Sie sind auf Seite 1von 11

NanaStructured Materials. Vol. 8. No. 4. pp. 43545.

1997
Ekevia Science Ltd
0 1997 Acta Metallurgica Inc.
Printed in the USA. All rights rcswed
0965-9773/97 $17.00 + .OO

Pergannon

PI1 SO9659773(97)00189-X

FTIR STUDY OF A NANOSTRUCTURED ALUMINUM NITRIDE


POWDER SURFACE: DETERMINATION OF THE ACIDIC/BASIC
SITES BY CO, CO2 AND ACETIC ACID ADSORPTIONS
M.-I. Baraton, X. Chen**, K.E. Consalves*
LMCTS, ESA 6015 CNRS, 123 Av. Albert Thomas, F-87060 Limoges, France
*Polymer Program at the Institute of Materials Science and Department of Chemistry,
University of Connecticut, Storm, CT 06269, USA
*Current Address: Office Imaging Research and Technology, Eastman Kodak Company,
Rochester, NY 14650-2129, USA
(Accepted April 20,1997)
Abstract--The surface contamination of nanosizedpowders is a criticalfactor influencing
the overall material properties. For non-oxide ceramics, hydrolysis leads to a surface reactivity
close to that of the corresponding oxide. However, the surface behavior of these non-oxide
nanostructured ceramics often results in chemical spectficities. A thorough investigation of the
surface composition and reactivity of a nanostructured aluminum nitride powder, already started
inprevious papers, is pursued here. The Fourier transform infrared spectrometric study of carbon
monoxide, carbon dioxide and acetic acid adsorptions on the activated aluminum nitride surface
revealed the presence of Brensted and Lewis basic sites. Furthermore, the aluminum nitride
sulfate chemical spectficity with respect to the ~alumina surface was demonstrated. 0 1997Acta
Metallurgica Inc.

INTRODUCTION
The surface of materials is the key point in many technological processes and, therefore, the
knowledge of the chemical composition of the first atomic layer becomes critical for the
understanding of surface and interface phenomena. For example, the dispersion of powders in a
given solvent requires the knowledge of powder surface charge properties which control the
adsorption of additives and dispersion forces. The charge behavior is thought to be determined by
the ratio between acidic and basic groups (1). For nanosized powders, all these dispersion
phenomena become even more crucial and strongly depend on their synthesis process. Furthermore, the fundamental question on how the surface of a non-oxide material is contaminated by
atmospheric water, oxygen or carbon dioxide becomes of tremendous interest for nanosized
powders. Indeed, the large surface to bulk ratio makes any contamination and oxidation of the
surface critical factors influencing the nanosized powder properties.
In order to pursue the investigation of the chemical composition of and the charge
distribution on the nanostructured AlN powder surface (2,3), the chemical interactions with CO,
435

436

M-l BARATON,X CHENAND KE GONSALVES

CO2 and CH3COOH molecules were analyzed in situ by FTlR spectrometry. The use of these
molecules which are acidic and basic probes of different strengths, along with the comparison with
the well-known y-alumina surface, will bring a more complete picture of this aluminum nitride
surface to light.
EXPERIMENTAL
The chemical synthesis as well as the bulk characterization of the nanostructured aluminum
nitridepowderhas been previously reported (2,3). Carbon monoxideandCOz gases were supplied
by Alphagaz and acetic acid by Prolabo (BP. Normapur) with no further purification. Our
experimental setup for surface characterization has been described elsewhere (4,5). As a reminder,
the Fourier transform infrared spectrometer, along with the specific attachment, allows one to run
in siru experiments. The powder, slightly pressed into thin grid-supportedpellets,can be thermally
treated from room temperature up to 873 K under atmosphere, vacuum or controlled pressures of
gases or liquid vapors. As previously explained (3), the sample is thermally activated to eliminate
the physisorbed and the weakly chemisorbed species on the surface before the addition of the
probe-molecules. In order to avoid any temperature effect, all the spectra were recorded at room
temperature. Therefore, for brevity purposes, the following sentence ...after evacuation at 473
K... will stand for ...after evacuation during heating at 473 K and subsequent cooling at room
temperature.. throughout the text. The spectra presented in the figures are difference spectra.
Consequently, all the bands correspond to modifications of the AlN surface as follows: the positive
bands reveal appearing or increasing species, whereas the negative bands reveal disappearing or
decreasing species.
RESULTS AND DISCUSSION
Carbon Monoxide Adsorption

