You are on page 1of 11

Carbon, Vol. 32, No. 5, pp.

759-769, 1994
CoDyright0 1994Elsevier Science Ltd
Printedk &eat Britain. All rights reserved
0008~6223/94
$6.00 t .OO

Pergamon
000%6223(94)E0029-X

REVIEW ARTICLE

SOME ASPECTS OF THE SURFACE CHEMISTRY


CARBON BLACKS AND OTHER CARBONS
Institut fi.ir Anorganische

OF

H. P. BOEHM
Chemie der Universitiit Miinchen, Meiserstrasse 1,
80333 Miinchen, Germany
(Received

4 January

1994)

Abstract-A
review is given on the surface chemistry of carbon blacks and other carbons, in particular,
activated carbons. The main part is devoted to surface oxides with emphasis on the chemical methods
used in the assessment and identification of surface functional groups. Their formation under mild
conditions and the influence of water vapor and metal catalysts on the reaction with air (aging of
carbons) are described. Reaction with free organic radicals can be used for the functionalization of
carbon surfaces.
Key Words-Activated

carbon, carbon black, surface chemistry,

surface functionalization,

surface

oxides.

1. INTRODUCTION

Carbon blacks consist of spheroidal particles with a


pronounced ordering of the carbon layers (graphene
layers). The layers are wrapped around a very disordered nucleus with a preferential orientation parallel
to the particle surface[i,2].
High-resolution TEM showed[3] that the layers
are bent and curved, following the surface. They
are larger, therefore, than the crystallite dimension
L, of 1S-2.5 nm, as determined from line broadening in X-ray diffraction. The spherical primary particles are fused to branched chain-like structures by
deposition of such layers.
Since carbon blacks are produced from hydrocarbons, the dangling bonds at the edges of the carbon
layers are saturated mostly by hydrogen. Often, one
finds large polycyclic aromatic ring systems on the
surface that can be extracted with hot solvents (e.g.,
xylene). One suspects, therefore, that there are also
still larger molecules on the surface that are insoluble, and that there is a gradual transition in size
to the layers that can be recognized in HRTEM
photographs.
Other elements than hydrogen are also found in
carbon blacks. The most important of these is oxygen. Whereas sulfur and nitrogen originate from the
oil precursor, oxygen can also be taken up during
carbon black formation or storage. Much more oxygen is chemisorbed on heating carbon blacks in air
(or oxygen or by treatment with oxidizing media
such as HNO, or NaOCl solution. The surface oxides formed in these reactions have a pronounced
effect on the surface properties of the carbons.
Activated carbons consist of small layers stacks
that are less regularly organized. They are also
759

curved in part, and there is pronounced cross-linking. Due to the activation process, the layer packets
are separated by micropores, most of which seem
to be slit-shaped. Depending on the precursor, there
may also exist meso- and macropores. The color
blacks of high apparent surface area (BET surface
area) are microporous in a similar way.
The surface oxides are bound to the edges of the
carbon layers. It has been shown that basal planes
of graphite are attacked by molecular oxygen only
at their periphery or at defect sites such as vacancies[4-61.
Many oxygen-containing
functional
groups have been detected in the surface oxides of
carbon. Other elements, in particular halogens, can
be chemisorbed on carbon surfaces.
The surface properties ofcarbon blacks and other
types of carbon are influenced to a large extent by
the foreign elements fixed on the surface, in particular by oxygen. This also affects the behavior of
carbon blacks in practical applications.
The present paper gives an overview of the functional groups in surface oxides. The emphasis is on
the methods used for their identification because
some of the pertinent literature is now no longer
readily available. The experiments were performed
in part with carbon black, but activated carbons
have also been used because of their larger surface
areas and larger concentrations
of surface groups.
Activated carbons were prepared from carbonized
sugar char, resulting in very pure carbon materials.
Charcoals produced from wood (Eponit) or peat
(Norit) were also used: they were extracted with
hot hydrochloric acid and washed with hot water
until no chloride could be detected. The carbon
blacks Corax 3 (furnace black) and CK3 (similar to
channel black) were obtained from Degussa. The

760

H.P.

experimental methods are outlined briefly; they are


quite simple, and details are described in the original
literature.
2. SURFACE OXIDES

2.1 General
Carbons, including carbon blacks, can show basic or acidic pH values in aqueous dispersions, A
good correlation between pH and oxygen content
of carbon blacks has been found[7]. The dispersion
is the more acidic, the higher the oxygen content
is. The acidic surface properties are due to the presence of acidic surface groups. Such carbons have
cation exchange properties. Carbons with a low oxygen content show basic surface properties and anion
exchange behavior. The basic properties are ascribed to the presence of basic surface oxides, but
it has been shown that the 7~electron system of the
basal planes of carbon is sufficiently basic to bind
protons from aqueous solutions of acids&-IO].

2.2 Acidic s~r~uce oxides


The acidic surface oxides have been the subject
of many studies that have been summarized in several reviews[l l-181. Figure 1 presents several structures of oxygen functional groups that might be
found at the edges of graphene layers. Carboxyl
groups (a) might give carboxylic anhydrides (b) if
they are close together. In close neighborhood to
hydroxyl groups or carboxyl groups, carbonyl
groups might condense to lactone groups (c) or form
lactols (d). Single hydroxyl groups (e) on the edge
of aromatic layers would be of phenolic character. The existence of carbonyl groups is very plausible; they could come either isolated (f) or arranged
in quinone-like fashion (g). Obviously, other arrangements could be envisaged for quinone-type
functions. Finally, oxygen could simply be substituted for edge carbon atoms (h); such xanthene- or
ether-type oxygen is very di~c~t to detect.
The groups (a) to (e) react more or less weakly
acidic. Evidence for their existence has been found

