You are on page 1of 29

Microporous and Mesoporous Materials 65 (2003) 129

www.elsevier.com/locate/micromeso

Calcination behavior of dierent surfactant-templated


mesostructured silica materials
Freddy Kleitz, Wolfgang Schmidt, Ferdi Sch
uth

Max-Planck-Institut fur Kohlenforschung, Kaiser-Wilhelm-Platz 1, D 45470 Mulheim an der Ruhr, Germany


Received 27 March 2003; received in revised form 15 July 2003; accepted 25 July 2003

Abstract
The removal of the template by calcination from mesostructured M41S and SBA-type silica materials was studied by
combining high temperature X-ray diraction, thermogravimetrydierential thermal analysis and mass spectrometry,
allowing detailed in situ investigations during the thermal treatment. A comparison was made between materials with
dierent mesoscopic structures, resulting from dierent synthesis routes and chemical treatment. The in situ XRD
studies showed a strong increase in scattering contrast observed for the low angle reections occurring when the
template is removed from the inside of the pores, irrespective of the type of mesostructure. In agreement with the XRD
investigations, the TGDTA/MS experiments proved that the removal of the surfactant is a stepwise mechanism.
Marked dierences in the scattering contrast variations and chemical reactions were observed depending on the synthesis conditions and the type of surfactant, which highlight the role of the silicasurfactant interfaces. MCM-41 and
MCM-48 materials synthesized in the presence of alkyltrimethylammonium surfactant under alkaline conditions
showed a template removal mechanism based on an Hofmann degradation at low temperatures, followed by oxidation
and combustion reactions above 250 C. On the other hand, acidic conditions employed for the synthesis of SBA-3 type
materials seems to favor reactions of oxidations after the evaporation of water and hydrochloric acid at low temperature. In that case, large contraction of the hexagonal unit cell was usually observed. Most of the block-copolymer
template is removed from SBA-15 at lower temperatures, in a single oxidation step. The SBA-15 framework possibly
catalyzes the oxidation of the block copolymer template species. In addition, the presence of framework porosity or
pore connectivities seems to be responsible for the strong scattering contrast variations observed below 250 C. Residual carbonaceous species and water are removed from the structure upon heating from 300 C up to 550 C. During
this subsequent process a large contraction of the hexagonal unit cell is observed, possibly due to further framework
condensation.
In addition, a brief survey of the previous investigations reported in the literature related to the decomposition of
structure-directing agents is given.
 2003 Elsevier Inc. All rights reserved.
Keywords: Mesoporous silica; MCM-41; MCM-48; SBA-3; SBA-15; Template removal; Scattering contrast; High temperature X-ray
diraction

Corresponding author. Tel.: +49-208-306-2373; fax: +49-208-306-2995.


E-mail address: schueth@mpi-muelheim.mpg.de (F. Sch
uth).

1387-1811/$ - see front matter  2003 Elsevier Inc. All rights reserved.
doi:10.1016/S1387-1811(03)00506-7

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

1. Introduction
Ordered mesoporous materials [1,2] consist of
extended inorganic or inorganicorganic hybrid
arrays with exceptional long-range ordering,
highly tunable textural and surface properties, and
controlled pore size and geometry. Typically, the
structure of the pores is periodic and the pore size
distributions are narrow with pore sizes ranging
between 2 and 50 nm, being known as the mesopore range. During the last decade, intensive scientic eorts have been devoted to synthesis,
characterization and application of such ordered
mesoporous materials [36]. Increasing knowledge
on mesoporous and mesostructured materials has
led to an almost continuous development of new
processes and techniques for synthesis and modication, which overcome previous limitations.
Ordered mesoporous materials are now thought to
nd applications in elds as diverse as catalysis
[3,6,7], separation [8], delivery and release techniques [9], low-k dielectrics, sensors, and other
electro-optical technologies [1013]. Nevertheless,
many aspects of the processes involved in preparing and developing valuable porous materials still
require greater insights. The present article is
concerned particularly with one of the most important aspects of ordered mesoporous materials,
namely the removal of the liquid crystal template,
which is generally used to synthesize most of the
ordered mesoporous solids. Specically, the aim of
this work is to provide some insights on the
physical and chemical processes involved in the
calcination of well-documented mesostructured
M41S [1,14] and SBA-type [15,16] silica materials.
Templating comprises the use of synthesis solution and a template molecule or assembly of
molecules. A template is generally described as a
central structure around which a network forms.
The cavity created after the removal of the template should retain morphological and stereochemical features of the central structure [17,18].
The generation of ordered mesoporous materials is
possible via templating by self-assembled liquidcrystalline phases. When the material has reached
a sucient degree of condensation, the templating
molecules are no longer needed and can be removed to open the porous structure. Since some