Carbon monoxide is a soft Lewis base that can reversibly adsorb on metal oxide surfaces
through coordinatively unsaturated (cus) cations of high acidic strength (6,7). The weak
interaction results in an upward shift of the v(C0) frequency with respect to the gas phase,
depending on the nature of the surface cation andon its chemical environment. Although different
pressures of CO varying from 3 to 30 mbar were added to the activated AlN sample, no adsorption
was found. However, our previous study on pyridine adsorption showed that Lewis acidic sites
exist on the AlN surface (3). Unlike y-alumina (8-lo), only one type of acidic sites was observed
on the AlN surface corresponding to coordinative vacancies shared by Alrv and Alv~ cation pairs
(3). At room temperature, no strong Alrv3+ Lewis acidic sites was detected (3). Moreover, for
Al2O3, it has been proven that most of the Lewis acidic sites responsible for CO adsorption are
adjacent to strong basic sites (67). Because of the presence of nitrogen atoms on the surface, these
acidic-basic pairs might be less reactive on the AlN surface than on that of y-alumina and, as a
consequence, these sites may not interact with the CO soft base.
Carbon Dioxide Adsorption

Carbon dioxide is an acidic probe. But, in fact, C@ molecules can adsorb on positive and/
or negative surface sites in such different ways that many different adsorbed species can

437

FTIR STUDYOFA NANOSTRLCTURED


ALUMINUM
NITRICE
POWDER
SURFACE

4000

3000

2ooo

Wavenumber
(cm-1)
Figure 1. C@ adsorption (14 mbar, room
temperature (rt)). Difference spectra show
-ing the evolutions of the adsorbed species:
(a) after 5 minutes contact, (b) between 5
and 30 minutes contact, and (c) between 30
and 60 minutes contact.

3ooo

2ooo

Wavenumber
(cm-1)
Figure 2. CR adsorption (14 mbar, rt).
Difference spectral (a) after 60 minutes
contact and (b) after evacuation at rt.

simultaneously appear. These latter, which may result from chemical reactions with the surface,
strongly depend on the synthesisconditions and/or thermalpretreatments. Extensive studies of the
CQ adsorption types for metal oxides and particularlyfor aluminas can be found in the literature
(see for example Ref. 11 to 14).
Fourteen mbar of CQ2 were added to the activated AIN pellet (Figure la-c) and remained
in contact for 60 minutes. During the first 5 minutes (Figure la)?severalbands appear below 1800
cm corresponding to the immediate formation of new species. A strong absorption band at 2349
cm-l was observed. Between 5 and 30 minutes contact (Figure lb), the 2349 cm- band still
increases as well as other bands below 1700cm-*. During the last step (Figure Ic), the changes
are small as only the increase of three bands around 1653,1456 and 1300cm*was noted. After
60 minutes contact (Figure 2a), the changes with respect to the activated bare surface are strong
in the 1800-1000cm-l region and a distinct band is noted at 3611 cm-l.

438

M-l

BARATON,

X CHEN AND KE GONSALVES

An evacuation at room temperature removes most of the species and the surface completely
recovers after desorption at 473 K. We note that the NH2 surface groups in the 3200 cm- region
are not perturbed (Figures 2b. 3a,b).
The band at 2349 cm-l (Figure la), close to the frequency of the gaseous phase absorption
frequency, is related to end-on CO2 complexes on uncoordinated cations (Scheme 1.1). These
linearly adsorbed species are usually completely eliminated by evacuation at room temperature.
According to Morterra (13), this end-on Co;? adsorption, which should bring similar information
as the CO adsorption, actually involves more surface cationic sites. Indeed, COZ adsorption can
occur on partly activated alumina and also on Alv~ coordinatively unsaturated sites present on the
corundum surface.

