You are on page 1of 5

Carbon 40 (2002) 145149

Surface oxides on carbon and their analysis: a critical


assessment
H.P. Boehm*
Munchen

, Butenandtstr. 5 13, 81377 Munchen


, Germany
Department Chemie, Universitat
Received 15 March 2001; accepted 15 June 2001

Abstract
The methods for the determination of various types of oxygen surface functions on carbon materials are briefly described,
and their relative advantages and problems that may arise are discussed. Acidimetric titration techniques, IR spectroscopy,
XPS, thermal desorption spectroscopy, and electrokinetic measurements are described. 2002 Elsevier Science Ltd. All
rights reserved.
Keywords: C. Infrared spectroscopy; Temperature-programmed desorption; X-ray photoelectron spectroscopy; D. Surface oxygen complexes

1. Introduction
Many properties of carbon materials, in particular their
wetting and adsorption behavior, are decisively influenced
by chemisorbed oxygen. Oxygen in the surface oxides can
be bound in the form of various functional groups which
are similar to those known from organic chemistry. This
article deals mainly with high surface area carbon materials which consist predominantly of sp 2 -hybridized carbon
atoms (non-graphitized, turbostratic carbons). The surface
of such carbons is heterogeneous, it consists of the faces of
graphene sheets and of edges of such layers. The edge sites
are much more reactive than the atoms in the interior of
the graphene sheets, and chemisorbed foreign elements, in
particular oxygen, are predominantly located on the edges.
The surface oxides decompose to CO 2 and CO on
heating to high temperatures. Highly reactive sites remain
on the carbon surface which have free-radical character to
some relatively small extent [1,2]. After cooling to room
temperature, they can react with oxygen (air) or water
vapor, giving new surface oxides. A continuous, very slow
oxidation of the surface occurs after a first, rapid
chemisorption of oxygen. The presence of water is essential for this aging of the carbons [3,4]. Aging is reduced
after prior chenmisorption of hydrogen, e.g., at 9508C
[5,6]. Much more oxygen is chemisorbed at elevated
temperatures, e.g., 3004208C. Alternatively, surface ox*Fax: 149-89-2180-7492.
E-mail address: hpb@cup.uni-muenchen.de (H.P. Boehm).

ides can be created by treatment with liquid oxidants, e.g.,


aqueous solutions of H 2 O 2 , NaOCl, (NH 4 ) 2 S 2 O 8 , AgNO 3 ,
H 2 PtCl 6 , etc., at 201008C. Oxidation with HNO 3 is often
used because its oxidizing properties can be controlled by
concentration and temperature.
The various aspects of the surface chemistry of carbon
materials have been described in detail [1,79]. The
present review discusses the most frequently used methods
for the characterization of surface oxides.

2. Titration methods

2.1. Titration of acidic surface functions


The surface oxides on a carbon can have acidic as well
as basic properties and can be conveniently determined by
titration methods. Basic surface character and anion exchange properties are found when a carbon surface,
cleaned at ca. 9508C or higher in vacuo or under an inert
gas is exposed to air and aqueous acid (or water) after
cooling to room temperature. Oxygen and acid (or water)
are chemisorbed at the same time [911]. Surface oxides
created with oxygen at elevated temperatures (or by aging)
or with liquid oxidants are acidic in character and cause
cation exchange properties. Acidic and basic surface sites
coexist usually, but the concentration of basic sites decreases with increasing acidic character of the surface.
The acidic surface properties are caused by the presence
of carboxyl groups (also in the form of their cyclic

0008-6223 / 02 / $ see front matter 2002 Elsevier Science Ltd. All rights reserved.
PII: S0008-6223( 01 )00165-8

146

H.P. Boehm / Carbon 40 (2002) 145 149

Fig. 1. A few possible surface groups.