BOEHM

by classical chemical detection methods, such as


esterification with hot methanol (acid-catalyzed),
formation of acyl chlorides with thionyl chloride or
formation of methyl esters of the carboxyl groups,
and methyl ethers of the phenolic hydroxyl groups
with diazomethane; see below. Methyl esters and
methyl ethers can be differentiated by their different
resistance towards hydrolysis. These reactions were
supplemented by observation of the changes in neutralization adsorption behavior.
Obviously, the individual functional groups, such
as carboxyl groups, will exhibit a spread of their
dissociation constants, depending on the neighboring groups, the size of the graphene layers, etc.
Yet the acidity constants of carboxyl groups, lactones, or phenols differ over several orders of magnitude (see ref. [I8]), and it was established that the
various types ofgroups can be distinguished by their
neutralization behavior. At a given pH of the adjoining aqueous medium, practically all carboxyl groups
will be dissociated to carboxylate with the counter
ions dispersed in the diffuse double layer. It has
been found that the most convenient way of determining the concentration of free carboxyl groups
is to perform a neutralization adsorption experiment
with 0.05 M NaHCO, solution, separate the solution, and titrate the remaining Na+ ions (by adding
excess 0.05 M HCl to an aliquot, boiling off the
COZ, and back-titrate with standard NaOH). It was
shown that practically identical results are found as
with a pH-static titration to pH 8.2[19,20], or by
reaction with a solution of KI and KIO, and titration
of the liberated iodine[l2]. Such solutions establish
a constant pH of 7.5. With nonporous carbons, the
adsorption of diphenylguanidine
agrees, too, with
NaHCO, consumption[8], Direct potentiometric titration suffers from very slow establishment of the
ion exchange equilibria, particularly at higher pH
values[l9-211.
However, the fact that the acidity
constants cluster around discrete values can be
clearly seen in conductometric titration curves[22].
The best results were obtained with dilute NaOCH,
0

0
\\

(a)

(e)

C-O

lb)

if)

fd

Fig. 1. Possible structures of surface oxygen groups (see text).

(h)

c-o

011

761

The surface chemistry of carbon blacks and other carbons

in methanol; an example is shown in Fig.


2. There are distinct breaks in the curves, and as
shown in Table 1, they agree quite well with the
neutralization
values found with NaHCO, and
Na,C!03. In a very recent paper, Bandosz ef af.
described the analysis of titration curves that leads
to a resolution of the various acidity constants[23].
Lactones are weaker acids than free carboxyl
groups. This is clearly demonstrated by the fact
that the lactone ring in phenolphthaleine
is opened
(purple color) by sodium carbonate, but not by bicarbonate. Therefore, 0.05 N Na,C03 solution is
suitable for the determination of carboxyl groups in
lactone-like binding. The formation of lactols (Fig.
Id) from aromatic carboxylic acids with neighboring
carbonyl groups is well known, for example, for 2benzoylbenzoic
acid[24]:

The presence of analogous structures seems


quite plausible on the edge of a graphene layer:

It is generally assumed that phenolic hydroxyl


groups on the carbon surface react with strong
alkali (e.g., NaOH) analogously to free phenols.
Attempts have been made to verify by chemical
reactions the presence of such groups. An example
of such reactions is shown in Table 2. On reaction
with an etheric solution of diazomethane, CH2N2,
carboxylic acids form methyl esters of the acids,
and methyl ethers are produced from phenols.
Lactols form methyl esters, too[251. Ethers are
stable to hydroIysis by dilute acids, in contrast
to esters that are saponified. As shown in Table
2, the neutralization
values with NaOH have decreased to those found with NaZCOj, and the
difference is equal to the methoxyl content remaining after hydrolysis.
Acyf chlorides are formed in the reaction with
thionyl chloride, SOCl,. At reflux temperature,
SOCI, decomposes slowly to Cl? and other products,
and therefore some chlorine is also bound by addition or by hydrogen substitution. Only part of the
fixed chlorine can be hydrolyzed with alkali; the
larger part is resistent even to hot I M NaOH. Only
the chloride that can be recovered by hydrolysis
with alkali was equivalent to the carboxyl groups
(Table 3). This is what one would expect for the
reaction

Table 1. Comparison of the breaks in conductometric titration with


neutralization values (titration with 0.05 M Na+OMe- in MeOH)

Sample
Corax 3a, H.T. 1400C
ox. with (NH,),S,Os
Corax 3 H T. 3000C
ox. v&h &MnO,
Graphite wear dust
ox. with air at 420C

OH

-;:x!:-

Fig. 2. Conductometric titration of oxidized graphite wear


dust with 0.05 M NaOMe in MeOH.

solutions

C-O

1st

NaHC03

break
peq/g

uptake
Feq/g

2nd
break
peq/g

Na?CO,
uptake
CLeqg

65

56

85

92

10s

99

145

134

205

181

350

363

Corax 3 = furnace black (Degussa).

762

H. P.BOEHM

;:k+ COCI

2NaOt-l

irCOO-No+

Na+CI-

However, with NaOEt in some cases the consumption was less than with the carbon before chlorination (see Table 3), although one would expect an
unchanged consumption, as was observed with the
other samples:

COCI
+ NaOEt

i_x

determined with NaHCO,. It is known that phthalic


acid is converted to phthalic anhydride in the reaction with thionyl chloride. The acyl chloride must
have been formed in this case from the lactone-type
carbonyl functions.
Another attempt to identify carboxyl groups used
the Friedel-Crafts reaction with dimethylaniline and
the Schmidt rearrangement,
a special type of the
Curtius rea~angementr26~.
In both reactions, the
carboxyl groups were first converted to acyl chloride by reaction with thionyl chloride. After treatment with dimethylaniline and anhydrous aluminum
chloride in hot nitrobenzene as a solvent, and extraction with dilute hydrochloric acid, water, and
ethanoi, the nitrogen content corresponded to the
acyl chloride content after ~hlorination[Z6~. The
neutralization value with NaHCO, was decreased
by an equivalent amount, whereas the neutralization
of the other groups was unchanged within experimental limits. The results presented in Table 4 can
be explained by the reaction sequence