Scheme 1 Proposed mechanisms for CO2 adsorption

(I) ends-on complex

fl
8
g
11111111111111111111llllll
)y

3+ l1ll1lllllllllllllllllll

AIN Surface

(II) hydrogenocarbonyl ion

ON

-0

C-

\\

.ow

ONH
\ :llllllllllllllllllllllllll
lllllllllllllllll1lllllllll~

"\
1111111,111111111111111111(,
:llllllllllllllllllllllllll

AlN Surface

AlN St&ace

(III) bridged carbonate

j3+

11111111111111111

llllflllll:3+

11111111111111111111llll

AlN Surface

FTIR STUDYOFA NANOSTFIUCT~RED


ALUMINUM
NITRIDE
POWDER
SURFACE

439

The appearing band at 3611 cm-l along with the negative band around 3780 cm-l
corresponding to a decrease of the <<free>> hydroxyl groups (Figure 2a), bring the evidence of
the adsorption of Co2 on Al-OH surface groups, yielding hydrogenocarbonyl ion (bicarbonate)
(Scheme 1.II). This 3611 cm-l band, assigned to thev(OH) stretching vibration, is associated with
the bands around 1670-1640 cm-, that is vas(C=O), 1490-1440 cm-l, that is v&=0) and 13001200 cm-, that is &OH) (llJ5).
In Figure la, it can be noted that the band corresponding to
v&=0)
around 1650 cm- is double and that the 1480 cm- v,(C=O) band has a shoulder on the
low wave number side. The complexity of these bands also appears in Figure 3a, showing the
species disappearing under evacuation at room temperature. Therefore, similarly to y-alumina,
several types of bicarbonate species should be considered (13).
The formation of these bicarbonate groups can be confirmed by C@ adsorption on the
deuterated AlN surface. Indeed, the v(0I-I) stretching band around 3611 cm- should shift toward
lower wave number by a factor of 1.356, as should do the S(0I-I) band. Indeed, the AlN activated
surface was first deuterated by DZaddition at 873 K and then, 14 mbar of CO2 were added at room
temperature (Figure 4a,b). The decrease of the band at 2790 cm- corresponding to the v(OD)
mode of surface Al-OD groups (3784 cm-l in Al-OH) and the related increase of the 2670 cm-

4000

3000
2000
Wavenumber (cm-l)

Figure 3. CO2 desorption. Difference spectra: (a) evacuation at


rt and (b) evacuation from rt to 473 K.

M-tBAPATON,X CHEN AND KE

440

GONSALVES

TABLE 1
Main Species Formed on the AlN Surface by CO;! Adsorption
Frequency (cm- 1)

Species

Thermal stability

2349

end-on adsorbed CQ

- immediately formed
- eliminated by evacuation at rt

3611
1653
1480
1235

hydrogenocarbonyl
species 1

- immediately formed
- increase with time
- eliminated by evacuation at 473 K

3611
1695
1487
1270

hydrogenocarbonyl
species 2

- immediately formed
- increase slightly with time
- eliminated by evacuation at rt

1772
1200

carboxylate
species 3

- immediately formed
- eliminated by evacuation at rt

1767
1198

bridged carbonate
species 4

- immediately formed
- increase slightly with time
- eliminated by evacuation at 473 K

corresponding to the v(OD) mode of bicarbonate species are clearly visible. Moreover, by
comparing Figure 2a and 4a, we also note the disappearing of the bands at 1235 and 1297 cm-l
assigned to the 6(OH) bending mode in bicarbonate groups, shifted toward the obscure region of
the spectrum (probably around 980 cm-).
Therefore, we can conclude that two different hydrogenocarbonyl groups are formed whose
vibrational frequencies are given in Table 1.
Their thermal stability is slightly different: the species 1 seems to be more stable since it is
only eliminated at 473 K (Figure 3b) whereas species 2 disappears by pumping at room
temperature (Figure 3a). They may correspond to the interaction with different AI-OH surface
hydroxyl groups, but the weakness of the v(OH) bands makes their precise assignment difficult.
However, it is well known that on alumina, Co;? reacts with the most basic OH surface groups; that
is, the groups corresponding to Al~v-OH groups with the highest v(OH) frequencies (11). This is
indeed proven by the reaction of C& with the de&rated AlN surface, but it is also possible that
CO2 partly reacts with the Al-OH groups on the AIN surface whose v(OH) vibration absorbs at
3740 cm-.
The hydrogenocarbonyl groups are not the only species resulting from the reaction of CO2
with the AlN surface. A strong band at 1772 cm-l appears during the first 5 minutes (Figure la)
concomitantly with a shoulder at 1200 cm- l. Both bands decrease by evacuation at room
temperature (Figures 2b, 3a). According to Ref. 11 and 16, both frequencies could be assigned
either to bridged carbonate (or organic-like) or to carboxylate species. The latter species,