anhydrides), lactones or lactols (see Fig. 1), and hydroxyl


groups of phenolic character. These groups differ in their
acidities and can be differentiated by neutralization with
0.05 N solutions of NaHCO 3 , Na 2 CO 3 and NaOH, respectively. The groups have been identified by other chemical
methods used in organic group analysis [9,11,12]. Still
higher base uptake than with NaOH is observed with
alcoholic 0.1 N sodium ethoxide; this can be explained by
the presence of reactive carbonyl groups which form the
sodium salt of a hemiacetal, =C(OEt)(O 2 Na 1 ). In a
simple way, the carbons are agitated with an excess of the
bases, and the excess is determined by back titration after
equilibration (see Refs. [11,12] for experimental details).
The acidity of a given functional group depends on its
chemical environment, i.e., the size and shape of the
polyaromatic layers, the presence and position of other
substituents, and the charge of neighboring dissociated
groups. However, the differences in acidity of the various
types of functional groups seem to be sufficiently large to
allow differentiation by the simple titration method, e.g.,
the difference between NaOH and Na 2 CO 3 consumption
corresponds to the weakly acidic phenolic groups. It has
been shown by careful, continuous titration with alkali that
several peaks appear in the distribution curve of acidity

constants (Fig. 2). They agree quite well with the titration
data [13,14]. Such determinations are very time-consuming
since equilibration in direct titration is very slow [1315].
This may be caused by slow diffusion in narrow pores, if
present, and by slow hydrolytic ring opening of carboxylic
anhydrides and lactones. The method is limited to a range
of pKa (or pH) values between 3.5 and 10.5 because of the
buffering effect of water at very high or very low pH
values. Differences in the number and positions of peaks in
the distribution curve have been observed for differently
pretreated carbons.

2.2. Titration of basic surface sites


While the nature of the acidic surface sites is quite well
understood, the origin of surface basicity is still under
discussion. Continuous titration showed the existence of
three peaks in the pK distribution curves [16]. One reason
for a basic behavior of carbon surfaces may be the p
basicity of the exposed graphene layers [1719]. However,
this basicity is relatively weak. The chemisorption of
oxygen together with acid suggests that the basicity may
be due to oxygen functional groups, and the existence of
pyrone-type structures on the edges of the polyaromatic

Fig. 2. Distribution of acidity constants for an activated carbon oxidized with nitric acid (curve taken from Ref. [13], Fig. 8).

H.P. Boehm / Carbon 40 (2002) 145 149

147

C=O vibrations of quinones or isolated carbonyl groups


[2428]. Peaks at 10001300 cm 21 are ascribed to CO
single bonds [24,25]. Most carbons contain hydrogen, and
band of CH stretch and wagging vibrations are observed.

4. X-ray photoelectron spectroscopy

Fig. 3. Pyrone-type structure.

layers has been suggested [20]. However, it has been


objected that g-pyrone is a much too weak base to account
for the observed strong peak at a pKa of |8.5 of the
conjugate Brnsted acid [16]. Support for the hypothesis of
pyrone-type structures came from theoretical calculations
which showed that the base strength increases strongly
when the carbonyl group and the ring oxygen of a pyronetype structure are distributed on polycyclic aromatic
compounds, e.g., phenanthrene, as shown in Fig. 3 [21,22].
Such structures may have basicities even stronger than that
of pyridine [21].
Most carbons contain more oxygen than can be accounted for by the observed functional groups. The
difference may be due to ether-type oxygen or to carbonyl
groups that do not react with NaOC 2 H 5 .