COOEt
+

Ph-NMe,

Na+CI-

R-COOH + R-COCl-

A plausible explanation is that anhydrides of two


neighboring carboxyl groups had formed, which
form ester plus carboxylate:
0
Z,
~

c
II
0

,O

+ NaOEt

R-CO-C,H,-NMe,.
The Schmidt rea~angement involves reaction of the
acyl chloride with an aicoholic solution of sodium
azide, NaN,. The resulting acyl azide decomposes
to form a urethane, which is stable in acidic media,
but decomposes with alkali to give an amine. Thus,
again, the carboxyl groups are destroyed:
NaNl

R-COOH + R-COCl----+
COO-No+
cx

R-CON,2
+H,O

R-NH-COOEt

-CO,,-EtOH R-NH2'

COOEt

The loss in NaOEt consumption corresponded


quite well to half the content of carboxyl groups

The reaction was followed by determinations of


the nitrogen and ethoxy contents, as well as the
changes in neutralization behavior after each step.

Table 2. Influence of reaction with di~omethane on the neutralization behavior of oxidized carbons
(the activated carbons from carbonized sugar char were oxidized with 0, at 400C)
Neutralization
Treatment

NaHCO,

Sugar char, H.T. 950C


after oxidation
methylated with CHzNz
methylated and hydrolyzed with
hot HCI solution

200
210

Na$O,
430
410
diff.:

Sugar char, H.T. 1100C


after oxidation
methylated with CH,N,
crethylated and hydrolyzed with
hot HCl solution

160
170

0CH3

in peqig

320
320
diff.:

NaOH

content
pmol/g

720
410

0
720
290

310
690
330
360

0
720
380

163

The surface chemistry of carbon blacks and other carbons


Table 3. Neutralization

behavior of oxidized carbons after reaction with thionyl chloride (oxidation with OZ at 400C)
Change in consumptn.
after SOCI? reaction

Sample
Sugar char
H.T. 1100C
Sugar char
H.T. 1100C
Eponit
H.T. 1100C
Activ. sugar char
H.T. 950C
Activ. sugar char
H.T. 950C
Eponit,
add. activation

Group
I
peq/g

Group
II
peqig

Hydrolyzable
Cl
peqlg

160

150

160

+I80

230

210

210

+200

870

430

+410

600

460

480

+450

-340

600

350

350

+350

-310

660

680

520

+580

-340

NaOH
peq/g

NaOEt
peq/g

Group I = NaHCO, neutralization.


Group II = NarCO,-NaHCO, neutralization

The results are a bit complicated because lactones


may form ethyl esters,and carbonyl groups may react with NaN, in hot ethanol. When the reaction
was performed at room temperature, one nitrogen
atom was taken up, and one carboxyl function was
lost for each acyl chloride group formed originally[26,27].
Phenols are very weak acids, neutralized by
NaOH solutions. In contrast to carboxylic acids,
they can be reacted with 2,4-dinitrofluorobenzene
(DNFB) or with p-nitrobenzoyl
chloride
(pNBC)[26]. As shown in Table 5, the bound quantities agree very well with the difference between
NaOH and Na&Os consumption.
Using a still stronger base than NaOH, sodium
ethoxide, Na+OEtt, in ethanol, additional groups
are detected. We have shown[26] that equivalent
quantities of Nat and OEt- are bound by the carbons in this reaction. The assumption that this is
due to the formation of sodium salts of hemiacetals
from carbonyl groups

is based on identical conversions after reaction with


hydroxylamine that lead to oximes (Table 6).

The carbonyl groups are susceptible to reduction


with nascent hydrogenl261. This results in a decrease of NaOEt consumption. After reaction with
zinc and hydrochloric acid, the sodium ethoxide
uptake was reduced to that of sodium hydroxide in
some cases. In one such case, the weakly acidic
carboxyl function (neutralized by Na*CO,) had increased in acidity and reacted with NaHCO,, too.
This observation points to an interrelationship
between the carbonyl groups and neighboring carboxyl
groups. The quantity of active carbonyl groups was
also reduced after reaction with NaBH, or LiAlH,.
Pairs of carbonyl groups arranged at the periphery of the graphene layers in such a way that a
system of conjugated double bonds can be formally
drawn in, will behave similar to quinone functions. Reduction peaks in cyclovoltammetry experiments have been ascribed to such quinone functions[28-311. Also, some chemical evidence for
their existence has been presented[32,33].
In our early experiments we observed that frequently equivalent quantities of the various groups
are found after strong oxidation that results in partial
gasification of the carbons. This observation led us
to speculate on a model in which carboxyl groups,
lactols, and phenols are grouped in immediate proximityl341. It was thought that this surface complex
is an intermediate step in the removal of carbon
atoms from the layer edge. It must be stressed,

Table 4. Results of the Friedel-Crafts reaction with dimethylaniline


(from refs. 126,271)
NaHCO?
neutralization
Sample
Eponit ox. with OZ
Eponit
ox. with (NH&&Ox

before
peqig

after
yeqlg

Cll after
reaction w.
SOClz
peq/g

460
590

280
280

230
310

N content
after
reaction w.
Ph-NMe*
pmol/g
220
330

764

H.P.
Table 5. Reactions

BOEHM

of phenolic surface groups on oxidized carbons


(oxidation with O2 at 400C)

Sample
Carbon black CK3
Activ. sugar char
H.T. 950C
Sugar char, H.T. 1100C
Activ. sugar char
H.T. 1100C