composite mesophases can contain as much as


55% of organic material by weight, the removal
procedure of the organics is of utmost importance
in the preparation. Moreover, this step can considerably alter the nal properties of the desired
materials. Ecient template removal and faithful
imprinting have been shown to depend largely on
the nature of the interactions between the template
and the embedding matrix, and the ability of the
matrix to adapt to the template. Ideally, after removal of the core molecules from the surrounding
matrix the shape of the voids that remain should
reect the shape of the template.
The most common method used in laboratories
to remove the template is calcination. In this
method, the as-synthesized materials are alternatively heated in owing nitrogen, oxygen or air,
burning away the organics. Any necessary chargecompensating counterions are supplied from the
decomposition of organics. Usually, the heating
rates required are slow with heating ramps such as
1 C/min up to 550 C, followed by an extended
period of heating at a temperature plateau (48 h).
Calcination of as-synthesized mesophase containing large amounts of carbonaceous species can
leave carbon deposits or coke as a contaminant in
the porous materials, and pore blocking may occur. Generally, when the template is removed by
calcination, the low angle reection intensities increase, the structure may shrink, and the mesoscopically ordered structure could be dramatically
aected [1921]. MCM-41 was originally calcined
at 540 C in N2 for 1 h and then in O2 for 6 h [14].
The reported framework condensation increases
from Q3 /Q4 of about 0.67 in the as-synthesized
MCM-41 precursor (29 Si NMR data) to about 0.25
after calcination. Chen et al. [22] calcined MCM41 samples at 540 C in air for 10 h with a slow
heating rate (1 C/min). They observed up to 25%
decrease of the unit cell constant depending on the
synthesis conditions, a fact that is in strong contrast with crystalline silicates which change very
little upon heating. The reported conditions for
calcination were found to vary widely. However, a
standard general procedure that can be used to
calcine mesostructured silica materials is performed under air at 550 C for 5 h with a heating
rate of 1 C/min. Thermogravimetry was used by

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

Chen et al. [22] to study the thermal behavior of


surfactant containing MCM-41 mesophase. They
recognized three distinct stages of weight loss: (1)
25150 C, due to the desorption of water; (2) 150
400 C, caused by combustion and decomposition
of the template, and (3) above 400 C attributed at
that time only to water loss upon silanol condensation. The rst investigations on calcination were,
however, already described in the early report
from the Mobil scientists on MCM-41 [14]. They
studied the removal of the alkylammonium template by thermogravimetric analysis combined
with temperature programmed amine desorption
analysis (TPAD). The measurements were carried
out on as-synthesized aluminum containing MCM41 from room temperature to 900 C under a ow
of He with titration of the evolving base. They
observed two main weight loss maxima. The molecular weight of the decomposing species in the
low-temperature weight loss was calculated to be
312 g/mol, which is close to the sum of the molecular weights (283 g/mol) expected for decomposition of C16 H33 (CH3 )3 N to hexadecene (224 g/
mol) and trimethylamine (59 g/mol). The amine
desorption analysis suggested the association of
C16 H33 (CH3 )3 N with siloxy groups. The authors
proposed that since the siloxy groups are stronger
bases, they could promote the Hofmann elimination at lower temperatures. An important note is
that the Hofmann elimination of the structuring
molecules is commonly suggested in the cases of
zeolites such as MFI-type ones (ZSM-5 or silicalite-1) [23,24]. Parker et al. [25] proposed a general
mechanism for the thermolysis of the tetrapropylammonium hydroxide that is occluded in the
zeolites with MFI structure. The rst step of this
mechanism is the Hofmann elimination producing
propylene, tripropylamine and water. It is followed by successive b-eliminations. In addition,
depending on the framework structure and compositions, several other types of chemical reactions
may occur in porous solids (e.g. dehydroxylation,
decomposition, cracking, oligomerization, rearrangements) [26,27].
Corma et al. [28] carried out in situ IR studies
of the thermal desorption of the template between
200 and 500 C with a heating rate of 10 C/min.
By increasing the temperature, they observed a