FTIR STUDYOF A NANOSTRUCTURED


ALUMINUM
NITRIDEPOWMR SURFACE

4000

2000
3000
Wavenumber (cm- I)

l( jot 1

Figure 4. CO:! adsorption (14 mbar, rt) on


deuterated AN Difference spectra: (a) after
60 minutes contact and (b) after evacuation
at rt.

441

3000

2000
Wavenumber (cm- 1)

Figure 5. CHsCOOH adsorption (4 mbar, rt).


Difference spectra. (a) after 30 minutes contact
followed by an evacuation at rt, (b) after
evacuation at 423 K, (c)after evacuation at
573 K, and (d) after evacuation at 873 K.

expected to have a low thermal stability, should form by interaction of C@ with strong Lewis
acidic sites which, in the present case, have not been evidenced by CO adsorption. On the other
hand, bridged carbonates are supposedly thermally stable. Since the pyridine adsorption showed
the presence of A13+acidic sites (3), we think that carboxylate groups (species 3) are likely weakly
bound to the surface and quickly eliminated. However, the formation of bridged carbonates
(species 4) (Scheme l.III), involving two adjacent Al~v cations, must also be considered (13).
Indeed, these above-mentioned bands decrease under evacuation at room temperature, but a pair
of bands at close frequencies (1767 and 1198 cm-l) is still present and only disappears at 473 K
(Figure2b,3b). Asalastremark,thebandat
1297cm-* (Figure2a)isbroadandonlypartofit(1270
cm-) can be definitely assigned to the &OH) vibration of species 2. The spectrum in Figure 4b
clearly shows that after CO2 evacuation, a band at 1354 cm-l exists on the deuterated AlN surface
and could be a component of this broad 1297 cm-l band. It is hard to precisely assign this band
by itself, but we cannot discard the presence of monodentate carbonates whose v,,(C=O)

M-l BAFIATON,
X CHENAND KE GONSALVES

442

-a

4000

2000
Wavemmber (cm-l)

3000

r
:::
f
p
IO

Figure 6. CH3COOH desorption. Difference spectra. (a) evacuation from rt to 423 K, (b)
evacuation from 423 K to 573 K, and (c) evacuation from 573 K to 873 K.

frequency should absorb around 1550 cm-l, overlapped by the strong absorption band of
bicarbonate groups. Table 1 summarizes the species formed by reaction of C@ with the activated
AlN surface.
Acetic Acid Adsorption

AceticacidlikeformicacidisaB~nstedacidicmQleculebutcanalsoprobetheLewisacidic
and basic sites. The strength of the interaction may irreversibly contaminate the surface.
The AlN activated surface was subjected to 4 mbar of acetic acid at room temperature for
45 minutes and the gaseousphase in excess was eliminated by subsequent evacuation. Ike groups
of very strong bands appear around 1600 and 1450 cm-l and are assigned to carbonyl stretching
vibrations (Figure 5a). Moreover, the formation of hydrogen bonds is indicated by the increasing
broad band centered at 3230 cm-l and by the decreasing band centered at 3740 cm-l assigned
respectively to the v(OH) vibration of hydrogen bonded and free OH surface groups. Independently from the very intense carbonyl bands, a weaker band, overlapping the broad3230 cm- band,
has its absorption maximum at 3233 cm-r (Figure 5a). The bands in the 3050-2950 cm-* region
are assigned to the v(CH) vibrations of the CH3 groups in acetic acid.

FTIR STUDYOFA NANOSTRUCTURED


ALUMINUM
NITRIDEPOWDERSURFACE

443

An evacuation at 423 K (Figure 5b) removes most of the hydrogen-bonded species.