Difference
NaOHNa2C03
uptake
weoig

Hydrol.
resistent
OCH,
gr.
fimol/g

Conversion

with

DNFB
wmol/g

p-NBC
~molig

320
610

330
620

130
620

290
-

420
390

430
-

420
370

400

however, that very frequently other distributions


of acidic surface functions are observed, especially
after relatively mild oxidation treatments. In the
more recent literature, such equivalences of different surface groups are rather rare. With carbon
blacks, there are often fewer carboxyl groups than
reactive carbonyl groups. In the early stages of oxidation with solutions of (NHJ2S205, KMnO,, etc.,
at room temperature, only weakly acidic carboxyl
groups and reactive carbonyl groups are formed
(lactol functions)[271. Oxidation with concentrated
nitric acid often produces acidic compounds of relatively small size that go into homogenous solution
on treatment with dilute alkali[20].
Usually the carbons contain more oxygen than
can be explained
by the detected functional
groups[12,35,36]. This oxygen is usually ascribed to
ether-type oxygen (Fig. lh) without further proof.
Determination
of active hydrogen by reaction
with methyl magnesium iodide or lithium methyl
was found to be low compared to the content of
carboxylic and phenolic groups on carbon black
Spheron 6, whereas good agreement was found with
agraphite[37]. The reason is that part of the methane
may be adsorbed on the carbon black surface.
Therefore, with higher-surface-area carbons, active
hydrogen can only be determined by isotope exchange. Table 7 shows that the contents of active
hydrogen agreed quite well with the NaOH uptake
with carbon black, as well as with a graphite wear
dust produced by milling graphite under argon[22].

2.3 Basic surface oxides


Basic surface oxides are always present on a
carbon surface. When a carbon is heated to ca.
1000C in vacua or under an inert gas, the existing
surface compounds are almost quantitatively
decomposed. When this carbon is exposed to dry oxygen after cooling to room temperature, some oxygen
is chemisorbed. After submersing this carbon under
aqueous acids, the same quantity of oxygen again
is taken up, and approximately one equivalent of
acid per chemisorbed oxygen atom is bound at the
same time (Fig. 3)[39]. The bound anion of the acid
can be exchanged for other anions. Water is a sufficantly strong acid, and OH- ions are bound when
the reaction is conducted in pure water, giving rise
to an alkaline pH of the dispersion. Some H202 is
formed during the second chemisorption reaction,
but carbon surfaces catalyse its decomposition and
it decomposes rapidly[39,40]. The chemisorbed HCl
or H,O are desorbed on outgassing in vucuo, even
at room temperature[41].
Garten and Weiss, who studied the basic surface
oxides in the 195Os[42,43], ascribed the basic properties to chromene-like structures. Voll and Boehm
concluded on the basis of a few chemical reactions
that y-pyrone-like structures, as shown in Figs. 4
and 5, are more plausible[44]. The ether-type oxygen can easily be replaced by nitrogen in the reaction
with ammonia. The hydroxyl groups can be methylated with diazomethane,
whereas the anion exchange property is preserved.

Table 6. Reactions of carbonyl groups on oxidized carbons (oxidation


with 0, at 400C if not stated otherwise)

Samnle
Sugar char, H.T. 95OC,
CO*-activated
Sugar char, H.T. 1100C
Eponit, H.T. 1100C
(ox. with (NH&S20s)
Carbon black CK3

Bound
-0Et
groups
pmol/g

Fixed N
after react.
with
NH20H
peq/g

670

680

660

480
630

460
640

430
580

210

180

Difference
NaOEt-NaOH
uptake
ueolg

765

The surface chemistry of carbon blacks and other carbons


Table 7. Active hydrogen on oxidized carbon surfaces. In the Zerewitinov method, the volume of
methane that forms in the reaction with a solution of CH,MgI is measured. The isotope exchange was
performed by exchange with deuterium from D20, as described in detail in ref. [38].

Sample
Graphite wear dust
ox. with air at 420C
ox. with NaOCl
ox. with (NH4)2S?08
Corax 3
ox. with (NH&SI08
Corax 3, H.T. 1100C
ox. with (NH&O8
Corax 3, H.T. 1400C
ox. with (NH&O8

Surface
area
m?!g

Method of
determination

Active H
fimolig

NaOH
uptake
i-=q&

275
345
330
n.d.

isotope exch.
isotope exch.
isotope exch.
Zerewitinov

670
1040
1440
130

580
1025
1650
200

n.d.

Zerew~tjnov

228

206

n.d.

Zerewitinov

94

123

As mentioned earlier, hydrochloric acid is also


adsorbed on the basal planes of carbons and graphite[8-IO]. This is due to the basic character of aromatic 7r electrons. Also in Fig. 3, the HCI adsorption
isotherm is a little higher than the oxygen-uptake
curve. By potentiomet~c titration two types of proton-binding sites were found on a carbon fiIm1451.
One corresponded to a base with a mean basicity
constant pKb = 6.6, while the second site was a
very weak base (pKb > 11)[451. Table 8 shows that
some hydrochloric acid is adsorbed even on graphite
powders with a very small surface area of the prism
faces. The acid uptake did not correlate well with
the BET surface areas, however. Papirer et nl.[36]
observed, in contrast, that the concentration of basic groups on oxidized carbon blacks was proportional to the surface area. They concluded further
that one of the oxygen atoms of a pyrone group is
quite heat-resistant, and is desorbed as CO and CO1
only between 800C and 950C. It was confirmed
that one oxygen atom is fixed at 100C for each basic
site formed.

Fig. 3. 0, and HCI consumption on immersion of outgassed aciivated carbon from sugar char in dilute hydrochloric acid. (Outgassing at 95OC, HCI adsorption by pHmetric titration to constant pH).

