Vous êtes sur la page 1sur 38

This article was downloaded by: [North Carolina State University]

On: 16 April 2013, At: 01:24


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number:
1072954 Registered office: Mortimer House, 37-41 Mortimer Street,
London W1T 3JH, UK

Catalysis Reviews: Science


and Engineering
Publication details, including instructions for
authors and subscription information:
http://www.tandfonline.com/loi/lctr20

IR Studies of NH3, Pyridine,


CO, and NO Adsorbed on
Transition Metal Oxides
a a
Mayfair C. Kung & Harold H. Kung
a
Chemical Engineering Department and the
Ipatieff Laboratory, Northwestern University,
Evanston, Illinois, 60201
Version of record first published: 13 Dec 2006.

To cite this article: Mayfair C. Kung & Harold H. Kung (1985): IR Studies of NH3,
Pyridine, CO, and NO Adsorbed on Transition Metal Oxides, Catalysis Reviews:
Science and Engineering, 27:3, 425-460

To link to this article: http://dx.doi.org/10.1080/01614948508064741

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/


terms-and-conditions

This article may be used for research, teaching, and private study
purposes. Any substantial or systematic reproduction, redistribution,
reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make
any representation that the contents will be complete or accurate or
up to date. The accuracy of any instructions, formulae, and drug doses
should be independently verified with primary sources. The publisher
shall not be liable for any loss, actions, claims, proceedings, demand, or
costs or damages whatsoever or howsoever caused arising directly or
indirectly in connection with or arising out of the use of this material.
Downloaded by [North Carolina State University] at 01:24 16 April 2013
.
CATAL REV.-SCI. E N G . , 27(3). 425-460 (1985)

IR Studies of NH3. Pyridine.


CO. and NO Adsorbed on
Transition Metal Oxides
Downloaded by [North Carolina State University] at 01:24 16 April 2013

.
MAYFAIR C K U N G AND HAROLD H K U N G .
Chemical Engineering Department and the Ipatieff Laboratory
Northwestern University
Evanston. Illinois 60201

I . INTRODUCTION ..................................... 426


I1 . AMMONIA A N D PYRIDINE ............................ 426
A . Modes of Adsorption ............................. 427
B . Differences between N H , and Pyridine ............ 428
C . Effect of Sample Pretreatment .................... 429
I11 . co A N D NO ......................................... 436
A . Bonding of CO to Surfaces ....................... 437
B . Bonding of NO to Surfaces ....................... 449
C . Surface Properties Deduced from
Adsorption of CO and NO ....................... 450
IV . COADSORPTION OF SMALL MOLECULES .............. 454
V . CONCLUSION ........................................ 455
REFERENCES ........................................ 456

425
Copyright 0 1985 by Marcel Defier. Inc . 016 1-4940/85/2103~425S3.50/0
426 KUNGANDKUNG

I. INTRODUCTION

Chemisorption of small molecules is often used as a probe for


the surface properties of transition metal oxides. By probing
the interaction of molecules with the surface, information is often
obtained on the oxidation state, the coordination symmetry, the
degree of coordination unsaturation of the surface cations, the
acid-base properties of the surface hydroxyl groups, and the
presence and the nature of surface Lewis acid and Bmnsted acid
Downloaded by [North Carolina State University] at 01:24 16 April 2013

sites. This information is deduced from experimental measure-


ments of the adsorption isotherms, the heats of adsorption, the
thermal desorption spectra, and the vibrational spectra of the
adsorbate. Until recently, when high resolution electron energy
loss spectroscopy became available, vibrational spectra were ob-
tained with infrared spectroscopy. Laser Raman spectroscopy
has seldom been used because of the low Raman scattering cross
section of most molecules.
Infrared spectroscopy has been used to study practically all
kinds of adsorbates, including basic molecules of N H , and pyri-
dine, carbon oxides (CO and CO,) , nitrogen oxides ( N O , N,O,
NO,), alkenes, alcohols, and acids. In this review, emphasis
is placed on the molecules that yield information on the acid-
base and on the coordination properties of the surface sites on
transition metal oxides. Thus the discussions are confined pri-
marily to studies involving NH,, pyridine, C O , and NO. CO,
is not included because a recent, extensive review has been
published [ 11. Emphasis is also placed on work published after
1970 because an excellent account of the literature before then
has already appeared in the book by Little [ 21.

11. AMMONIA AND PYRIDINE

These two molecules are regarded as basic molecules because


they possess a lone pair of electrons at the nitrogen available for
donation to other species during bonding, and because they can
accept a proton from other species. Thus, they have been used
to probe for the presence and the nature of surface acid sites.
In this section, information obtainable from the study of the in-
teraction of these molecules with transition metal oxide surfaces
is reviewed and discussed. The mode of adsorption of these
molecules, their differences, and the importance of surface pre-
treatment in affecting the mode of bonding are discussed.
IR STUDIES OF N H , , PYRIDINE, CO, AND NO 427

A. Modes of Adsorption

N H , and pyridine bond to the surface in three different modes.


The first mode involves transfer of proton from surface hydroxyl
to the adsorbate and occurs with the surface acting as a B r ~ n s t e d
acid. The second mode involves coordination of the adsorbate
through the electron lone pair to the metal of the oxide and takes
place with the solid acting as a Lewis acid. Such bonding is
known as o-donation bonding because the pair of electrons in
the bond is donated entirely by the adsorbate. Hydrogen bond-
Downloaded by [North Carolina State University] at 01:24 16 April 2013

ing is the third mode. It can occur through the interaction of


the nitrogen with the hydroxyl group on the surface o r , for N H , ,
through interaction of the anions of the oxide with the hydrogen
of the adsorbate. Hydrogen bonding is the weakest mode of in-
teraction. Hydrogen-bonded molecules can be pumped off readily.
The presence of Bransted acidity is revealed by the character-
istic band for the NH,+ ion located between 1400-1480 cm- [ 21.
Generally the identity of the infrared bands of adsorbed N H , are
established through comparison with the infrared bands of N H ,
in solution. For this purpose, a detailed table prepared by
Tsyganeko [ 31 of N H , infrared absorption bands in solutions of
various proton-accepting and proton-donating solvents can be
used. For Lewis acidity, Morimoto et al. [ 41, after reviewing
various studies of N H , adsorbed on metal oxides, have concluded
that for N H coordinated to surface metal ions, vas = 3330-3380
cm- (gas phase 3444 cm-1, v s = 3260-3280 cm- (gas phase 3336
.
cm-) Additional hydrogen-bonding interaction with the neigh-
boring anions would lower these stretching frequencies [ 51. This
follows from the fact that the stretching frequencies in amine com-
plexes decrease as the metal-nitrogen bond order increases [ 61.
When hydrogen bonding to the surface hydroxyl groups is the
only interaction with the surface, the NH, infrared bands appear
near 3400 and 3200 cm- [ 41.
For pyridine, according to Knozinger [71, the ring vibration
mode of 19b and 8a are most affected by the nature of intermo-
lecular interactions via the nitrogen lone pair electrons. These
two modes are observed at 1440-1447 and 1580-1600 cm-, respec-
tively, for hydrogen-bonded pyridine, at 1535-1550 and about
1640 cm- for the pyridinum ion, and at 1447-1464 and 1600 to
1634 cm- for pyridine coordinatively bonded to Lewis acid sites.
While the general classification of the bonding mode is possible
from the band frequencies, it does not seem possible to correlate
the exact band position with the nature of the metal ion and its
428 KUNGANDKUNG

coordination. This is because Gill et al. [ E l found, in their


studies of pyridine complexes with metals of coordination num-
bers four and six, that there is no systematic change in the
band position with respect to the nature of the central metal
atom or the types of ligands bonded to the metal atom. This con-
clusion has been substantiated by Zecchina et al. [91 for chemi-
sorption.