decrease of the interaction of the silanol groups


with the template molecules. Their analysis of the
organic fragments detected above 400 C suggests
that part of the carbon chain has been cracked and
removed. In addition, the appearance of IR bands
assigned to the RNH
3 of the protonated amine
supports the proposed mechanism of Hofmann
degradation of the template.
We described previously the temporal evolution
of hexagonal mesophases of silica (Si-MCM-41),
titania and zirconia as a function of temperature
[19]. Detailed in situ XRD studies with a high
temperature XRD chamber system were conducted
in conjunction with TGDTA and MS. The thermal behavior of cationic surfactant templates in the
mesostructured systems has been analyzed. In the
particular case of Si-MCM-41, an initial change
occurred up to 250 C with an increase in intensity
of all reections, with the (1 1 0) and (2 0 0) reections increasing later and at a higher rate than
the (1 0 0) reection. Above 300 C, changes were
less pronounced and the intensities remained unchanged while the sample was kept at 550 C. The
TGDTA/MS data showed that the decomposition
mechanism in air involves three steps. An initial
endothermic eect, between 150 and 250 C, was
assigned to Hofmann elimination of trimethylamine, leading to a hydrocarbon chain. During this
step (below 250 C), 46% of template was removed
by evaporation of the alkene resulting from the
Hofmann degradation. Keene et al. [29] used
sample controlled thermal analysis (SCTA) coupled with a mass spectrometer to carefully eliminate the organic surfactant and study the evolved
gases in situ during the thermal decomposition of
the template. They found hexadecene to be the
major evolved product in this range of temperatures, conrming the elimination of the trimethylamine and the formation of the alkene by Hofmann
degradation. They obtained additional evidence
for the presence of hexadecene by collecting the
intermediate evolved species during calcination,
and further characterizing them by GC/MS, and
1
H and 13 C NMR. The second step appearing in
the temperature range of 250300 C was shown
to be exothermic and originate from a carbon
chain fragmentation. Several fragments assigned to
shorter chain lengths appeared in this interval

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

accompanied with early oxidation processes producing CO2 and H2 O. We proposed that this step,
corresponding in the XRD to a less pronounced
increase in intensity, results from a successive carbon chain fragmentation or decomposition, with
early oxidation of dierent fragments [19]. Here,
some cracking reactions on the hydrocarbon chain
may have also occurred. Finally, oxidation occurring at 320 C converted the remaining organic
components to carbon dioxide and water.
The removal of triblock copolymers species
(BCP) from as-synthesized SBA-15 was also investigated by TGDTA. Two main processes were
observed: at low temperature (80 C), desorption
of physisorbed water takes place, followed at
higher temperature (145 C) by the exothermic
decomposition of the template [16,30]. The authors suggested that this relatively low temperature decomposition compared to that of cationic
surfactants or the pure block copolymer may be
catalyzed by the inorganic framework.
In contrast to silicates, other compositions are
usually more sensitive to thermal treatments and
calcination can result in breakdown of the structural integrity. Hydrolysis, redox reactions or phase
transformations account for this lower thermal
stability [31]. Therefore, the removal of the surfactant by thermal treatment happens to be more
dicult in the case of non-siliceous mesostructured
materials. We showed recently that template removal appeared to be completely dierent for the
titanium- and zirconium-based materials synthesized with cationic surfactants: a single-step complete oxidation of the surfactant was observed
around 300 C in TGDTA/MS. This was usually
accompanied with a drastic decrease in d-spacing
and initial sharp increase in reection intensity in
the XRD pattern, which generally led to the loss of
the highly ordered mesostructure [19]. Nevertheless, various non-siliceous mesoporous materials
with well-ordered hexagonal and cubic phases
could be successfully obtained [20,3237]. In addition, transition metal-based mesoporous materials could be also synthesized in the presence of
block-copolymers [38], followed by extraction of
the template or calcination at temperatures below
400 C, these temperatures often being the upper
thermal stability limit.