Oxidized carbons always contain basic sites in


addition to the acidic functions. However, their
quantity is usually less than with the original carbon.
It was found that the acid take-up increased when
acidic groups on Pd- or Pt-loaded carbons were reduced in hydrogen at 80-500C (see Fig. 611461.Hydrogen is chemisorbed as H atoms on Pt or Pd, and
can diffuse from the metal surface to the surface of
the support (hydrogen spillover)[47]. This provides
a relatively mild way of reducing surface oxides, as
compared to hydrogen treatment of the metal-free
carbons,
which becomes effective only above
4OO*C.The carboxyl groups/lactols were more susceptible to reduction than the phenofic groups[48].
Obviously, basic surface groups form at the same
edge sites as acidic groups, but their quantity remains lower than that of the destroyed acidic sites.
2.4 Sp~ctroscopi& rn~~~o~s
There have been numerous attempts to study the
surface groups by spectrometric
methods, especially by infrared spectroscopy. This is made very
difficult by the strong IR absorption of carbon. The
first confirmation of the presence of carboxyl groups
was found with dispersions of highly oxidized fineparticle-size
color blacks[49-5 I I. Progress was
made by application of the ATR technique[52] and
especially by FT spectrometers[53,54].
Adsorption
peaks at ca. 1000 cm- and ca. 1200 cm- indicate
the presence of C-O single bonds. The spectrum
of HOOK-oxidized graphitized
Thornel carbon
fibres showed peaks at 1705, 1730, and 1750 cm-,

(1)
Fig. 4. Proton addition to y-pyrone-like

(n)
structures.

746

H. P. BOEHM

Table 8. Adsorption of HCI from 0.05 M solution on graphites and carbon blacks
Surface area
m*/g

HCI uptake

13
7.5
14
42

2s r 2
34

Graphites
Kropfmiihl AF (natural)
Lonza KS 75 (synthetic)
Lonza KS 15
Lonza KS 75/KM
Carbon blacks
CK3
Corax 3
Corax 3, H.T. 9OOC
Corax 3, H.T. 3000C

Fig. 5. Possible structures of basic surface sites on a


graphene layer, derived from the y-pyrone structure.

EiLeo/g

::

77
84
87
62

40
36
64
27

which were assigned to free carboxyl groups, esters,


and lactone groups, respectively[54]. An absorption
at 1640 cm- was explained by carbonyl groups situated near hydroxyl groups (enols). However, we
observed the same frequency with violanthrone (dibenzanthrone),
a polycyclic system with nine condensed rings and two carbonyl groups. An extended
review on this subject has been published[45], but
unfortunately there are not many studies on carbon
black surfaces.
X-ray photoelectron spectroscopy (XPS, ESCA)
has also been used for oxidized carbons, mainly
carbon fibres. Due to the high electronegativity
of
oxygen, the 0 1s signal is not very sensitive to the

way in which the oxygen is bound. One fmds essentially two different binding energies for oxygen singly and doubly bonded to carbon[SS]. More information can be gained from satellites of the C Is peak
at higher binding energies[55,56]. The shifts from
the main peak range from 1.6 eV for carbon atoms
bonded to OH groups to 4.0 eV (carboxyl groups
or esters) and even higher. XPS is not very accurate
for quantitative determinations, and it is used mainly
in studies of low-surface area carbons such as carbon fibers.
Thermodesorption
spectroscopy is not very well
suited for the determination
of the functional
groups. Carboxylic groups are least stable, and they
decompose with formation of C02. However, this is
not possible when carboxylic anhydrides are formed
first, and they will produce CO2 plus CO in this

Acidic

Basic

450

surface

groups

1 psq/gf

groups

[peq/gf

T-_q,i

350
:

250
,50

surface

)
0

ox

ox-to,,,

100

200

r;l
t

\
350

O--U Norit,

Nod:'

300

400

1~

o.,.o Norlt, ox

1
250

500
Temperature

TNorl:'

100

ox-:,,t

200

300

400

[C ]

Fig. 6. Reduction of acidic surface sites on treatment with hydrogen (left side) and concurrent increase
in basic surface sites (right side). Open symbols: Activated carbon Norit, oxidized with 0,: Filled
symbols: Norit loaded with 200 pmol/g of platinum (80% dispersion at 150C; there was some smtering
at higher temperatures).

, D

500

The surface chemistry of carbon blacks and other carbons

case. The TDS patterns for CO* and CO are affected


by the type of carbon used, the oxidation conditions,
and the way the TDS experiment is carried out[35].
The probability
of secondary
reactions of the
evolved gases is drastically enhanced with porous
carbons. Especially in micropores, CO* might react
with the carbon surface to give CO, and at lower
temperatures CO might react with surface oxygen
complexes to give CO,[57].
2.5 Oxidation under mild conditions
Formation of surface oxides is an activated process. It has been shown that no oxygen is chemisorbed at low temperatures (below 200 K) on freshly
produced carbon surfaces[58]. Surface oxidation
with molecular oxygen is fairly rapid above 3OoC,
but obviously surface oxides must be formed more
slowly at lower temperatures. This phenomenon of
aging of carbon materials (activated carbons and
carbon blacks) is well known in the relevant industries. The isoelectric point of a carbon black heattreated at 1500C shifted from pH > 10.5 to pH
5.8 after one years storage in ambient air[59]. The
presence of water vapor significantly accelerates the
surface oxidation[60-621.
The surface becomes
more hydrophilic when covered with surface oxides,
and the adsorption properties of activated carbons
are greatly affected[60,62].
We have treated an activated carbon (Norit) and
a furnace black (Corax 3) either in a drying oven at
110C in ambient air or in air of 70% relative humidity at 60C. The increase in acid surface functions
was determined by titration with 0.05 N NaOH. As
is shown in Fig. 7, acidic surface groups are formed
in both cases with creation ofrelatively many phenolit groups[48]. The rate of their formation decreases
gradually, but the surface was not saturated even
after 70 days. With humid air the oxidation was
much faster with the activated carbon, although the
reaction temperature was lower. With the carbon
black, slightly fewer acidic groups were found
after reaction in moist air at 60C than in air of
low relative humidity at 110C. Very likely this
difference arises because much more water is

NaOH uptnke

Norit

NaOH uplnke

@mWgl
&lo
I

..