B . Differences between N H and Pyridine


Downloaded by [North Carolina State University] at 01:24 16 April 2013

Although the two molecules have been used to probe similar


surface properties, their properties are not identical. One dif-
ference between N H , and pyridine is their relative sizes. On
Cr,O, [5] the amount of irreversibly adsorbed N H , is 5.9 mole-
cules/nm2 and that of pyridine is 2.4 molecules/nm2. This cor-
relates well with their molecular cross-sectional area of 0.127
nm2 for N H , and 0.313 nm2 for pyridine. This size difference
has been used to explain the observation that pyridine can only
detect one type of Lewis acid site on CrpO, whereas the smaller
N H , molecule can detect two types of Lewis acid sites.
Another difference between the two molecules is their relative
basicities. In an aqueous solution, ammonia is a stronger base
than pyridine. The pKb for ammonia is around 9 while that for
pyridine is about 5 [ 101. However, in the gas phase the basicity
of pyridine is significantly greater than that of NH, [ l l , 121.
For metal surfaces it could be argued that gas-phase basicity is
a more appropriate measure of the basic strength of the adsor-
bate, as no solvation effect comes into play during adsorption
1131. For metal oxides, the situation is probably complicated.
If adsorption is via coordination to the metal cation, the gas-
phase basicity should again be the appropriate one to apply.
For example, on Fe,O, 141, pyridine is adsorbed more strongly
(desorbed more slowly) than NH, onto Lewis acid sites. I f hy-
drogen bonding or adsorption onto a BrBnsted acid site takes
place, the interaction would involve neighboring anions or hy-
droxyl groups. Such interaction has similarities with the inter-
action of a molecule with solvents in solution. For example, on
all three forms of oxidized, reduced, or sulfided molybdena-alu-
mina catalysts, NH,' ions were detected by infrared. On the
contrary, except for the oxidized form, no pyridinum ion was
detected [151. On Fe20,, hydrogen-bonded N H , causes a greater
perturbation of the hydroxyl group infrared spectrum than pyr-
idine [14].
IR STUDIES OF NH3, PYRIDINE, CO, AND NO 429

C. Effect of Sample Pretreatment

Listed in Table 1 are the different modes of bonding of N H ,


and pyridine with different transition metal oxides as inferred
from the observed IR frequencies. Confirming similar observa-
tions on these and other systems [2, 15a1, the data show the
critical role the surface hydroxyl groups play in influencing the
oxide surface properties. The extent of coverage of the surface
with hydroxyl groups in turn depends on the pretreatment con-
Downloaded by [North Carolina State University] at 01:24 16 April 2013

ditions. Different pretreatment conditions have resulted in dif-


ferent observations on the same oxide. Infrared spectroscopy
can monitor the amount of hydroxyl groups on the surface by
the intensity of their absorption peaks. In many cases, iso-
lated hydroxyls can be identified as their frequencies occur
above 3700 cm-' [ 241.
The extent of hydrogen bonding depends on the concentration
and the nature of the surface hydroxyl groups. The dependence
is different for different oxides. For example, adsorbed pyridine
hydrogen-bonded to hydroxyl groups of TiO,, as indicated by a
band around 1595 cm-l, is only observed when the evacuation
temperature is kept low (298 K ) [ 171. At this low temperature
the concentrations of adsorbed water and surface hydroxyl are
high. When these concentrations are lowered by evacuation at
473 K , there is hardly any indication of hydrogen bonding of
pyridine. Conversely, on a-Fe,O,, increasing the pretreatment
temperature increases the extent of hydrogen bonding [ 141. A
sample heated in oxygen at 973 K exhibits more hydrogen-bonded
pyridine but less surface hydroxyl than a sample heated at 713 K .
Thus, for a-Fe203, more hydrogen bonding occurs when there
are more isolated hydroxyl groups.
The presence of hydroxyl groups is not only important for hy-
drogen bonding but is also important for Brensted acidity. On a
ZnO sample that is fully hydroxylated, significant formation of
NH4+ is observed [ 41. But when the hydroxyl density is less
than half of the fully hydroxylated surface, hardly any NH4+ is
observed. On oxidized molybdena- , chromia- , rhenia- , and
tungsta-alumina catalysts, Segawa et al. found that the pyri-
dinium band intensity is increased after exposing the catalysts
to water [ 211.
Lewis acidity can be created by dehydroxylating the surface.
The formation and destruction of these sites can be observed by
monitoring both the intensities of the surface hydroxyl bands as
well as the bands of N H , and pyridine coordinated to such sites.
Downloaded by [North Carolina State University] at 01:24 16 April 2013

l a
TABLE 1 W
0

NH, and Pyridine

Modes of
bondinga
Adsor - Final oxide pretreatment
Oxides Refs. bate HB LA BA Frequency (cm-) Structure conditionsb

TiO, (anatase) 3 NH,c + 3360 OH.. .NH, or 723-773 K evacuation,


H,NH.. .O cooled in 0,, then
+ 3395, 3280, 1065 evacuation (dehy-
+ + 3170, 1605 Ti+NH,H. ..O droxylated)

TiO, (rutile) 3 NH, + 3310, 1620 OH.. .NH, or Same as above


H,NH.. .O
+ + 3400, 3350, 3250, TicNH2H.s -0
1605, 1180
+ + 3400, 3350, 3160, Ti+NH,H.. .O
1605, 1230
x
C
Z
TiO, (anatase 16 NH, + 3250 473 K evacuation (par- n
and rutile) tially dehydroxylated) *z
+ 3385, 3245, 1190 473 K evacuation (par-
U
tially dehydroxylated) x
+ 3300, 3140, 1220 673 K evacuation (com- C
plete dehydroxylated) z
n
Downloaded by [North Carolina State University] at 01:24 16 April 2013

TiO, (anatase 16 PY -k
and rutile) + 1440, 1600 473 K evacuation
+ 1440, 1605 673 K evacuation

TiO, (anatase) 17 PY + 1595 Room temperature evacua-


(3 wt%c1 im- tion
purity) + 1580 Weakly held 673 K evacuation
"liquid- like"
pyridine
+ 1606 673 K evacuation

ZrO 3 NH, + 3410, 1140 OH.. .NH3 723-773 K evacuation


i- 3380, 3280, 1610, (dehydroxylated)
1230, 1190

v2 5 18 NH3 + 1413 673 K evacuation

v205 19 NH3 + 1425 723 K in 0,


+ 1620, 1260

Cr20 3 20 NH 3 + 3370, 3275, 1610, High temerature 0 ,


(oxidized) 1255 treatment

Cr20 3 5 NH 3 + 3400, 3330 773 K evacuation

(continued)
Downloaded by [North Carolina State University] at 01:24 16 April 2013

la
TABLE 1 (continued) w
hJ

Modes of
bondinga
Adsor- Final oxide pretreatment
Oxides Refs. bate HB LA BA Frequency (cm-l) Structure condition sb

CrZo 3 5 PY + 3130, 3050, 2985, 773 K evacuation


1598, 1567, 1480,
1439, 1220, 1140,
1063, 1040
Cr,0,/Al,03 21 PY + + 773 K evacuation, cooled
(oxidized) to 473 K for adsorption

Cr,O /A1,0 21 PY + Reduction temperature


(reduced) 773 K , H,

CrO,/SiO, 22 PY -I 1612 1023 K evacuation followed


(reduced 1 by 623 K CO reduction,
then evacuation
19 NH3 + 1090, 1145 High temperature 0, treat- R
Moo3 C
+ 1425 ment z
n
Moo3 43 + 1613, 1250 673 K in 0,.cooled under *z
+ 1450 vacuum to 93 K for ad- U
sorption ; sample has x
been pre-exposed to C
z
water D
IR STUDIES OF NH3, PYRIDINE, C O , A N D NO 433

X
4c
c,
.r(

m
00
gu
6
c
.r(
c
cd
v

cd
3
s c
s
.r(

5 s
X P
Downloaded by [North Carolina State University] at 01:24 16 April 2013

k
0
0 N
m z

2
a
9
3
c
0

d d d l-l

+ + +
+ + + t + + +

m
z h h h R h
z a a a a a

m In In rl rl 4
N rl d N N N

h h
Downloaded by [North Carolina State University] at 01:24 16 April 2013

TABLE 1 (continued)

