is quite dierent from that of a
network modier [40,41]. It has been suggested
that the addition of PbO and Al2 O3 will form P
OPb and POAl linkages by opening P@O
bonds of PO4 tetrahedra [42,43]. Similarly, such a
view point was also used to interpret the formation
of covalent POCu linkages in binary copper
phosphate glasses [23].
FTIR spectra (Fig. 3) indicate the structural
changes of the ternary samples with 50 mol%
P2 O5 . Absorption bands of (PO2 ) asymmetric
stretching modes and PO groups shift to lower
frequencies as the CuO content increase. Since the
electronegativity of Cu is larger than that of Na,
we expect that the copperoxygen bond is more
covalent than the sodiumoxygen bond, and the
phosphorus-oxygen bonds linked to copper ions
[PO(Cu)] are more ionic than
bonds. Furthermore, the decrease of (PO2 )as inbonds will
tensity conrms that O
be replaced by the POCu bonds as Na2 O is replaced by CuO.
Bands of POP asymmetric and symmetric
stretching modes shift to higher frequencies as the
CuO content increases. The band shift can be explained by an increase in the covalent fraction of
POP bonds, indicating that the POP bonds
are strengthened as the Na2 O is replaced by CuO.
Thus, Tg and Td increase and the chemical durability improves as the CuO content increases (see
Table 1). The properties of the P2 O5 Na2 OCuO
samples depend on composition, and change with
the amount of any one of the component oxides.
The systematic changes in the O 1s spectra in
Fig. 9 indicate that the formation of POCu
upon the subbonds, which replace
stitution of CuO for Na2 O. We attribute improved

P.Y. Shih et al. / Journal of Non-Crystalline Solids 244 (1999) 211222

chemical durability and increased Tg , Td , and


hardness of the samples is a result of increasing the
crosslinks between the phosphate chains.
Fig. 11 indicates that the fraction of POCu
decreases
bonds increases, whereas, that of
as CuO replaces Na2 O. Besides, the fraction of
bridging oxygen (POP) is almost unaected by
glass composition. It is interesting to note that the
fraction of bridging oxygen is almost identical and
constant at 33% independent of CuO content.
The role of copper cation in sodium metaphosphate glasses diers from that of the aluminum
cation. In the latter case, it has been suggested that
the fraction of POP and PO bonds both decrease and the fraction of POAl bonds increases
when Al2 O3 is added [15].
4.3. Samples with more than 50 mol% P2 O5
The FTIR spectra show that the stretching
band of P@O double bonds is near 1320 cm1 in
the ultraphosphate (Figs. 4 and 5). As Cu cations
replace Na cations in the phosphate glass structure, the POP bond strength increases because
of the greater eld strength of the Cu cation
compared with that of Na cation. Thus, the POP
asymmetric stretching band near 910920 cm1
shifts to higher frequency as Na2 O is replaced by
CuO.
Compared with the FTIR spectra of metaphosphate samples (Fig. 3), the bands around
1310 cm1 are broadened in the spectra of ultraphosphate samples especially of the samples with
CuO contents > 30 mol%. Since the absorptions
for P@O double bonds coexist with (PO2 )as in the
ultraphosphate samples, they overlap in the region
around 12601360 cm1 causing the broadened
bands. The absorptions for (PO2 ) asymmetric
stretching modes shift to lower frequency as the
CuO content increases. Thus, the substitution of
CuO for Na2 O will increases the widths of bands.
Fig. 12 shows the eect of glass composition on
and
the bridging (POP), non-bridging (
P@O) and asymmetric bridging (POCu) oxygen
contents measured from the decomposed O 1s
spectra of samples containing 60 mol% P2 O5 . It is
clear that the fraction of POCu bonds increases,
whereas, that of non-bridging oxygen (P@O and

221

) decreases as Na2 O is replaced by CuO. The


fraction of bridging oxygen (POP) is nearly independent of glass compositions 43% independent of CuO contents. This invariance is similar to
that for metaphosphate glasses.
5. Conclusions
(including
The fraction of POP and
P@O) as well as POCu bonds are composition
dependent. The FTIR results are in agreement
with the XPS analyses. In the ultraphosphate and
metaphosphate glasses, POCu bonds are
bonds, and the
formed, which replace
fraction of POP bonds is constant as Na2 O is
replaced by CuO. In the polyphosphate glasses,
the formation of POCu bonds replaces both P
bonds. The formation of P
OP and
OCu bonds in these glasses increases the crosslink
density in the glass network, therefore, chemical
durability and Tg of the glasses are increased.
Acknowledgements
The authors are grateful to the National Science
Council of the Republic of China for nancial
support, under grant NSC 87-2216-E-239-004.
Gratitude is also extended to Mr Rea-Chin Lee of
National Cheng Kung University for his kind help
in recording the XPS spectra.
References
[1] P.Y. Shih, S.W. Yung, C.Y. Chen, H.S. Liu, T.S. Chin,
Mater. Chem. Phys. 50 (1997) 63.
[2] P.A. Tick, Phys. Chem. Glasses 25 (1984) 149.
[3] L.M. Sanford, P.A. Tick, US Pat. 4,314,031 (1982).
[4] Y. He, D.E. Day, Glass Technol. 33 (1992) 214.
[5] H.S. Liu, P.Y. Shih, T.S. Chin, Phys. Chem. Glasses 37
(1996) 227.
[6] C.M. Shaw, J.E. Shelby, Phys. Chem. Glasses 29 (1988) 49.
[7] H. Yung, P.Y. Shih, H.S. Liu, T.S. Chin, J. Am. Ceram.
Soc. 80 (1997) 2213.
[8] M.R. Reidmeyer, M. Rajaram, D.E. Day, J. Non-Cryst.
Solids 85 (1986) 186.
[9] C.M. Shaw, J.E. Shelby, Phys. Chem. Glasses 29 (1988) 87.
[10] I.W. Donald, J. Mater. Sci. 28 (1993) 2841.