3. Infrared spectroscopy
Infrared spectroscopy became very useful in analysis of
the functional groups when the limitations due to the high
absorbance of carbon were overcome by the development
of FTIR spectroscopy. It is often used in combination with
diffuse reflectance spectroscopy (DRIFTS) or with total
reflection spectroscopy [2328].
The assignment of the absorption bands is based on
experience with molecular organic compounds. However,
there is often disagreement in their assignments. For
instance, an intensive band near 1600 cm 21 has been
explained by stretching vibrations of aromatic C=C bonds
which are polarized by oxygen atoms bound near one of
the C atoms. An alternative assignment to hydrogenbonded, highly conjugated carbonyl groups was only
recently refuted [24].
The spectra of apparently similarly pretreated samples
show often differences, especially in the C=O bands. Free
carboxylic acids absorb near 1700 cm 21 [25], while cyclic
anhydrides give rise to peaks at 1780 and 1840 cm 21 (sh)
[26]. The anhydrides can be hydrolyzed to the free acids;
the reaction can be reversible on heating. The evidence for
cyclic lactones is not as conclusive. They are described to
give a single peak at 1740 cm 21 [27] or 1760 cm 21 [28],
and a clear distinction from cyclic anhydrides is not
possible. Problems arise also in the assignment of bands to

In X-ray photoelectron spectroscopy (XPS, also ESCA),


core electrons are excited by X-ray irradiation to leave the
atoms. Their binding energy (b.e.) is derived from the
measured kinetic energies. XPS is surface-sensitive since
the escape depth of the photoelectrons amounts to only a
few atomic layers. The b.e. depends on the atomic species
but is also affected by the shielding of the nuclear charge
which is lowered or raised by bonding of the atom to more
electronegative or electropositive atoms, respectively. The
differences in b.e. for various binding states are quite small
compared to the line width, especially with electronegative
elements such as oxygen. A deconvolution of overlapping
peaks is necessary, therefore. However, the results of the
curve fitting are influenced to some extent by the somewhat arbitrary inputs for the number, shape and width of
the peaks. Although modern instruments provide a sufficient resolution also for O1s electrons [15,29], it is more
convenient to measure the C1s signal. Carbon atoms differ
in their b.e. depending on whether they are linked to one O
atom by a single bond (phenols and ethers), a double bond
(carbonyl groups), or two oxygen atoms (carboxyl groups,
lactones). The corresponding signals appear as satellites on
the high-b.e. side of the main C1s peak of the carbons
(Fig. 4). The observed ranges of b.e. are listed e.g., in
Refs. [15,27,29].
XPS requires calibration since charging of the sample
will influence the kinetic energies. Usually, the main C1s
peak of the carbon samples is taken with a b.e. of 284.6 eV
assigned to it. However, values differing by a few tenths of
an eV have been also used (see Ref. [29]). It has been
suggested to list rather the shifts relative to the main
signals instead of the absolute binding energies [29,31].
Alternatively, gold has been sputtered onto the samples
(84.0 eV b.e. for Au 4f 7 / 2 ). It is an advantage of XPS that
the relative surface concentrations of the various species
can be estimated from the peak sizes. The results can be
misleading, however, with porous samples when the
exterior surface is more strongly oxidized by aging.

5. Thermal desorption spectroscopy


While carboxyl groups evolve CO 2 at relatively low
temperatures, other groups are thermally more stable.
Therefore, thermal desorption spectroscopy (TDS), also
called temperature-programmed desorption (TPD), is used
for the study of surface oxides. The samples are heated in a
vacuum or in a helium stream with a constant heating rate

148

H.P. Boehm / Carbon 40 (2002) 145 149

Fig. 4. Typical C1s XPS spectrum of oxidized carbon fibers: (I) phenols, (II) carbonyl groups, (III) carboxyl groups, (IV) plasmon peak
(after Ref. [30], Fig. 3).