Modes of
bondinga
Adsor- Final oxide pretreatment
Oxides Refs. bate HB LA BA Frequency (cm-) Structure conditionsb

Re,O ,/A1,0 21 PY + H, reduction at 543 K


(reduced)

3 19 NH3 + 1620, broad 723 K in 0,


bands at
1100-1300

Fez0 3 14 NH3 + 3365, 3260, 713 K in 0,


1610, 1218,
1176
Perturbation of
OH group spec-
tra on Fe,O,

3 14 PY + 1592 973 K in 0,

COO 3 NH, + 1130 OH.. .NH3 493 K evacuation


-+ 3420, 3380, H,NH.. .O
3310, 1580,
1050
Downloaded by [North Carolina State University] at 01:24 16 April 2013

NiO 3 NH, + 3420, 3380, 3310, H,NH 0


... 493 K evacuation (par-
+ 1140 OH.. . N H , tially dehydroxylated 1

NiO 3 NH3 + 1610, 1180 843 K evacuation (com-


plete dehydroxylated)

ZnO 3 NH,~ + + 3420, 3380, Zn+NH,H.. .O 773 K evacuation


1630, 1220

ZnO 4 NH, + 3400, 3320 723 K , O , , cooled to room


+ 3350, 3275, 1580 temperature, evacuated
ND 3 2480 ND2 873 K , 0,, cooled to room
temperature, evacuated

ZnO 19 NH3 + 3335, 1625 723 K , 0,,cooled in 0,


+ 3350, 3275,
1610, 1255
+ 3225, 3175 Hydrogen -
bonded NH,
groups

aHB = hydrogen bonding, LA = Lewis acid, B A = BrBnsted acid.


bAdsorption at room temperature unless specified.
CFormation of N H , is inferred from the increase in the OH band intensity.
436 KUNGANDKUNG

On TiO,, Primet et al. I241 reported two types of sites created


through evacuation at high temperature. Site (a) is formed by
dislodging coordinatively bonded molecular water :

H
\ /H
0
Downloaded by [North Carolina State University] at 01:24 16 April 2013

Site (b) is formed when the temperature of evacuation is raised


to 673 K :

OH OH 0

The infrared frequencies of the N H , coordinated to the two sites


are different, being at 3385, 3245, and 1190 cm-' for Site ( a ) ,
and 3330, 3140, and 1220 cm-' for Site ( b ) .
The pretreatment conditions influence other factors besides
the hydroxyl groups. The types of surface oxygen generated
as a function of pretreatment conditions and their role in deter-
mining surface acidities are of interest. The following results
may be interpreted as the effect of oxygen. Rochester et al.
[14] reported that for a-Fe,O, heated in oxygen at 713 K , only
a weak CO, band at 1320 cm-I is seen on adsorbing CO,. How-
ever, for a-Fe20, heated in vacuo at 623 K for 4-6 h , CO, ad-
sorption results in several infrared bands assigned to carbonate,
,
carboxylate, and bicarbonate species [ 11. a-Fe,O heated at
713 K in oxygen interacts with pyridine coordinatively, but heat-
ing at 973 K in oxygen results in no evidence of Lewis acid sites.

111. CO A N D NO

The chemisorption of CO and NO has been commonly used


to probe for the nature of the surface transition metal ions.
The extensive knowledge pertaining to the interaction of these
molecules with metal atoms and ions makes their use attractive.
Attempts have been made to use these molecules to deduce
the charge, the degree of coordination unsaturation, and the
IR STUDIES OF NH,, PYRIDINE, C O , AND NO 437

symmetry of the surface metal ions. These attempts are reviewed


and discussed in this section. . The bonding of these molecules to
the surface will be first discussed in general terms, to be followed
by an individual examination of the different theories advanced to
understand their infrared absorption. Finally, various examples
will be discussed to demonstrate how information regarding the
Lewis acid sites can be obtained. Tables 2 and 3 list the various
infrared absorption bands of CO and NO, and are useful in un-
derstanding the structural make-up of the Lewis acid sites that
Downloaded by [North Carolina State University] at 01:24 16 April 2013

these molecules are bonded to.


The molecular orbitals of NO and CO are very similar. NO has
one more electron than CO that occupies an antibonding n*-orbi-
tal. They both form o-bonds with the oxide surface through their
lone pair electrons at N o r C (5a-orbital). In addition, electrons
from the surface cation can flow back to the antibonding n*-or-
bital of the adsorbate in what is known as n-donation. The man-
ner that NO or CO bonds to the surface cation depends on the
charge and coordination symmetry of the metal cation. The ratio
of o- to n-bonding can be qualitatively seen from infrared spec-
troscopy. ?r-donation causes an electron to fill the antibonding
orbital, weakening the N - 0 or C - 0 bond, and thus lowering
the corresponding infrared frequency.
Generally, NO adsorbs more strongly than CO on transition
metal oxides. For example, NO chemisorbs on Fe304 to about
monolayer coverage at 273 K and about 10 kPa [ 25, 261, when
the C O uptake is only about 10% of the NO uptake [ 271. On re-
duced Mo03/A12031281, Coo-MgO 1291, and CoO/SiO, 1301, NO
is adsorbed significantly more than CO.

A. Bonding of CO to Surfaces

The ground state for CO is z+ with the electronic configura-


tion of lo 20 30 40 I n 4 5 a 2 [ 311. Theoretical calculations in-
dicate that the effective charge on carbon is positive and that on
oxygen is negative [32, 331. However, using molecular beam
electric resonance spectroscopy, Muenter determined that CO has
a polarity of C-O+ and a very small dipole moment of 0.1222 D
[ 341. This is due to the fact that the lone pair electrons in the
5o orbital reside at the outside of the carbon atom.
Listed in Table 2 are CO absorption bands on different tran-
sition metal oxides. On many oxides a CO IR band is found at
frequencies substantially higher than the gas-phase frequency
of 2143 cm-l. This phenomenon was first noticed by Eischens and
Pliskin on NiO 1511. Since then this high frequency band has
been the subject of intensive studies [ 52-56, 56al.
Downloaded by [North Carolina State University] at 01:24 16 April 2013

&
TABLE 2 w
OD

co
Frequency Final pretreatment
Oxides Ref. (cm-l) Structure conditions

TiOz (anatase) 17 2188 Ti4+ cus site 473 K evacuation


2208. 2188 Ti4+ cus site 573 K evacuation

TiO, (anatase) 35 2185 Ordinarily adsorbed CO 673 K, 0,, followed by


2115 An intermediate form leading to 673 K evacuation
bicarbonate formation

v3 + / ~ i ~ 36 2174, 2187 I 0 873 K in CO, then evacu-


' 4 ated
I ,c
' 3+
-0-v-o-
/I
co
D l x
0 C
z
0
37 2190, 2179, 2184 On Cr2+ CO reduction at 673 K , *z
evacuated U
x
C
37 2200 On Cr3+ 1073 K evacuation, adsorp- z
n
tion at - 10C , 80 torr CO
Downloaded by [North Carolina State University] at 01:24 16 April 2013

Cr(II1) /SnO, 38 2195 673 K evacuation

Reduced C r O , / 39 2189, 2184 On C r 2 + 573 K CO reduction, cooled


SiO , in CO to 323 K , evacu-
ated, adsorption at 323 K

CrO,/Al,O, 40 2200 cm- On cr3+ 623 K in 0,,evacuated


Tail of 2200 ern- On Cr2+

Reduced CrO,/ 22 2191 On cr3+ 1023 K evacuation followed


SiO , 2181, 2186 On Cr2+ by 623 K CO reduction,
then evacuation

Mn(I1) /SiO, 41 N o adsorption 773 K , 5 X torr

Mn( 11) / S O , 41 2190.5 773 K. torr

o/ i
I
/
Mn /SnO 38 2200 Weakly adsorbed CO 693 K in 0,
1590, 1320, 1020 Bidentate CO,- bonded to Mn

(continued)
Downloaded by [North Carolina State University] at 01:24 16 April 2013