An alternative method for surfactant removal is


based on the extraction of the organic template.
This can be done either by liquid extraction
[22,39,40], acid treatment [41], oxygen plasma
treatment [41], or supercritical uid extraction
[42]. Dried as-synthesized MCM-41 samples are
usually extracted in acid solutions, alcohols, neutral salt solutions, ammonium acetate, or mixtures
of these. For example, Hitz et al. [40] showed that
an Al-MCM-41 sample could be extracted in
acidic media for 1 h at 78 C. Up to 73% of the
template could be removed by extraction with solutions of an acid or salt in ethanol. These authors
showed that when extracting with acidic ethanol,
ion exchange of the counter cations for protons
could be achieved simultaneously. Using strong
acids or small cations was proved to be more efcient for the extraction of the template in ethanol,
suggesting that the size and, thus, the mobility of
the cations in the close-packed micellar aggregates
is one of the factors determining the extent of extraction. Acids with low acid dissociation constant
such as CH3 COOH were less ecient. Moreover, it
seems that more polar solvents are superior to
dissolve the template ions. Accordingly, it is widely
suggested that an ion-exchange mechanism occurs
during solvent extraction of M41S-type materials.
The presence of cationic species in the extraction
liquid for charge balance is therefore essential for
the ion exchange. Various acidic media are used
for surfactant extraction, ethanolic HCl solutions
being the most commonly employed. Dierently,
the HMS [39] or SBA-types [15,16] frameworks
are considered to be relatively neutral, and the
resulting frameworksurfactant interactions are
weak. Such weak electrostatic interactions or hydrogen bonding are more favorable for surfactant
extraction even in the absence of cationic species
since counter cations are not needed. It is, for
example, possible to remove large amounts of cationic surfactant from an SBA-3 mesophase by extraction in boiling ethanol for a short time [43]. The
templates from mesophases obtained with longchain amines [39], as well as transition metal-based
mesophases obtained from the ligand-assisted
method [4446] are usually readily extracted. Also
block copolymers could be extracted from SBA-15
using acidic ethanol solutions for short times and

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

low temperatures [47,48]. Hence, extraction might


provide an alternative to calcination especially in
the case of non-siliceous mesophases, which show
a poor thermal stability. However, the possibility
of extraction of the template molecules depends
strongly on the nature of the interactions between
template and inorganics, and the eciency of this
method relies on a balance between extraction
time and temperature as well as on the composition and concentration of the extraction solution.
Another method to remove the template from
MCM-41 was pioneered recently by Keene et al.
[49,50]. Ozone was used to remove the organic
surfactant species at room temperature from an assynthesized mesophase to form mesoporous
MCM-41. As-synthesized MCM-41 was treated by
ozone using a UV lamp whose wavelength was
known to produce ozone from atmospheric oxygen. The pore size of the resulting ozone-treated
sample was apparently larger, the pore size distribution narrower and the hexagonal long-range
ordering of the pores seems to be improved compared to MCM-41 calcined in an ordinary box
furnace. The unit cell parameter observed for the
ozone treated sample was found to be the same as
for the initial mesophase, which is in contrast with
calcination or ion exchange of the template. The
ozone treated samples seem to exhibit a higher
SiOH group density than the calcined samples.
UVozone treatment was subsequently applied to
remove non-ionic surfactants from mesostructured
silica thin lms. Brinker and coworkers [51]
showed that room temperature UVozone treatment provides an ecient way for the removal of
the template while simultaneously stabilizing the
inorganic silica framework into a well-dened
mesoscopic morphology. Their results established
that ozone treatment leads to complete removal of
the template, strengthens the inorganic framework
and renders the thin lm surfaces highly hydrophilic. The main advantages of ozone treatment
over conventional thermal treatments are that
elimination of organic molecules at room temperature might be applicable to thermally unstable
mesostructured materials, and that no organic
solvents are needed. However, it seems that the
rst ozone treatments performed on titania-based