(often 10 K / min), and the evolved gases, H 2 O, CO 2 , CO


and H 2 , are determined, mostly by use of a quadrupole
mass spectrometer. The peaks shift to higher temperatures
with increasing heating rate. They are usually very broad,
often with pronounced tailing, and there is considerable
overlapping. A deconvolution of the spectra into separate
peaks has been described [32]. Techniques to achieve
better resolution and to determine the heat of activation for
the decomposition reactions have been described recently
[33].
It is generally assumed that each type of surface group
decomposes to a defined product, e.g., that CO 2 derives
from carboxyl groups and CO from carbonyl and hydroxyl
groups and ether-type oxygen. A good correlation has been
found of the NaOH consumption with the CO 2 -forming
complexes on the surface of an activated carbon [34]. The
results of TDS are not always unambiguous, however. Two
adjacent carboxyl groups may be first dehydroxylated to
the cyclic anhydride which, in turn, decomposes to CO
plus CO 2 . This decomposition occurs at higher temperature than that of free carboxyl groups [32,35]. The
degradation of lactols will probably also produce both
gases. A cyclic lactone can either give one CO 2 molecule
or two CO molecules, with both reactions possibly running
concurrently. Furthermore, secondary reactions cannot be
excluded. Diffusion of the evolved gases is rather slow in
narrow pores, and CO molecules may react to CO 2 with
surface-bound oxygen, or CO 2 molecules hitting the pore
walls may form two CO molecules. Freshly created active
surface sites or free radical sites can facilitate such
reactions.
Carbons oxidized with liquid oxidants usually contain
relatively more carboxyl groups than O 2 -oxidized samples.
Part of these carboxyl groups decompose already at
temperatures significantly below 3008C.
It is not surprising that the TPD spectra published in the
literature often differ in details, since they are influenced
by the pore structure of the carbons and experimental
parameters such as the heating rate. Nevertheless, they can

provide a good overview of the samples surface properties.

6. Electrokinetic measurements
A carbon surface of acidic or basic character is surrounded in aqueous suspension by a diffuse cloud of
dissociated H 1 or OH 2 ions, respectively, and pH values
of ,7 or .7, respectively, are measured in such suspensions. The pH returns to near neutral, however, after
sedimentation of the carbon particles when the carbon
sample is electrolyte-free and purified water is used [17].
In the presence of an electrolyte, ion-exchange results in a
permanent pH change. The surface charge depends on the
pH of the surrounding electrolyte. There is a pH value,
called the point of zero charge (PZC) at which the net
surface charge is zero. The ZPC can be easily determined
by a method called mass titration [36]. It is based on the
fact that the pH of an electrolyte changes in the direction
of the ZPC on contact with a solid powder.
The charged particles move in an applied electric field
(electrophoresis). A thin water layer, containing a part of
the diffuse cloud of dissociated H 1 or OH 2 ions, adheres
to the particles and moves with it. The charge and potential
at its boundary determine the electrokinetic phenomena.
The electrokinetic potential (or z-potential) can be calculated from the measured electrophoretic mobilities. The pH
of zero z-potential is the isoelectric point (IEP). It is not
identical with the ZPC, but usually not very far from it
with non-porous carbon materials. The observed ZPC
range from pH |2 to pH |10.5. With porous carbons,
however, the IEP values are often considerably lower than
the ZPC because the electrokinetic behavior is determined
by the charge on the external surface of the carbon
particles which is usually oxidized by aging [37]. The PZC
is determined, in contrast, by the much larger internal
surface of the pore walls which are oxidized much more
slowly since diffusion of oxygen in narrow pores is very

H.P. Boehm / Carbon 40 (2002) 145 149

slow. Thus, activated carbons with a predominantly basic


surface may show an IEP in the acidic range.

References
y Leon
CA, Radovic LR. In: Thrower PA, editor,
[1] Leon
Chemistry and physics of carbon, vol. 24, New York: Marcel
Dekker, 1994, pp. 213310.
[2] Lewis IC, Singer LS. In: Walker PL, Thrower PA, editors,
Chemistry and physics of carbon, vol. 17, New York: Marcel
Dekker, 1981, pp. 188.
[3] Billinge BHM, Docherty JB, Bevan MJ. Carbon
1984;22(1):839.
F, Rivera-Utrilla, Joly JP, Moreno-Castilla
[4] Carrasco-Marn
C. J Chem Soc, Faraday Trans 1996;92(15):277982.
[5] Verma SK, Walker Jr. PL. Carbon 1992;30(6):83744.