TABLE 2 (continued) P
sr
0

Frequency Final pretreatment


Oxides Ref. (cm-l) Structure conditions

Fe(I1) /SiO, 41 2172 0 773 K evacuation at 5 X


I / / / torr or torr
Ic
1 /2+
-0 -Fe - 0 -

Fe( 111) /SnO, 38 1590, 1305, 1030 Bidenrate C03- bonded to Fe 693 K in 0,
Co(I1) /Si02 41 2177 0 0 1173 K evacuation at 5 X
14 I 4 torr
IC I C
1 /2+ I /2+
-0-co-0- -0-co-0-

3
2178, 2188 1173 K evacuation at %lo- a3
P torr
I C b
1/2+ z
-0- co-c=o 3
/I T
-4
p I z
3
Downloaded by [North Carolina State University] at 01:24 16 April 2013

21 95 1173 K evacuation at -lo-'


torr

NiY 42 2217 Site I1 or 111 of zeolite 773 K evacuation

NiO (nearly 44 2125, 2075 ,c= 0 ,C=O 473 K evacuation


stoichiometric)
Ni' andlor Ni---C=O
"cz s 0 '..c=o

1910, 1880 Ni
',CEO and/or Ni---- C S O
Ni'

1730, 1160 O r g a n i c -like carbonate


1655, 1550, 1280 Bulk carbonate groups
1630, 1318 B i d e n a t e carbonate
~~

(continued)
Downloaded by [North Carolina State University] at 01:24 16 April 2013

44 2

8+

zr
a0
ln
o m 0
c-d
dlN
I
N

d
%
z
a
c

N
d
0

&i
KUNG AND KUNG
Downloaded by [North Carolina State University] at 01:24 16 April 2013

NiO /SiO 97 21 90 Ni'+ 673-873 K H, reduction


2150 N i+ m
c3
2080-2120 Nis+ C
E
M
CUY 46 2160 On Cu+ CO reduction at 723 K m
0
Cu( 11) /SnO, 38 2121 638 K in 0, rrl
z
X
CuO /SiO, 47 21 70 co+ 393 K evacuation
2140 cd
1675 Bicarbonate +c
1585 Carbox ylate E
U
c1
1350 Carbonate z
m
Oxidized Cu 48 21 35 18 torr minimum oxidation,
c1
film 2113 then evacuation 0

CUO/Y -A1,03 49 21 30 cu+- C E O 773 K in 0,evacuation at 9


z
293 K U
z
0
ZnO 50 21 92 673 K evacuation, 0, at
40 torr , then evacuation
Reduced MOO3 / 95 2190 Mo4+ (possibly 3+ or 2+) 773 K and higher H, re-
AIZO 3 duct ion
Reduced MOO / 98 21 90 773 K CO or H, reduction
A1203
sr
sr
w
444 KUNGANDKUNG

TABLE 3
Infrared Frequencies of Adsorbed NO

Frequencies
Species Solid n (cm-l) Ref.

NO Gas 1876 59

NO Condensed matrix 1872, 1883 59


Downloaded by [North Carolina State University] at 01:24 16 April 2013

cis- (NO 1 Condensed matrix 1768, 1776


(asym 1
1862, 1866
(sym)

trans- (NO) Condensed matrix 1740

,,NO C r in NaY zeolite 3 1900, 1775 60


n+'
M .\'. C r in NaX zeolite 3 1895, 1770 60
'. NO C r in NaX zeolite 2or 3 1875, 1750 60

Reduced CrO,/SiO, 1875, 1745 61

Cr203 3 1875, 1745 62

Reduced C r O , 1865, 1735 63

Reduced C r O ,/SiO, 2 1865, 1747 64

Reduced CrO,/SiO, 3 1880, 1755 64

Cr(II1) /SiO, 3 1880, 1753 65

Cr(I1) /Si02 2 1875, 1745 65

Reduced CrO,/SiO, 2 1865, 1747 66

Reduced Cr0,/Si02 3 1880, 1755 66

,
Reduced CrO /SiO, 3+ 1887, 1765 66

(continued)
IR STUDIES OF NH,, PYRIDINE, CO, A N D NO 445

TABLE 3 (continued)

Frequencies
Species Solid n (cm-') Ref.

C r 3 + on A1,O surface 3 1940, 1820 40

Cr2+-Cr2+ or Cr2+-Cr3+ 2/3 1755 40

c r 3+- c r 3+
Downloaded by [North Carolina State University] at 01:24 16 April 2013

3 1775 40

Reduced MoO,/ZrO, 1810, 1710 67


(strong)

Reduced MOO /SiOz 1810, 1710 67


(weak)

Mo/AI,O, (high tempera- <6 1800, 1695 68


ture evacuation) ((1% Mo
adsorb 1

Reduced MOO /A120 1817, 1713 69

W (CO) 6-A1,0 3 1795, 1685 70

Reduced Fe,O /SiO , 1910, 1810 71


Oxidized Fe,03/Si02 1750, 1700 71

Fe2+/zeolite Y 1917, 1815 72

CO'+/S ~ O , 1874, 1785 32


Reduced Co2+/Sn02 1840, 1765 32

CoAl,O4 1860, 1780 68

c0304 1865, 1780 68

Co/Si02 1875, 1797 73


C0/A1203 (52% CO) 1860-1870, 68
1780-1795
Co2+/zeolite Y 2 1910-1830 74
(continued)
446 KUNGANDKUNG

TABLE 3 (continued)

Frequencies
Species Solid n (ern-') Ref.

M~+-NO Reduced Cr0,/Si02 1800 61

Reduced CrO, /Si02 2 1815 66


Downloaded by [North Carolina State University] at 01:24 16 April 2013

Reduced CrO,/Al,O, 2 1880 40

,
Reduced CrO /A1,0 , 3 1905 40

c r 3 +in ~ 1 ~ surface
0 , 3 1875 40

,
Reduced Fe20 / S O 2 1750 71

Reduced Fe20,/Si02 2 1830 71

Oxidized Fe203/Si02 3 1820 71

,
0 xidi zed Fe20 / Si0 1810 67

Reduced Fe20,/Si02 1810 67

Oxidized Fe3+/Sn02 3 1832 32

Reduced Fe3+/Sn02 3 1770 32

Reduced Fe3+/Sn02 2 1720 32

Fe2+/zeoliteY 2 1845 72

Fe2+/zeolite Y 2 1870 72

Fe2+/zeolite Y 2 1767 72

Fe2+/zeolite Y 3 1860 72

CoA1204 2 1875 68

c0304 1850 68

(continued)
IR STUDIES OF NH,, PYRIDINE, CO, A N D NO 447

TABLE 3 (continued)

Frequencies
Species Solid n (cm-') Ref.

Ni2+/Sn02 2 1870 32

N 10 1820 76

,,
Downloaded by [North Carolina State University] at 01:24 16 April 2013

Reduced NiO /A1,0 2 1870-1880 97


octahedral N i 6+ 1900-1915

Oxidized Cu 2+/ SnO , 1880 32

CuO /Al,O , 1875 77

C uAlO
, 1900 77

Cu2+ dispersed in A120, 1920 77

Reduced CuO /A1,0, 1780 78

Reduced Cu/zeolite Y 1740 78

zfl, 1900 76

M"+--No
+ Reduced CrO,/SiO, 1810 64

CuO /SiO, 1890 47

ZrO 2260 76

M"+-No- Fe20 3 1595 33

Reduced Fe,03/Si0, 1735 67

Reduced Cr,O, 1305 79

By comparing the band frequencies for a number of metals, ox-


ides, and ion-exchanged zeolites, Lokhov et al. [ 541 substantiate
earlier proposals that the frequency for M2+C0 is above 2170 cm-',
M'CO in the region of 2120-2160 cm-', and M'CO below 2100 cm-l.
448 KUNGANDKUNG