mesophases led mainly to uncontrolled ozonation


resulting in a highly exothermic reaction and the
loss of mesoscopic order [52].
In the present contribution, we choose to focus
on materials based on silica. The rst candidates
for in situ investigation of the removal of the
templating species are materials that are synthesized under alkaline conditions (MCM-41 and
MCM-48). Acid-prepared mesotructures such as
SBA-3 and SBA-15 will then be discussed. Comparison is made between materials with dierent
mesoscopic structures, synthesized with surfactant
with dierent chain lengths, and resulting from
dierent synthesis routes and chemical treatment.
The removal of the template by calcination is
studied by combining high-temperature X-ray
diraction, thermogravimetrydierential thermal
analysis and mass spectrometry, allowing in situ
investigations during the thermal treatment.

2. Experimental section
2.1. Materials
All materials described in this section are synthesized according to published standard procedures. The calcinations were performed at 550 C
for 5 h in all cases.
MCM-41-type materials were all prepared
within 2 h according to the method described by
Gr
un et al. [53]. The synthesis is based on the use
of TEOS (0.05 mol) as the silicon source, with
ammonia (0.14 mol) as the catalyst and an aqueous solution of surfactant (6.6 103 mol in 120 g
H2 O). n-Alkyltrimethylammonium bromides of
dierent alkyl chain lengths, CnTAB with n 12
18, were used as template. The original Gr
un
synthesis was modied in 1999 [54] and Si-MCM41 samples with apparent increased structural order were achieved by aging the materials in the
mother liquor at 90 C for 7 days. n-Alkyltrimethylammonium bromides can be substituted by
an equimolar amount of n-hexadecylpyridinium
chloride (cetylpyridinium chloride, CPCl) for an
alternative MCM-41 synthesis. The as-synthesized
materials are subsequently dried at 90 C overnight.

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

High quality MCM-48 is obtained following the


hydrothermal method described by Fr
oba et al.
[55] using TEOS (0.02 mol) as the silicon source,
and CTAB (0.013 mol) as the template in the
presence of KOH (0.01 mol) and water. As-synthesized materials were isolated after 35 days of
hydrothermal treatment at 115 C, washed with
H2 O and dried.
As-synthesized SBA-3 is obtained at room
temperature by slow addition of TEOS (0.096 mol)
to an aqueous acidic (HCl) solution of CTAB
(0.012 mol) according to the standard procedure
[15].
SBA-15 is normally synthesized according to
the acidic synthesis route using EOPOEO triblock copolymers of the pluronic-type as structure
directing agents [16,30]. SBA-15 was prepared with
TEOS as the silicon source and EO20 PO70 EO20
(P123) as the template in an aqueous HCl solution,
according to Zhao et al. [16]. The synthesis is
carried out for 24 h at 40 C followed by 24 h at
90 C.
The detailed structural parameters resulting from the in situ XRD investigations and the
physicochemical parameters obtained by N2 sorption of all materials described in this report are
summarized in Tables 1 and 2. The results are in
line with the data published previously on these
materials.
2.2. Characterization methods
The purpose of the present study is to investigate the removal of the template upon thermal
treatment from mesostructured frameworks. For
this, in situ high temperature powder X-ray diffraction methods have been used.
The in situ high temperature measurements
were recorded on a Stoe STADI P hh powder
X-ray diractometer in reection geometry (Bragg
Brentano) using Cu-Ka12 radiation with secondary monochromator and scintillation detector.
A high temperature X-ray diraction chamber
(Johanna Otto HDK S1), with a Pt/Rh heating
element as a sample holder, was mounted on the
goniometer. XRD patterns were typically recorded
with an automatic divergence slit conguration
(ADS, receiving slit xed at 0.8 mm), except SBA-