Continuous evacuation at increasing temperatures causes slight frequency shifts (Figure 5c,d). At
the end of the deisorption process (Figure 5d), the temperature reaches 873 K but the surface does
not recover. Twlo bands are still evident at 1575 and 1480 cm-l.
The obvious conclusion is that acetic acid reacts with the AlN surface, yielding carbonyl
containing species irreversibly linked to the surface. Even when the AlN pellet is brought back to
the ambient atmosphere anddesorbed again, these carbonyl species are still grafted on the surface.
The adsorption of formic acid on alumina revealed the irreversible presence of formate ions
characterized by two absorption bands at 1620 and 1344 cm- (17). On metals, the decomposition
of formic acid crm proceed, leading to hydrogen and carbon dioxide. It has also been proven that,
in addition to this chemisorption on metal, the silica support can be involved dissociating formic
acid as water and carbon monoxide (17). But, in the present case, no evidence of gaseous or
adsorbed CO2 and/or CO has been found at room temperature. Then, the subsequent desorption
process we carried out under dynamic vacuum eliminated the potential gaseous phases, anyway.
However, there is no doubt about the dissociative character of the adsorption. Indeed, the
v(C=O) band in free acetic acid is at 1728 cm. As soon as the acid in excess is eliminated (Figure
5a), this band nlo longer exists. On the contrary, the characteristic bands of the carbonyl group
appear. According to Ref. 18, the A splitting frequency between v&=0)
and v,(C=O) gives a
good indication on the bonding geometry. In the present case, this splitting is around 100 cm-
depending on the desorption temperature and can correspond to either bridged acetate (A=200 cm> or bidentate acetate (A c 80 cm-) (Scheme 2).

Scheme

2 Proposed mechanisms

for acetic acid adsorption

(V) bidentate acetate

A3+

lllilllllll1llllllllllllll

llllllllll1lllllllllllll

AIN Surface

(VI) bridged acetate

YH3
0

;3+

lllllllllllllllll

&-+so
llllllllll:3+

111,,1111111111111~1)111

AlN Surface

444

M-l BAPATON.
X CHENANDKE GONSALVES

Between 423 and 573 K, the desorbed species are characterized by the major bands at 1630
cm- (shoulder), 1600,1479 and 1422 cm (Figure 6b) assigned to differently coordinated acetate
ions (18-20). Part of the Al-OH surface groups (3740 cm-l negative band) participates in the
reaction forming these acetate ions. On the other hand, the quite stable band appearing at 3233 cm1 is not related to hydrogen bond since the latter interaction does not persist above 423 K (Figures
5b, 6a,b). This band couldratherbe assigned to av(NH) stretching. Indeed, by analogy with silicon
nitride, whose surface has been proven to present SisN sites (21), Al3N basic sites may exist on
the AlN surface. Moreover, the mechanism of methanol adsorption on AlN has also given
presumption of such sites, by the appearing of a similar band around 3220 cm- (3). The acidic
attack on the AlN surface is very strong, thus an opening of these Al3N groups could happen
(Scheme 3) leading to new >NI-Igroups and new Al acidic sites allowing the bonding of the acetate
groups in the bidentate geometry. These newly formed species are stable up to 573 K (Figures 5c,d,
6b,c).
At 873 K (Figure 5d), bidentate or bridged acetate groups are still bound onto the surface.
The v(CH) stretching bands in the 3000-2900 cm-l region, always very weak with respect to the
carbonyl bands, can no longer be discriminated from the noise.
CONCLUSION
The AlN surface reactivity toward CO2 and CHsCOOH probe-molecules was observed to
be strong and mostly basic. The Bronsted basicity of the surface generates hydrogenocarbonyls,
whereas the Lewis basicity is responsible for the formation of both carbonate and acetate groups.
Nevertheless, the presence of Al 3+ Lewis acidic sites is also proven by the formation of
carboxylate. Moreover, coordinative vacancies on Altv-Alw cationic pairs have been proven to
be strong acceptor sites towardpyridine molecules. But, because none of these acidic site adsorbs
CO, it may be concluded that, compared to the y-alumina surface, these sites are perturbed by the
presence of nitrogen atoms on the surface. Even though the NH2 surface groups are not involved
in interactions with any probe-molecule, the nitrogen atoms play an indirect role in the AlN surface