They argued that for Mo, dative and back bonding are present,
and the importance of back n bonding is seen by the stretching
frequency which is located below that of gas phase CO ( 2 1 4 3
.
cm-l) A s the metal becomes more oxidized, less electron is
available for backbonding and the increased contribution from
o-bonding is reflected by the increased CQ stretching frequency.
Hence, at high oxidation states, the metal acts as a weak Lewis
acid.
Hush and Williams [ 551 and Larsson et al. [ 561 have attempted
to explain the high frequency CO band in more mathematical
Downloaded by [North Carolina State University] at 01:24 16 April 2013

terms. Hush and Williams, using a Finite Field CNDO-I1 treat-


ment, examined the effects of strong electric field on equilibrium
nuclear configurations, vibrational frequencies, and vibrational
transition intensities. They argued that for the three canonical
representations of CO , (1) would be favored by a positive field
along the C-0 axis and , conversely , ( 3) would be favored by a
negative field along the same axis. They proposed that for a

weakly adsorbed CO molecule that has its C-0 axis perpendicu-


lar to the oxide surface, a negative shift (to lower frequency)
in the CO stretching frequency is due to adsorption of the car-
bon end on an anionic surface site, o r the oxygen end on a cat-
ionic site. A large positive shift in the infrared frequency would
indicate adsorption of the carbon end on a cationic site or the
oxygen end on an anionic site, (There i s yet no experimental
evidence that CO adsorbs with the oxygen end.) They demon-
strated a reasonable correlation of their theoretical work with
the experimental results of Angel1 and Schaffer [ 4 2 ] . Harrison
et al. 1381 studied CO adsorption on cations exchanged onto
tin(1V) oxide gels. By assuming Cu to be in the Cu+ form and
Mn to be in the Mn3+ form, they too demonstrated that the CO
stretching frequencies are dependent on the electric field strength
in the manner predicted by Hush and Williams.
Larsson et al. [ 561 , however, have difficulty in accepting the
fact that for high positive field along the C-0 axis, the CO dis-
tance , as predicted by Hush and Williams , tends to increase after
a minimum is passed. They argued that the high positive electric
field favors the canonical form -CEO+, resulting in a shorter in-
stead of a longer CO bond. They in turn proposed a purely elec-
trostatic model to understand the high frequency CO band. They
IR STUDIES OF N H , , PYRIDINE, CO, AND NO 449

argued that coordination of CO is to the metal ion a s is seen by


the strong dependence of the stretching frequencies on the na-
ture of the cation. In their model the effective positive charge
is on carbon, but the dipole moment is negative (i.e., C - O + )
due to the localization of the lone pair electrons outside the car-
bon atom. Employing the empirical relation between force constant
and interatomic distance as described by Badger [ 571, they calcu-
lated the CO stretching frequencies for different oxides, which
are found to deviate from experimental results by no more than
Downloaded by [North Carolina State University] at 01:24 16 April 2013

three standard deviations [ 561.

B. Bonding of NO to Surfaces

Theoretical analysis of the infrared band of NO adsorbed on


solid oxides has not been available. However, there exists a
comprehensive theoretical analysis of infrared absorption bands
of NO as a ligand in transition metal complexes in solution by
Enemark et al. [ 581. These authors used the molecular orbital
correlation method to analyze the structure, bonding, and re-
activity of the metal nitrosyls. They considered M(N0)n as a
single entity and studied the perturbation of this moiety by the
field imposed by other ligands attached to the metal. The ground
state of the M-NO group may vary from a linear to a strongly
bent configuration. They pointed out that the energies of the
metal d-orbitals and the a*-orbitals of the NO ligand are simi-
lar. This has the consequence that minor changes in the metal
atom or its environment may cause dramatic changes on the ef-
fective charge on the NO group. Thus they cautioned that U N O
is not sufficient to predict the linearity of the M-N-0 group.
They did, however, arrive at some generalizations. It appears
that complexes with VNO > 1850 ern-' always have a linear MNO
grouping. All six- and five-coordinated {MNOIX groups are
linear when X 5 6 (where X is the number of electrons on the
metal when the nitrogen ligand is formally considered N = O + ) .
Four-coordinated complexes with X = 10 are linear in tetrahedral
geometry and bent in square planar geometry. Applications of
these generalized predictions to NO adsorbed on solids have
been made and will be examined in detail later.
Table 3 lists the frequencies of the NO bands. The structure
of the M(NO), complexes are classified as suggested by the au-
thors. It is seen that NO has a tendency to adsorb in pairs.
Zecchina et al. [64] ascribed this to the fact that a stability fac-
tor is gained through the mutual interaction of the unpaired elec-
tron on each NO. Whether the pair of adsorbed NO molecules
450 KUNGANDKUNG

exists in the form of a dimer or dinitrosyl is still under substan-


tial discussion, and in some studies the investigators stated that
they do not attempt to make any distinction between the two forms
71, 721. Zecchina et al. [ 6 4 ] favor a strongly coupled dinitrosyl
complex. They based their argument on infrared absorption of
coordination complexes containing NO ligands and the total ab-
sence of absorption in the 1250-1350 cm- region, which excludes
the presence of a hyponitrite species N 2 0 2 - . Kazusaka and Howe
1701 also favor the dinitrosyl species for their alumina-supported
Downloaded by [North Carolina State University] at 01:24 16 April 2013

catalysts prepared from molybdenum hexacarbonyl , tungsten hexa-


carbonyl, and chromium hexacarbonyl. The adsorbed NO pa?-
they observed is stable to quite a high temperature, and the two
nitrosyl bands have approximately the same intensity.
On the other hand, Kugler et al. [ 611 compared the coupled
nitrosyl bands on a silica-supported chromia catalyst with those
from N 2 0 2 dimer and found that the ratio of the symmetric to
asymmetric stretching are very similar. Thus they concluded
that the bands are from the cis-N202 dimer. Millman and Hall
1941 also arrived at the same conclusion from their study using
isotopically labeled NO. In this study a reduced molybdena-alu-
mina catalyst was first half-saturated with 1 5 N 0 . The catalyst
was then exposed to 1 4 N 0 . Examination of this catalyst by in-
frared spectroscopy yields no bands attributable to a %O-l4NO
complex. Since such isotopic mixing is expected for dinitrosyl
complexes, the observation supports a dimeric species. The same
observation was made independently by Peri [ 9 5 , 961 on a reduced
molybdena-alumina catalyst. However, Peri cautioned that this
result could also be achieved (though less likely) for a dinitrosyl
species if the adsorption sites differ greatly in their binding en-
ergies for NO or in their accessibility.
It is also noticed that on a substantial number of occasions,
NO absorbs at frequencies higher than 1850 cm-. Enemark et al.
[ 581 have indicated that in solution, metal nitrosyl compounds
with infrared absorption frequencies > 1850 cm- all possess
linear M-NO groups.

C. Surface Properties Deduced from


Adsorption of CO and NO

Many interesting properties of Lewis acid sites can be deduced


from the infrared spectra of NO and CO adsorbed on transition
metal oxides. The validity of the proposed properties hinges on
the interpretation of the infrared spectra. Unfortunately, some-
times numerous bands arise from the adsorption of a single mole-
cule, and unambiguous assignment of such bands becomes a chore.
IR STUDIES OF NH,, PYRIDINE, CO, A N D NO 451

In such cases, information from other measurements, such as


Mossbauer spectroscopy, UV reflectance, quantitative measure-
ment of the amount adsorbed, temperature programmed desorp-
tion, electron paramagnetic resogance, and others may be help-
ful in the identification of infrared bands.
Listed below are some of the possible information derivable
from NO or CO adsorption studies on transition metal oxides.
They will be discussed in detail with examples:
Downloaded by [North Carolina State University] at 01:24 16 April 2013

( 1 ) The importance of pretreatment and loading


( 2 ) The nature of the adsorption site:
( a ) The oxidation state of the cation
(b) The geometry of the adsorbate-surface complex