.-

comx

@mWgl
.
,I

-time+
Fig. 7. Formation of acidic surface groups during lowtemperature oxidation of an activated carbon (Norit) and
a furnace black (Corax 3); open symbols. For comparison,
the same carbons loaded with 200 pmol/g of palladium
were also studied (filled symbols).

767

adsorbed in the micropores of the activated carbon


than on the more hydrophobic
surface of the
nonporous carbon black.
For comparison, we also studied the same carbons after loading with 200 pmol/g of palladium by
incipient-wetness
impregnation with H2PdCl,. Although it is known that palladium catalyses the gasification of carbons with oxygen above 350C[63],
the catalytic effect observed at 60-110C (Fig. 7)
was impressive. It was very significant with the carbon black.
Liquid oxidizing agents are often used for the
production of acidic surface oxides. Concentrated
nitric acid is usually used at its reflux temperature,
and-as
shown above-solutions
of (NHJ2S208,
NaOCl, or KMnO, can be used at room temperature. Puri also found surface oxidation with KBrO,
or even KNO, solutions[64]. Therefore, it is not
surprising that carbon surfaces can also be oxidized
with metal cations of a sufficiently high oxidation
potential. It has been observed that hydrated Ag+
ions are reduced to metallic silver[65-671. Analogously, PdCI:- ions are reduced to Pd0[48], and
iron(III) ions are reduced to iron(I1). Using a series
of different metal ions, Fu et al. showed experimentally that the extent of oxidation depends on the
redox potential of the system, which is dependent
not only on the element, but also on temperature,
pH, and the presence of complexing ions[67]. After
oxidation with a weaker oxidizing agent, additional
surface oxides can be produced by action of
stronger oxidants[67].
2.6 Influence of surface oxides on the
surface properties
Surface oxides provide hydrophilic sites on a
hydrophobic surface. A high concentration of chemisorbed oxygen makes a carbon black hydrophilic,
and it disperses very well in water. The dissociation
of carboxylic surface groups facilitates the dispersion by creating a negative surface charge[l4]. Carbon blacks oxidized with ozone form spontaneously
colloidal dispersions on immersion in water[68]. The
polar and the hydrophobic parts of a carbon black
surface can be separately determined by microcalorimetric determination
of the heat of reaction for
the preferential adsorption of n-butanol (for polar
sites) and n-dotriacontane,
C,*H, (for the unpolar
surface),
on the carbons
immersed
in nheptane[69].
The carboxylic groups on oxidized carbon surfaces produce cation exchange properties, and carbons always exhibit an anion exchange capacity because basic surface oxides are always present when
carbons have been exposed to the atmosphere.
However, the concentration of basic surface sites
is relatively small, whereas considerable cation exchange capacities can occur. This can have consequences for the application of color blacks in printing inks, paints, etc.
The surface charge of carbons and the zeta poten-

768

H. P. BOEHM

tials in aqueous dispersions are determined by the


nature of the surface groups and the pH. Temperature and the concentration
of non-potential-determining ions have a smaller effect (provided there
is no specific adsorption). The isoelectric point of
a color black, Printex U, was found at pH 4.5 2
0.2. After heat treatment under N, at llOOC, that
is, thermal destruction of most surface oxides, it
shifted to pH 8.8; after treatment at 1500C it was
at pH > 10.5, and even higher after graphitization
at 27OOC[59]. On oxidation with air at 37OC, new
acidic functions were formed with a consequent shift
of the IEP to pH 5.7 after 2 h and pH < 3.0 after
60 h. These measurements were performed by electrophoresis, but mass titration provides a convenient way to determine IEPs if sufficient quantities
of material are available[70].
3. BINDING OF NITROGEN AND HALOGENS
Carbon blacks contain small concentrations
of
nitrogen. This is very likely bound substitutionally
at the edges of the graphene layers. Nothing is
known about the solubility of nitrogen in the interior
of graphene layers (e.g., in graphite). It is well
known that diamonds frequently contain substitutionally dissolved nitrogen.
On treatment of carbons with ammonia at elevated temperatures (e.g., 600-9OOC), nitrogen is
bound[71-731. This reaction has been studied with
carbon blacks as well as with activated carbons.
Photoelectron spectra show two main nitrogen N
1s signals. The peak with a binding energy of ca.
400 eV is assigned to amine-type nitrogen, and the
signal at 398-399 eV is generally explained by pyridine- or acridine-type nitrogen[71,73]. These studies
have been performed mainly because the catalytic
activity of carbons (e.g., in oxidation reactions) is
drastically increased by incorporation of nitrogen.
At temperatures above 600C some carbon is gasified in the reaction with NH,, and microporosity
develops. But also at 6OoC, there is an increase in
HCl neutralization, indicating that new basic surface
sites, presumably amine groups, were formedl741.
Halogens react with carbon blacks by substitution of hydrogen,
and hydrogen halide is
evolved. Donnet reported that all hydrogen was removed from the surface of carbon black Vulcan
6 after treatment with chlorine at 75OC[15]. The
maximum chlorine uptake was observed at 4505OOC[13,75,76]. The bound chlorine is hydrolyzed
only to a small part by hot 1 M sodium hydroxide[76], but it can be removed by heating in hydrogen[77,781.
A study of chemisorption of chlorine at 500C
has been performed using a carbon black and carbon
cloth that had been outgassed at 1000C to remove
all chemisorbed oxygen and most of the hydrogen[77]. It was concluded that chlorine reacts stepwise; first it is chemisorbed by addition, possibly
to double bonds. Subsequently, in the major step,

it is bound substitutionally with HCl evolution, and


finally it removes hydrogen as HCl without Cl
chemisorption.
Bromine can be chemisorbed in a similar way,
but the bound quantities are lower.