15 samples, which were measured in a xed slit


conguration (FS). XRD patterns were typically
recorded in a range of 18 (2h) with step 0.05
(2h) and time/step 4 or 8 s. SBA-15 was measured in a range of 0.83 (2h) with step 0.02
(2h) and time/step 5 s.
For in situ high temperature experiments, all
the samples in the as-synthesized form containing
the templating species were ground prior to analysis. A small amount of ethanol or hexane was
used to disperse the materials homogeneously on
the sample holder. The samples were heated stepwise with a heating rate of 5 C/min, up to the nal
calcination temperature. This temperature was
maintained for several hours, then the sample was
cooled down. XRD measurements were performed
every 50 C during the heating process, every hour
during the isothermal heating, and at various
temperatures upon cooling. To adopt the usual
oven calcination conditions, all the measurements
were carried out in air. The thickness of the sample
preparation was about 0.20.3 mm, and the sample surface was homogeneous. The preparation
has to be very thin to avoid any temperature gradient in the sample during thermal treatment.
Therefore, the sample temperature is considered to
be close to the temperature of the Pt/Rh band
sample holder. Phenomena of hot spots and
rapid overheating of the sample bed (glow eects)
[56] can be neglected. For a diractometer, the
instrumental error limits the precision of the dspacings measured at low angle for each mesostructured materials of interest. Discrepancies may
originate from dierences in sample preparation,
since dierences in scattering volume and sample
packing density can lead to dierent diraction
patterns in terms of signal-to-noise ratio. Moreover, the accuracy, with regard to the positions of
the reections and their intensity, for measurements performed with material deposited on the
Pt/Rh band is strongly inuenced by the homogeneity and the thickness of the preparation.
Therefore, materials treated in situ and under
conventional conditions in a box furnace were
compared to test the reproducibility and validity
of the measurements. Even when signicant lattice
shrinkage is observed upon the removal of the
template, the matching positions of the low angle

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

Table 1
Results of the in situ XRD investigations carried out on dierent surfactant-templated silica mesophases
Materials

aassynthesized
(nm)

acalcined (550 C)


(nm) (%)a

I(1 1 0):I(2 0 0)
as-synthesized

I(1 0 0):I(1 1 0)b


calcined (RT)

I(1 1 0):I(2 0 0)
calcined (RT)

10
4
7.3
9.7

1.1
0.9
1
1.1

15
5
10.5
6.5

2
1.1
1.55
1.6

C12-MCM-41
C14-MCM-41
C16-MCM-41
C18-MCM-41

3.73
4.17
4.64
5.25

3.51
3.82
4.16
4.80

C14-MCM-41-aged
C16-MCM-41-aged
C18-MCM-41-aged

4.35
4.75
5.46

4.15 (4)
4.6 (3)
5.15 (5)

10.5
6.85
6.25

1.2
1.2
1.25

9
4.2
5

1.3
1.4
1.4

4.45
9.60
4.53
4.40
12.11

3.95 (11)
8.33 (13)
3.73 (18)
3.8 (14)
10.52 (13)

6.3
5.9c
12
13
16.9

0.9

1.3
1.5
0.8

6.75
11c
37.5
25
15.9

1.3

1.5
1.85
1.35

CPCl/MCM-41
MCM-48
SBA-3
SBA-3 extracted
SBA-15d

(6)
(8)
(10)
(9)

I(1 0 0):I(1 1 0)b


as-synthesized

Lattice contraction (%).