The importance of the pretreatment conditions with respect to


the surface properties cannot be overstated. With respect to
NO and C O , pretreatment affects the presence of surface oxy-
gen or hydroxyl that cause reactions to occur after adsorption.
When a metal oxide is present in the oxidized form, sometimes
CO can react to form carbonate and bicarbonate 138, 441, and
NO can react to form nitrite and nitrate [ 38, 47, 801. CO can
also react with surface hydroxyl groups to form formate [ 811,
The critical dependence of surface properties on pretreatment
is clearly illustrated by Rebenstorf and Larsson [ 41, 451. It was
shown that after heating samples of metal ions supported on SiO,
to 773 K under a vacuum pressure that ranged from around lo-'
to 5 X t o r r , substantial changes in the infrared absorption
spectra were observed. For Mn(I1) / S O , samples heated at 5 X
torr, no CO adsorption is detected. However, if the pre-
treatment i s performed at around lo-' torr , infrared absorption
spectrum of MnCO complex is detected. The infrared spectra of
CO adsorbed on Co/SiO, heated at lo-' torr and Co/SiO, heated
at 5 X 10- torr vacuum are also different.
Amberg and Seanor 1821 found that the stretching frequency
of CO depends on the pretreatment of ZnO. ZnO evacuated at
40OoC followed by an oxygen treatment at 4 O O O C yields a CO band
at 2212 cm-'. If. after the oxygen treatment, a brief evacuation
is performed and the sample is cooled down to room temperature
in vacuo, the CO frequency shifts to 2187 cm-l.
Related to the problem of pretreatment is sample loading.
Topsae et al. [ 681 found with infrared spectroscopy that changes
in the environment of the cobalt ions can be detected as a func-
tion of loading. In samples of low loading ( 5 2 %Co) , Co ions are
located in the octahedral vacancies of A1,0,, and the adsorbed
NO exhibits an infrared spectrum that in general is similar to
452 K U N G AND K U N G

that for CoA1,0,. The changes in the infrared spectra on de-


creasing Co2+loading suggest that the cobalt in the more dilute
samples have less electrons available for back-donation to the
adsorbed NO. A s the cobalt concentration is increased to >2%,
the infrared spectra becomes similar to that of NO adsorbed on
Co30,.
Adsorption of small molecules on transition metal oxides also
yields information about the position of these sites with respect
to the surface. High temperature treatment on reduced CrO,
Downloaded by [North Carolina State University] at 01:24 16 April 2013

converts the surface ions into ?'better shielded species" [ 221.


The result of such conversion is that CO cannot interact with
these sites any longer and NO interacts to form a monomer in-
stead of a dimer. On the other hand, for CoA1204.pretreat-
ment at a higher temperature increases NO adsorption 1681. At
high temperature the fraction of Co located in the octahedral
site is high [ 8 3 ] , and NO is adsorbed readily on Co located in
octahedral sites [ 841.
The nature of the active site for adsorption is governed by
the composition of the cations in the oxide, their charge, their
coordination symmetry, and their degree of coordination unsat-
uration. The formal oxidation state of the cation may be deter-
mined with infrared spectroscopy. A s mentioned earlier, Lokhov
et al. [ 541 bracketed the frequency range for M'CO , M'CO , and
M*CO. Rebenstorf et al. found excellent correlation between
CO frequencies and the formal charge on the metal cation [ 361.
Even a small remnant of Cr3+ on reduced CrO,/SiO, can be de-
,
tected [ 641 because the infrared frequencies of Cr3+(NO) and
C r 2 + ( N 0 ) , are different.
The oxidation state of the same cations can also be probed by
its inability to adsorb NO or CO in certain oxidation state. NO
adsorbs strongly on Fez+ but only weakly on Fe3+ [ 8, 71, 851.
Similarly, NO adsorbs strongly on partially reduced tungsten
oxide but not on fully oxidized W6+ [ 8 6 ] . Cu+ adsorbs CO
strongly, Cu2+ weakly [ 491. The reason for the inability of
Fe3+and W6+ to adsorb NO may be related to their degree of
coordination unsaturation rather than their electronic struc-
ture.
The degree of coordination unsaturation of the metal cation
as well as the coordination symmetry can be deduced from the
infrared frequencies of the adsorbed NO or CO. Yuen et al.
[ 711 , using infrared and Mossbauer spectroscopy, determined
that Fez+ of high coordination adsorbs NO to form a Fe2+-N0
complex that is more linear than Fe2+ of low coordination.
Tanabe et al. observed on reduced Fe203/Si02two infrared
bands of adsorbed NO at 1810 and 1735 cm-' [67], which were
IR STUDIES OF NH,, PYRIDINE, C O , A N D NO 453

assigned to adsorption on Fez+ in sixfold and fourfold sites, re-


spectively [711. When Fe2+ is only located in octahedral sites
as in a SnO, supported sample, only the 1810 cm-' band is de-
tected 167, 711.
Segawa et al. [ 721 , studying NO adsorption on Fe2+-exchanged
zeolite Y catalysts, assigned the 1767 cm-' band to NO adsorbed
on Fe2+located in Site 111 of the zeolite, where Fez+ is coordinated
to five oxygen anions. Four of the oxygen ions are in a square
planar configuration. NO stretching band at 1845 cm-' is as-
Downloaded by [North Carolina State University] at 01:24 16 April 2013

signed to Fe2+ at Site 11, where the cation is coordinated to three


oxygen anions in a triangular configuration. Ghiotti e t al. [66]
distinguished four types of C r 2 + in their calorimetric and spec-
troscopic studies of NO adsorption on silica- supported chromium
ions. The most exposed Type A C r 2 + is coordinated to only two
oxygen surface atoms. Type B Cr2+ is coordinated to three oxy-
gen anions. The most shielded C r 2 + (C, and C,) have three and
four oxygen ligands.
NO has a tendency to adsorb in pairs on transition metal ox-
ides. The ratio of the intensities of the asymmetric to symmetric
stretching of the pair is related to the angle between the two
M-NO bonds [ 871. This relationship has been widely utilized to
understand the geometry of metal nitrosyl complexes [ 68, 73, 741.
Chemisorption of CO can also provide information on the co-
ordination symmetry of the surface metal ions. Rebenstorf and
Larsson [41] observed that at low temperatures (-145OC), some
divalent, coordinatively unsaturated cations supported on silica
adsorb up to four CO molecules per ion. At room temperature,
usually one or two CO molecules are coordinated to these surface
cations. They compiled a list of the number and relative intensi-
ties of infrared bands of CO ligands of transition metal complexes
having one to four CO ligands to help interpret the adsorption
spectra for divalent Mn, N i , C r , Fe , and Co supported on SiO,
1411. A similar approach was also applied to trivalent metal cat-
ions on silica [ 361.
Finally, a surface dinuclear chromium complex has been de-
tected by CO adsorption at low temperature, which yields a com-
plex of [22, 371

This complex has been used to explain the 2095 cm-' band [ 221.
454 KUNGANDKUNG

IV. COADSORPTION O F SMALL MOLECULES

Additional o r sometimes different information can be obtained


when different molecules are coadsorbed onto surfaces, When
two different molecules are adsorbed together, the following may
occur :

(1) They do not affect each other.


( 2 ) They affect the strength andlor the type of bonding of each
other. This is detected as changes in the frequencies, in
Downloaded by [North Carolina State University] at 01:24 16 April 2013

the extinction coefficients, and lor as appearance or disap-


pearance of bands.
(3) They react with each other.