[6] Menendez
JA, Phillips J, Xia B, Radovic LR. Langmuir
1996;12(18):440410.
[7] Bansal RC, Donnet JB, Stoeckli F. Active carbon. New
York: Dekker. 1988:27118, 259333.
[8] Donnet JB, Bansal RC, Wang MJ. In: Carbon black, New
York: Dekker, 1993.
P, editor, Graphite and precursors,
[9] Boehm HP. In: Delhaes
Amsterdam: Gordon and Breach, 2001, pp. 14178.
[10] Voll M, Boehm HP. Carbon 1970;8(6):74152.
[11] Boehm HP. Carbon 1994;32(5):75969.
[12] Boehm HP, Diehl E, Heck W, Sappok R. Angew Chem, Int
Ed Engl 1964;3(17):66978.
[13] Contescu A, Contescu C, Putyera K, Schwarz JA. Carbon
1997;35(1):8394.
[14] Bandosz TJ, Jagiello J, Contescu C, Schwarz JA. Carbon
1993;31(7):1193202.
[15] Biniak S, Szymanski J, Siedlewski J, Sviatkowski A. Carbon
1997;35(12):1799810.
[16] Contescu A, Vass M, Contescu C, Putyera K, Schwarz JA.
Carbon 1998;36(3):24758.
[17] Boehm HP, Voll M. Carbon 1970;8(2):22740.

149

y Leon
CA, Solar JM, Calemma V, Radovic LR.
[18] Leon
Carbon 1992;30(5):797811.
MA, Menendez

[19] Montes-Moran
JA, Fuente E, Suarez
D. J
Phys Chem B 1998;102(29):5595601.
[20] Voll M, Boehm HP. Carbon 1971;9(4):4818.

MA.
[21] Suarez
D, Menendez
JA, Fuente E, Montes-Moran
Langmuir 1999;15(11):3897904.

MA.
[22] Suarez
D, Menendez
JA, Fuente E, Montes-Moran
Angew Chem, Int Ed Engl 2000;39(7):13769.
[23] Sellitti C, Koenig JL, Ishida H. Carbon 1990;28(1):2218.
[24] Fanning PE, Vannice MA. Carbon 1993;31(5):72130.
[25] Zawadzki J. In: Thrower PA, editor, Chemistry and physics
of carbon, vol. 21, New York: Marcel Dekker, 1989, pp.
147380.
[26] Meldrum BJ, Rochester CH. J Chem Soc, Faraday Trans
1990;86(10):18814.

MV, Carrasco-Marn
F.
[27] Moreno-Castilla C, Lopez-Ramon
Carbon 2000;38(14):19952001.
[28] Zhuang Q-L, Kyotani T, Tomita A. Energy Fuels
1994;8(3):7148.
[29] Papirer E, Lacroix R, Donnet JB. Carbon 1994;32(7):1341
58.
[30] Yue ZR, Jiang W, Wang L, Garner SD, Pittman CU. Carbon
1999;37(11):178596.
[31] Nanse G, Papirer E, Fioux P, Moguet F, Tressaud A. Carbon
1997;35(2):17594.
JJM.
[32] Figureido JL, Pereira MFR, Freitas MMA, Orfao
Carbon 1998;37(9):137989.
MA, Carrasco[33] Haydar S, Moreno-Castilla C, Ferro-Garca
F, Rivera-Utrilla J et al. Carbon 2000;38(9):1297
Marn
308.
[34] Otake Y, Jenkins RG. Carbon 1993;31(1):10921.
F, Maldonado-Hodar

[35] Moreno-Castilla C, Carrasco-Marn


FJ,
Rivera-Utrilla J. Carbon 1998;36(1):14551.
[36] Noh JS, Schwarz JA. Carbon 1990;28(5):67582.

y Leon
CA, Radovic
[37] Menendez
JA, Illan-Gomez
MJ, Leon
LR. Carbon 1995;33(11):16559.

You might also like