4. REACTIONS

WITH FREE RADICALS

The surfaces of carbons react with free radicals,


and this reaction can be used for the grafting of
functional molecules or polymers to the carbon surface. The fixation of 2-isobutyronitrile
radicals,
(CH&-CN,
on the carbon surface can be easily
determined from their nitrogen content[79]. The radicals are produced by heating a solution of azo-diisobutyronitrile.
Similarly, the carbons react with
3,5-dichlorobenzoyl
peroxide or lauryl peroxide.
The radicals seem to attack especially quinone-type
functions, and radicalic sites may be produced on
the carbon surface[80]. Benzoate groups are found
on the surface after reaction with dibenzoyl peroxide, and these can be hydrolyzed to surface phenolic
groups[81].
Polystyrene can be grafted to carbon black surfaces by starting radicalic polymerization in styrene
solutions in the presence of carbon blacks. It has
been found that carbon blacks with quinone oxygen
inhibit polymerization initially until all quinone is
converted to hydroquinone[8 1,821. Prior hydrogenation of the quinone groups eliminates the inhibition period. These and other grafting reactions are
well described in Donnets review[l5].

REFERENCES

H. P. Boehm, Z. Anorg. A&. Chem. 297,315 (1958).


W. M. Hess, and L. L. Ban, J.
Appl. Crystallogr. 1,1 (1968).
P. A. Marsh, A. Voet, T. J. Mullens, and L. D. Price,
R. D. Heidenreich,

Rubber

Chem.

Technol. 43. 470 (1970).

J. M. Thomas, In Chemist& and~Phy&s

of Carbon,

Vol. 1(edited by P. L. Walker, Jr.), p. 121-202. Marcel


Dekker, New York (1965).
5. G. R. Hennig, In Chemistry and Physics of Carbon,
Vol. 2 (Edited by P. L. Walker, Jr.), pp. l-49. Marcel
Dekker, New York (1966).
6. R. T. Yang, In Chemistry and Physics of Carbons,
Vol. 19 (Edited by P. A. Thrower), pp. 163-210. Marcel Dekker, New York (1984).
7. M. L. Studebaker. Rubber Chem. Technol. 30. 1400
(1957).
8. D. Rivin, Proc. 5th Conference on Carbon, University
Park, PA (1963), Vol. 2, p. 199. Pergamon Press, New
York.
9. H. P. Boehm, Furbe und Lack 79, 419 (1973).
10. C. A. Leon v Leon. J. M. Solar. V. Calemma and
L. R. Radov& Carbon 30, 797 (1492).
11. H. P. Boehm, E. Diehl, and W. Heck, Reu. G&n.
Caout. 41, 461 (1964).
12. H. P. Boehm, E. Diehl, W. Heck, and R. Sappok,
Angew.

Chem.,

Internal.

Ed. Engl. 3, 669 (1964).

13. H. P. Boehm, In Advances in Catalysis (Edited by


D. D. Eley, H. Pines, and P. B. Weisz), Vol. 16, pp.
179-274. Academic Press, New York. London (1966).
14. J. B. Donnet, Carbon 6, 161 (1968).

769

The surface chemistry of carbon blacks and other carbons


IS. J. B. Donnet, Tanso 88, 12 (1977).
16. B. R. Puri, In Chemistry andPhysics ofCarbon, Vol. 6
(Edited by P. L. Walker, Jr.), pp. 191-282. M. Dekker,
New York (1970).
Pressures 22.
17. H. P. Boehm. Hiah Temperatures-Hkh
275 (1990).

18. C. A. Leon y Leon and L. R. Radovic, In Chemistry


and Physics of Carbon, Vol. 24 (Edited by P. A.
Thrower), pp. 213-310. Marcel Dekker, New York
(1992).
19. Y. Matsumura, S. Hagiwara, and H. Takahashi, Carbon 14, 163 (1976).
20. S. Neffe. Carbon 25, 441 (1987).
21. A. S. Arico, V. Antonucci, M. Minutoli, and N. Giordano, Curhon 21, 337 (1989).
22. G. Palkowitsch, Doctoral thesis, Univ. of Heidelberg
(1969).
23. T. J. Bandosz, J. Jagiello, C. Contescu, and J. A.
Schwarz, Carbon 31,-l 193 (1993).
24. M. S. Newman and C. W. Muth. /. Amer. Chem. Sot.
73, 4627 (1951).
25. G. M. Badger, J. E. Campbell, J. W. Cook, R. A.
Rapfael, and A. 1. Scott, /. C/tern. Sot,. 1950, 2326.
26. H. P. Boehm, E. Diehl, and W. Heck, Proc. 2nd Int.
London Carbon & Graphite Conf, Sot. Chem. Ind.,
London (1966), p. 369.
27. W. Heck, Doctoral thesis, Univ. of Heidelberg (1966).
28. H. V. Drushel and J. V. Hallum. J. Phvs.
, Chem. 62.
1502 (1958).
29. B. D. Epstein, E. Dalle-Molle, and J. S. Mattson,
Carbon 9, 609 (1971).

30. K. Kinoshita and J. A. S. Bett, Carbon 11,403 (1973).


31. J. P. Randin and E. Yeager, J. Electroanal. Chem.
58, 313 (1975).
32 J. B. Donnet, J. Lahaye, and J. Schultz, Buh. Sot.
Chim. France

33 Y. Matsumura

1966, 1769.

and H. Takahashi,

47. D. J. Suh, T.-J. Park, and S.-K. Ihm, Carbon 31,427


(1993).
48. T. Kuretzky, Doctoral thesis, Univ. of Munich (1993).
49. J. B. Donnet. F. Hueber. N. Perol. and J. Jaeger, J.
Chimie Phys. 60, 426 (1963).
50. J. B. Donnet, Carbon 6, 161 (1968).