Case (1) does not introduce new insight on the nature of the
surface and will not be discussed further. The other two cases
are more interesting. They will be discussed using a number of
specific examples.
The first example is Mo/A120, 1861. When NO is adsorbed first,
followed by pyridine, the absolute integrated intensity of the NO
bands remain constant, indicating that NO is not displaced by
pyridine. The NO bands, however, are broadened and shifted
to a lower frequency by about 50 cm-l. The shift in frequency
implies that even though pyridine is probably adsorbed on the
alumina, its presence causes increased electron back-donation
into the antibonding orbitals of the NO adsorbed on the molyb-
denum ion. The ratio of the integrated intensities of the asym-
metric to symmetric stretch increased with pyridine coverage.
This suggests that coadsorbed pyridine somehow hinders the
dimeric NO-NO interaction.
On ZnO, when CO is adsorbed onto a hydrogen-covered sur-
face, the OH band shifts to a higher frequency and two new Zn-H
bands appear. The new Zn-H bands represent Zn-H complexes
with one or both of its neighboring Zn atoms coordinated to CO
1501. When CO and COP are coadsorbed on ZnO 1881, the CO fre-
quency is shifted from 2183 to 2212 cm-. This shift reflects the
fact that CO, enhances Lewis acidity of the Zn ions that initially
possess two coordinatively unsaturation sites.
On a reduced Cr03/Si02sample with preadsorbed NO, adsorp-
tion of NH, or pyridine lowers the NO stretching frequency, Both
N H , and pyridine donate electrons to the solid and thus increase
back-bonding contribution to the n*-orbitals of NO. When CO is
adsorbed onto CrII with preadsorbed NO, it becomes an electron
acceptor and its stretching frequency is lowered with a simulta-
neous increase in the NO stretching frequency 1891.
IR STUDIES OF N H 3 , PYRIDINE, CO, AND NO 455

Coadsorbed molecules sometimes react. On ZnO and Cu- ZnO


when H , and CO are coadsorbed, two small bands are observed
at 2661 and 2770 cm-l, even at 273 K 1901. These bands are
more prominent on Cu-ZnO than ZnO, and are ascribed to formyl
formation. The 2770 and 2661 cm- are assigned to V C H and two
other bands at 1520 and 1370 cm- to v c s [ 911.
At elevated temperatures on CuO/Al,O, [ 781, Cr,O,/Al,O, [ 781,
Cr,O,/SiO, [ 921, and unsupported chromia 1931, coadsorption of
CO and NO leads to the formation of isocyanate. The isocyanate
Downloaded by [North Carolina State University] at 01:24 16 April 2013

species is adsorbed on C r on the unsupported catalyst and mostly


on C r for Cr,O,/SiO, catalyst. Its infrared band is located at
around 2210 cm-l. On C r 2 0 3 /A1203, a rapid migration of the
isocyanate to alumina occurs leading to bands a t 2262 and 2242
cm-l.

V. CONCLUSION
From a review of the recent literature, it is clear that proper
interpretation of the infrared spectra of adsorbed N H , , pyridine,
CO, and NO can provide important information of surfaces of
transition metal oxides. The spectra of adsorbed N H , and pyri-
dine provide information on the presence of Lewis and /or Br6nsted
acid sites and the properties of the surface hydroxyl groups. The
spectra of adsorbed CO and NO are useful for probing the surface
Lewis acid sites, which are mostly surface-exposed cations. The
formal oxidation state, the coordination symmetry, and the degree
of coordination unsaturation of the surface cation can be deduced.
However, while qualitative conclusions can be drawn with confi-
dence, quantitative interpretation is seldom achieved. For ex-
ample, while the formal oxidation state of a surface cation can be
deduced from the vibrational frequencies, it is much more difficult
to deduce the true ion charge o r s m a l l changes in the t r u e ion
charge. Similarly, the angle between the adsorbate and the sur-
face, or other information on the configuration of the adsorbate,
may not be determined unequivocally.
Another important point is the strong dependence of the s u r -
face properties on the pretreatment condition. The density and
the type of acid sites, the degree of hydrogen bonding with ad-
sorbates, the coordination unsaturation of the surface cations,
and the configuration of the adsorbate can all be affected. Thus,
attention must be paid to the pretreatment conditions for proper
comparison of experimental results among investigators. When
the results are carefully analyzed, valuable information on the
surface properties can be obtained.
456 K U N G AND K U N G

Acknowledgment

Work supported by the American Chemical Society, Petroleum


Research Fund.

REFERENCES

[l] G. Busca and V. Lorenzelli, Mater. Chem., 1, 89 (1982).


Downloaded by [North Carolina State University] at 01:24 16 April 2013

[2] L . H . Little, Infrared Spectra of Adsorbed Species,


Academic, New York, 1966.
[3] A . A . Tsyganeko, D . V . Pozdnyakov, and V. N . Filimonov,
J . Mol. Struc., 3, 299 (1975).
[4] T . Morimoto, H . Yanai, and M . Nagao, J . Phys. Chem.,
80, 471 (1976).
[5] K . Morshige, S. Kittaka, S . Katsuragi, and T . Morimoto,
J . Chem. SOC., Faraday Trans. I , l_s, 2947 (1982).
[6] K . Nakamoto, Infrared Spectra of Inorganic and Coordina-
tion Compounds, 2nd e d . , Wiley-Interscience, New York,
1970.
[71 H . Knozinger, Adv. Catal., 2, 184 (1976).
181 N . S. Gill, R. H . Nuttall, D. E . Scaife, and D . W . A .
Sharp, J. Inorg. Nucl. Chern., 18, 79 (1961).
[9] A. Zecchina, E . Guglielminotti, LT Cerruti, and S.
Coluccia, J . Phys. Chem., 76, 571 (1972).
[lo] E. P . Parry, J . Catal., 2, 371 (1963).
[ll] M . Taagelpera, W. G . Henderson, R . T. C . Brownlee,
J . L . Beauchamp, D . Holtz, and R. W. Taft, J . A m .
Chem. SOC., 94, 1369 (1972).
[12] W. B . Jensen, The Lewis Acid-Base Concepts, An Over-
view, Wiley, New York, 1980.
[13] T C . Stair, J. Am. Chem. SOC., 104, 4044 (1972), and
references therein.
[14] C . H . Rochester and S. A . Topham, J . Chem. S o c . ,
Faraday Trans. I , 1_5, 1259 (1979).
[15] J. Valyon, R . L. Schneider, and W. K . Hall, J. Catal.,
85, 277 (1984).
[15a] B . Peri, J. Phys. Chem., e, 220 (1965).
[16] M. Primet, P. Pichat, and M. Mathieu, - Ibid., 75, 1221
(1971).
[17] C . Morterra, G. Ghiotti, and E . Garrone, J. Chern. Soc.,
Faraday Trans. I , I_, 2102 (1980).
[l8] M . Takagi-Kawai, M . Soma, T . Onishi, and K . Tamaru,
J . Phys. Chem., so, 430 (1976).
IR STUDIES OF NH,, PYRIDINE, CO, AND NO 457

r191 Yu. V . Belokopytov, V . M. Kholyavenko, and S . V. Geri,


J. Catal., so,
1 (1979).
[ 201 Yu. V. Belokopytov, V. A. Kuznetsov, M . K . Kholyavenko,
and S . V . Gerei, E d . , 44, 1 (1976).
1211 K . Segawa and W. K . Hall, K d . , I_s, 133 (1982).
[ 221 A . Zecchina, E . Garrone, G . Ghotti, and S. Coluccia, &
Phys. Chem., 79, 972 (1975).
[ 231 A . A . DavydovTReact. Kinet Catal. Lett., - 19, 377
(1982).
Downloaded by [North Carolina State University] at 01:24 16 April 2013

I 241 M. Primet, P . Pichat, and M . Mathieu, J . Phys. Chem.,


75, 1216 (1971).
[ 251 c.Otto and M. Shelef, J. Catal., 5,184 (1970).
261 C . R. F . Lund, J . J. Schorfheide, and J . A . Dumesic,
Ibid., 57, 105 (1979).
-
[ 271 J . E . Kubsh, Y . Chen, and J . A . Dumesic, e., 71, 192
( 1981) .
[ 281 H. C . Yao, Ibid., 70, 440 (1981).
[ 291 A . P. Hagan-. G . Lofthouse, F . S. Stone and M . A .
Trevetnan, in Preparation of Catalysts I1 (B. Delmon,
P. Grange, P. A . Jacobs, and G . Poncelet, e d s . ) ,
Elsevier, Amsterdam, 1979, p. 4177.
1301 B . Rebenstorf, Acta Chem. Scand., Ser. A, - 31, 208 (1977).
[311 -
W . M . Huo, J . Chem. Phys., 43, 624 (1965).
S. Forsen and B . Roos, Chem. Phys. Lett., 5, 128 (1970).
r321
331 E. R. Davidson, J . Chem. Phys., 46, 3320 (1967).
341 J. S. Muenter. J . Mol. Spectrosc., 55, 490 (1975).
r351 K . Tanaka and J. M . White, J . Phys. Chem., 86, 4708
(1982).
361 B . Rebenstorf, M . Berglund, R . Lykvist, and R . Larsson,
Z. Phys. Chem., N.F., 126, 47 (1981).
[ 371 B . Rebenstorf and R . Larsson, Z. Anorg. Allg. Chem.,
478, 119 (1981).
1381 G
. .P Harrison and E . W . White, 3 . Chem. SOC.,Faraday
Trans. I , 74, 2703 (1978).
[ 391 D . D . Eley, C . H . Rochester, and M . S. Scurrel, J.
Catal., 2, 20 (1973)-
[401 J. B . Peri, J . Phys. Chem., 2, 588 (1974).
1411 B. Rebenstorf and R . Larsson, Z . Anorg. Allg. Chem.,
453, 127 (1979).
1421 C . L , Angel1 and P . S . Schaffer, J . Phys. Chem., - 70, 1413
(1966).
1431 R. P. Groff, J. Catal., 86, 215 (1984).
1441 E . Guglielminotti, L. Cerruti, and E . Borello, Gazz. Chim.
Ital., 107, 503 (1977).
4 58 K U N G AND K U N G