222

P.Y. Shih et al. / Journal of Non-Crystalline Solids 244 (1999) 211222

[11] C.M. Shaw, J.E. Shelby, J. Am. Ceram. Soc. 71 (1988)


C252.
[12] R.K. Brow, C.M. Arens, X. Yu, D.E. Day, Phys. Chem.
Glasses 35 (1994) 132.
[13] R.K. Brow, J. Am. Ceram. Soc. 76 (1993) 913.
[14] R.K. Brow, R.J. Kirkpatrick, G.L. Turner, J. Am. Ceram.
Soc. 76 (1993) 919.
[15] R.K. Brow, R.J. Kirkpatrick, G.L. Turner, J. Am. Ceram.
Soc. 73 (1990) 2293.
[16] B.C. Bunker, G.W. Arnold, M. Rajaram, D.E. Day, J. Am.
Ceram. Soc. 70 (1987) 425.
[17] R.W. Larson, D.E. Day, J. Non-Cryst. Solids 88 (1986) 97.
[18] D.E. Day, J. Non-Cryst. Solids 112 (1989) 7.
[19] M.A. Tindyala, W.R. Ott, Am. Ceram. Soc. Bull. 57 (1978)
432.
[20] A.S. Shawoosh, A.A. Kutub, J. Mater. Sci. 28 (1993) 5060.
[21] B.-S. Bad, M.C. Weinberg, Glass Technol. 35 (1994) 83.
[22] B.-S. Bae, M.C. Weinberg, J. Am. Ceram. Soc. 74 (1991)
3039.
[23] M.A. Salim, G.D. Khattak, M. Sakhawat Hussain, J. NonCryst. Solids 185 (1995) 101.
[24] E.E. Khawaja, M.N. Khan, A.A. Kutub, C.A. Hogarth,
Int. J. Electronics 58 (1985) 471.
[25] P.Y. Shih, S.W. Yung, T.S. Chin, J. Non-Cryst. Solids 224
(1998) 143.
[26] H. Zheng, M.W. Colby, J.D. Mackenzie, J. Non-Cryst.
Solids 127 (1991) 143.

[27] A. Mekki, D. Holland, K. Zig, C.F. McConville, Phys.


Chem. Glasses 39 (1998) 183.
[28] J. Koo, B-S. Bae, H-K. Na, J. Non-cryst. Solids 212 (1997)
173.
[29] R.F. Bartholomew, J. Non-Cryst. Solids 7 (1972) 221.
[30] E.E. Khawaja, S.M.A. Durrani, F.F. Al-Adel, M.A. Salim,
M.S. Hussain, J. Mater. Sci. 30 (1995) 225.
[31] R.K. Brow, D.R. Tallant, S.T. Myers, C.C. Phifer, J. NonCryst. Solids 191 (1995) 45.
[32] J.J. Hndgens, S.W. Martin, J. Am. Ceram. Soc. 76 (1993)
1691.
[33] K. Meyer, J. Non-Cryst. Solids 209 (1997) 227.
[34] L. Montagne, G. Palavit, G. Mairesse, Phys. Chem.
Glasses 37 (1996) 206.
[35] D.E. Corbridge, J. Appl. Chem. 6 (1956) 456.
[36] X.J. Hu, D.E. Day, Phys. Chem. Glasses 31 (1990) 183.
[37] H.S. Liu, T.S. Chen, S.W. Yung, Mater. Chem. Phys. 50
(1997) 1.
[38] J.O. Byun, B.H. Kim, S.K. Hong, H.J. Jung, S.W. Lee,
A.A. Izyneev, J. Non-Cryst. Solids 190 (1995) 88.
[39] R.K. Brow, D.R. Tallant, J.J. Hudgens, S.W. Martin,
A.D. Irwin, J. Non-Cryst. Solids 177 (1994) 221.
[40] R. Gresch, W. Muller-Warmuth, H. Dutz, J. Non-Cryst.
Solids 34 (1979) 127.
[41] E.C. Onyiriuka, J. Non-Cryst. Solids 163 (1993) 268.
[42] U. Selvaraj, K.J. Rao, J. Non-Cryst. Solids 104 (1988) 300.
[43] S. Chakraborty, A. Paul, J. Mater. Sci. Lett. 8 (1989) 1358.

Vous aimerez peut-être aussi