51. E. Papirer, E. Guyon, and N. Perol, Carbon 16, 133


(1978).

52. J. S. Mattson and H. B. Mark, J. Co//. Interf. Sci. 31,


131 (1969).
53. J. M. OReilly and R. A. Mosher, Carbon 21,47 ( 1983).
54. C. Sellitti, J. L. Koenig, and H. Ishida, Carbon 28,
221 (1990).

55. C. Kozlowski and P. M. A. Sherwood, J. Chem. Sot.,


Faraday

Trans. 180, 2099 (1984); 81, 2745 (1985).

56. J. B. Donnet and G. Guiloain. Carbon 27.749 (1989).


57. P. J. Hall, J. M. Calo, and W. D. Lilly, Proc. Carbon
88, Internat. Carbon Conf., Newcastle-upon-Tyne
(1988), p. 77.
58. G. H. Fedorov, Yu. A. Zarifyants, and V. F. Kiselev,
Zh. Fiz. Khim. 37, 2344 (1963).

59. A. C. Lau, D. N. Furlong, T. W. Healey, and F.


Grieser. Coil. Surf. 18, 93 (1986).
60. B. H. M. Billinge, J. B. Docherty, and M. J. Bevan,
Carbon 22, 83 (1984).

61. V. R. Deitz, Carbon 25, 31 (1987).


62. L. B. Adams, C. R. Hall, R. J. Holmes, and R. A.
Newton, Carbon 26, 451 (1988).
63. F. Carasco-Marina, J. M. Solar, and L. R. Radovic,
Ext. Abstr.,

Carbone

90. Internat.

Carbon

Conf.,

Paris (1990). p. 672.


64. B. R. Puri, Proc. 5th Conf on Carbon, University
Park, PA (1961), Vol. 1, p. 165. Pergamon Press, New
York, Oxford (1962).
65. B. R. Puri, S. Singh, and 0. P. Mahadjan, lndien .I.
Chem. 3, 54 (1965).

Carbon

17, 109

(1979).

34 H. P. Boehm, Angew. Chem., Internal. Ed. Engl. 5,


533 (1966).

35 H. P. Boehm and G. Bewer, Proc. 4th Internat. London Carbon and Graphite Conf., 1974. Sot. Chem.
Ind., London (1976), p. 344.
36. E. Papirer, J. Dentzer, S. Li, and J. B. Donnet, Carbon
29, 69 (1991).
37 S. S. Barton and B. H. Harrison, Curbon 13, 283
(1975).

38. H. P. Boehm and H. Knozinger, In Catalysis-Science


und Technology Vol. 4 (Edited by J. R. Anderson
and M. Boudart), pp. 39-207. Springer-Verlag, Berlin,
Heidelberg, New York (1983).
39. M. Voll and H. P. Boehm, Carbon 8, 741 (1970).
40. E. C. Larsen and J. H. Walton. J. Phvs. Chem. 44.
70 (1940).
41. H. P. Boehm and M. Voll, Carbon 8, 227 (1970).
42. V. A. Garten and D. E. Weiss, Austral. .I. Chem. 10,
309 (1957).
43. V. A. Garten and D. E. Weiss, Rev. Pure Appl. Chem.
(Australia) 7, 69 (1957).
44. M. Voll and H. P. Boehm, Carbon 9, 481 (1971).
45. J. Zawadzki. In Chemistry and Physics of Carbon,
Vol. 21 (Edited by P. A. Thrower), pp. 147-380. Marcel Dekker, New York (1989).
46. T. Kuretzky and H. P. Boehm, Proc. Carbon 92,
Internut. Curbon Conf., Essen, Germany (1992), p.
260.
,

66. P. G. Hall, P. M. Gittins, J. M. Winn, and J. Robertson, Carbon 23, 353 (1985).
67. R. Fu, H. Zeng, and Y. Lu. Carbon 31, 1089 (1993).
68. J. B. Donnet and E. Papirer, Bull. Sot. Chim. France
1965, 1912.

69. A. J. Groszek, Carbon 25, 717 (1987).


70. J. S. Noh and J. A. Schwarz, Carbon 28,675 (1990).
71. R. Kurth, B. Tereczki, and H. P. Boehm, Ext. Abstr.
15th Bienn. Conf. on Carbon, Philadelphia, PA (1981),
p. 244.
72. H. P. Boehm, G. Mair, T. Stohr, A. R. de Rincon,
and B. Tereczki, Fuel 63, 1061 (1984).
73. B. Stohr, H. P. Boehm, and R. Schlogl, Carbon 29,
707 (1991).

74. T. Stohrand H. P. Boehm, Proc. Carbon 86, Internut.


Carbon Conf., Baden-Baden (1986), p. 354.
75. H. P. Boehm, A. Clauss, and U. Hofmann, Proc. Srd
Bienn. Conf. on Carbon, Buffalo, NY (1957), p, 241.
Pergamon Press, New York, London (1959).
76. B. R. Puri and R. C. Bansal, Carbon 5, 189 (1967).
77. H. Tobias and A. Soffer, Carbon 23, 281 (1985).
78. M. Gurrath and H. P. Boehm, Ext. Abstr., 21st Bienn.
Co& on Carbon, Buffalo, NY (1993), p. 462.
79. J. B. Donnet and G. Henrich, Bull. Sot. Chim. France
1960, 1609.

80. K. Ohkita, N. Tsubokowa, E. Saitoh, M. Noda, and


N. Takashima, Carbon 13, 443 (1975).
81. K. Ohkita and N. Tsubokowa, Carbon 10,631 (1972).
82. J. B. Donnet, L. Geldreich, G. Henrich. and G. Riess,
Reu. GPn. Cuout. 41, 519 (1964).

You might also like