t 451 B . Rebenstorf and R . Larsson, 2. Anorg. Allg. Chem.,


453, 139 (1979).
[ 461 Yu. A . Lokhov and A . A . Davydov, Kinet. Katal., g,
1239 (1979).
471 J. W. London and A . T. Bell, J. Catal., 31, 32 (1973).
[ 481 H . G . Tompkins and R . G. Greenler, Surf. Sci., 2, 194
(1971).
[ 491 Yu. A . Lokhov, 2. Musil, and A. A . Davydov, Kinet.
Katal., 2, 207 (1979).
F. Boccuzzi, E. Garrone, A . Zecchina, A . Bossi, and
Downloaded by [North Carolina State University] at 01:24 16 April 2013

[ 501
M . Camia, J. Catal., Ll, 160 (1978).
t 511 R. P. Eischens and W. A . Pliskin, Adv. Catal., 9, 662
(1957).
521 J. H. Taylor and C . H. Amberg, Can. J . Chem., - 39, 535
(1961).
[ 531 R. A . Gardner and R . H . Petrucci, J. A m . Chem. Soc.,
82, 5051 (1960).
541 y u . A . Lokhov and A . A . Davydov, Kinet. Katal., 21,
1093 (1980).
551 N . S . Hush and M. L . Williams, J. Mol. Spectrosc., 2,
349 (1974).
I561 R . Larsson, R . Lykvist, and B . Rebenstorf, 2. Phys.
Chem. (Leipzig) , 263, 1089 (1982).
56al J. B. P e r i , J . C a t x . , @, 89 (1984).
1571 R. M. Badger, J. Chem. Phys., 3, 710 (1935).
581 J. H. Enemark and R . D . Feltham, Coord. Chem. Rev.,
138, 339 (1974).
591 W . G . Fately, H . A . Bent, and B . Crawford Jr., J. Chem.
-
Phys., Ll, 204 (1959).
t 601 J. Pearce. D . E . Sherwood, M. B . Hall, and J. H .
Lunsford, J. Phys. Chem., 84, 3215 (1980).
[611 E. L . Kugler, R . J . Kokes, and J . W. Gryder, J. Catal.,
36, 142 (1975).
621 Naccache and Y . Ben Taarit, J. Chem. Soc., Faraday
Trans. I , 69, 1475 (1973).
631 E. L . KugKr, A . B . Kadet, and J . W. Gryder, J. Catal.,
41, 72 (1976).
641 A1 Zecchina, E . Garrone, C . Morterra, and S. Coluccia,
J. Phys. Chem., 79, 978 (1975).
1651 D . D . Beck and J. H . Lunsford, J. Catal., @, 121 (1981).
[ 661 G . Ghiotti, E. Garrone, G . Della Gatta, B . Fubini, and
E . Giarnello, K . ,so, 249 (1983).
[ 671 K . Tanabe, H . Ikeda, T . Iizuka, and H . Hattori, React.
Kinet. Catal. Lett., - 11, 149 (1979).
IR STUDIES OF NH,, PYRIDINE, C O , AND NO 459

N . Topsee and H . Topsse, J . Catal., 75, 354 (1982).


J. Valyon and W. K . Hall, Ibid., 84, 216 (1983).
A . Kazusaka and R . F . Howe, Ibid,, 63, 447 (1980).
S. Yuen, Y. Chen, J. E . K u b s c J . A . Dumesic, N . Topsee,
and H . Topstae, J . Phys. Chem., 2, 3022 (1982).
K . Segawa, Y. Chen, J . E . Kubsh, W. N. Delgass, J. A .
Dumesic, and W. K . Hall, J. Catal., 76, 1 1 2 (1982).
[ 731 B . Rebenstorf, Acta Chem. Scand., Ser. A , 31, 208 (1977).
741 K . A . Windhorst and J. H . Lunsford, J . AmrChem. SOC.,
-
97, 1407 (1975).
Downloaded by [North Carolina State University] at 01:24 16 April 2013

L. Cerruti and E . Guglielminotti, Discuss. Faraday SOC.,


52, 285 (1971).
[ 761 V . Pozdnyakov and V. N . Filimonov, Kinet. Katal., 14,
655 (1973).
771 Yu. A . Lokhov, L . N . Morozov, A . A . Davydov, and
V . V . Kostrov, z., g ,943 (1980).
1781 Yu. A . Lokhov and A . A . Davydov, E., 3,1239 (1979).
1791 A . A . Davydov, Yu. A . Lokhov, and Yu. M . Shakochikhin,
Ibid., 19, 532 (1978).
C B u s c a and V. Lorenzelli, J. Catal, 12, 303 (1981).
M. H e and J . G. Ekerdt, g., 87, 381 (1984).
C . H , Amberg and D . A . Seanor, Proceedings of t h e 3rd
International Congress on Catalysis, Amsterdam, 1965,
p . 450.
1831 C . Angelletti, F. Pepe, and P . Porta, J. Chem. SOC.,
Faraday Trans. I , 76, 244 (1980).
[ 841 H . C . Yao and M . Shelef, in The Catalytic Chemistry of
Nitrogen Oxides (R. L , Klemesch and J. G . Larson, eds.) ,
Plenum, N e w York, 1975, p. 45.
G . W . Poling and R. P. Eischens, J . Electrochem. SOC.,
113, 218 (1966).
[ 861 KTSegawa and W. K . Hall, J . Catal., c. 221 (1982).
[871 F. A . Cotton and G . Wilkinson, Advanced Inorganic Chem-
-i s t r y , 3rd e d . , Wiley-Interscience, New York, 1972, p . 697.
[ 881 J . C . Lavalley, J . Saussey, and T . Rais, J . Mol. Catal.,
17, 289 (1982).
[a91 Garrone, G. Ghiotti, S . Colucci, and A . Zecchina, J.
Phys. Chem., 79, 984 (1975).
[ 901 J. Saussey, J . 7 . Lavalley, J . Lamotte, and T . Rais, &
Chem. SOC., Chem. Commun. a 5 , 278 (1982).
911 J. C . Lavalley, J. Saussey, and T . Rais, J . Mol. Catal.,
-
17, 289 (1982).
921 J . Rasko and F. Solymosi, J . Chem. SOC., Faraday Trans.
-I , -76, 2383 (1980).
-
460 KUNGANDKUNG

[ 931 F . Solymosi and J . Rasko, J. Catal., 65, 235 (1980).


[ 941 W. S. Millman and W . K . Hall, J. Phys. Chem., - 83, 427
(1979).
t 951 J. B. Peri, E d . , g ,1615 (1982).
[ 961 J. B. Peri, Prepr., Div. Pet. Chem., Am. Chem. SOC.,
23, 1281 (1978).
t 973 B . Peri, J. Catal., 86,84 (1984).
[ 981 W . S . Millman, M. Crespin, A. C . Cirillo, S . Abdo, and
W . K . Hall, -
Ibid., so, 404 (19'79).
Downloaded by [North Carolina State University] at 01:24 16 April 2013

Vous aimerez peut-être aussi