Scheme 3 Proposed mechanism

~N-AI.N,A,~-AI.N,A~r_~~.N,

for the formation of =NH group

~N.A~.N.*I_N.Al.N/~~_~~.~,

,Ll ~N.~,,N,A~N./;~,N,Al.N./i,~-~Al~N.i~,N,Al.N.il,Y,~~.N~,,
1111111111111111111
I llllllllllllllllllll
I 11111111111111111111

AlN surface

~N_*l.N,Al~~-Al.N,~~~.~,,~,

1111111111111111111
I llllllllllllllllllllll
11111111111111111111

AlN surface

,Al .,.~l.~,Al.N.~l,y,A~.~_~,,
1111111111111111111
I llllllllll1lllllllllll
11111111111111111111

AlN swface

FTIR STUDYOFA NANOSTRUCTURED


ALUMINUM
NITRIDEPOWDERSURFACE

445

chemistry. The weaker electronegativity of the nitrogen atom compared to oxygen implies a
different charge distribution and, thus, modifies the acid-base pair strength. Moreover, in the
presence of a strong Bronsted acid, nitrogen atoms from Al3N surface groups do participate to the
coordination as electron donor sites, thus strikingly demonstrating the specificity of the AlN
surface.
REFERENCES
1.
2.

3.

4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.

Liden, E., Journal of the European Ceramic Society, 1991,7,361.


Xiao, T.D.,, Gonsalves, K.E. and Strutt, P.R., Journal of the American Ceramic SocieQ, 1993,76,
987; Xiao, T.D., Gonsalves, K.E., Stnttt, PR, Chow, G.M. and Chen, X., Ceramic Engineering
Science Proceedings, 1993,14,1107.
Baraton, M.-I., Chen, X. and Gonsalves, K.E., Nanotechnology, Molecularly Designed Materials,
ACS Symposium Series 622, chapter 22, p. 312, eds. G.M. Chow and K.E. Gonsalves, ACS,
Washington, D.C., 1996.
Baraton, M.-I., Journal of High Temperature Chemical Processes, 1994,3,545.
Baraton, M.-I., Chen, X. and Gonsalves, K.E., Journal of Materials Chemistry, 1996,6, 1407.
Morterra, C., Bolis, V., Magnacca, G. and Cerrato, G., Journal of Electrically Related Phenomenon, 1993,64-65,235.
Ishida, S., Ixnamura, S. and Fujimura, Y., Reaction Kinetics Catafysis Letters, 1991,43,447.
Abbattista, F., Dehnastro, S., Gozzelino, G., Mazza, D., Vallino, M., Busca, G., Lorenzelli, V. and
Rarnis, G., Journal of Catalysis, 1989,117,42.
Ballinger, T.H. and Yates, J.T., Langmuir, 1991,7, 3041.
Zaki, M. and Knozinger, H., Spectrochim. Acta, 1987,43A, 1455.
Busca, G. and Lorenzelli, V., Materials Chemistry, 1982,7, 89.
Auroux, A. and Gervasini, A., Journal of Physical Chemistry, 1990,94,6371.
Morterra, C. and Magnacca, G., Catalysis Today, 1996,27,497.
Lavalley, J.C., Catalysis Today, 1996,27,377.
Baumgarten, E. and Zachos, A., Spectrochim. Acta, 1981,37A, 93.
Ramis, G., Busca, G. and Lorenzelli, V., Materials Chemistry, 1991,29,425 .
Little, L.H., Infrared Spectra of Adsorbed Species, Academic Press, London, 1966.
Nakamoto, K., Infrared and Raman Spectra of Inorganic and Coordination Compounds, Wiley,
New York, NY, 1986.
Tackett, J.E., Applied Spectroscopy, 1989,43,483.
Greenler, R.G., Journal of Chemical Physics, 1962,37,2094.
Busca, G.., Lorenzelli, V., Porcile, G., Baraton, M.-I., Quintard, I? and Marchand, R., Materials
Chemistry and Physics, 1986, 14, 123.

Das könnte Ihnen auch gefallen