The intensities were measured with a diractometer in reection geometry and automatic divergence slit (ADS) conguration
(average values with subtraction of the background scattering). The experimental error on the intensity ratios is estimated to be 10%.
c
Ratio (2 1 1):(3 3 2).
d
The XRD measurements were carried out in xed slit conguration (FS).
b

Table 2
Physico-chemical parameters observed for the calcined materials, obtained by nitrogen physisorption
Materials

BET surface
areaa (m2 /g)

Total pore volume, Vp (cm3 /g)

Pore size,
wBJH des: (nm)

Pore size,
wd b (nm)

Wall thickness,
bBJH c (nm)

Wall thickness,
bd d (nm)

C12-MCM-41e
C14-MCM-41
C16-MCM-41
C18-MCM-41

1035
1100
1130
995

0.51
0.61
0.78
0.79

2.06
2.12
2.47
2.90

2.55
2.89
3.31
3.83

1.45
1.7
1.7
1.9

0.96
0.93
0.85
0.97

C14-MCM-41-aged
C16-MCM-41-aged
C18-MCM-41-aged

910
1010
1015

0.57
0.80
0.85

2.32
2.82
3.62

3.10
3.67
4.16

1.85
1.8
1.55

1.05
0.93
0.99

CPCl/MCM-41
MCM-48
SBA-3f
SBA-15f

995
1175
1470
737

0.58
0.72
0.65
0:67Vpmeso
0:13Vpmicro

2.24
2.16
2.04
5.42

2.96

2.86
8.07g

1.7

1.7
5.10

0.99

0.87
2.45


1=2
Average BET surface area.
Vp q
Obtained from the geometrical model with equation wd cd100
, c 1:155 (hexagonal pores) and q 2:2 g/cm3 .
1 Vp q
c
Wall thickness calculated as a  wBJH .
d
Wall thickness calculated as a  wd .
e
Limit of the BET equation accuracy.
f
The presence of microporosity in the silica walls makes the use of the BET equation and the t-plot method likely inaccurate re
1=2
sulting in discrepancies.
Vpmeso
g
Obtained from the geometrical model with equation wSBA15
cd100
c 1:213 (circular pores) and
d
1=q Vpmeso Vpmicro
q 2:2 g/cm3 [86].
a

reections indicated a good reproducibility and


proved the in situ XRD measurements to be appropriate [57] (see for example inset in Fig. 10a). A

rough estimation of the measurement error during


the dierent calcination protocols can be made
giving average variations between in situ and

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

conventional calcination procedures of about 0.1


nm for the repeat distances at low angle. Furthermore, it is recognized that an absolute error of
about 0.1 nm due to the limits of the synthesis
reproducibility can be observed for the d value
measured for several syntheses of a same material.
It is therefore reasonable to assume an absolute
experimental error of 0.10.2 nm.
In order to study the chemical and physical
aspects of the surfactant degradation within the
mesopores, thermogravimetry (TG) in combination with dierential thermal analysis (DTA) experiments were carried out. The conventional
thermobalance was coupled with a quadrupole
mass spectrometer. This setup allows the characterization of the species evolved during thermal
treatment [58]. The thermogravimetric analyses
combined with dierential thermal analyses were
performed on a Netzsch STA 449 C thermobalance coupled with a Balzers Thermostar 442 mass
spectrometer (temperature of the transfer capillary
was 160 C). The measurements were carried out
under air with a heating rate of 5 C/min for assynthesized samples. All TGDTA and MS results
are reproducible within an error estimated to be
10 C.
The N2 sorption measurements were performed
on a Micromeritics ASAP 2010 adsorption unit.
Prior to the measurements, the calcined samples
were activated under vacuum for 5 h at 200 C.
The measurements were performed at 77 K using a
static-volumetric method. The empty volume was
measured with helium gas.

3. Results and discussion


3.1. Materials synthesized under alkaline conditions
(S I )
The mesophase formation under alkaline conditions is based on cooperative electrostatic interactions between negatively charged oligomeric
silicate species I and positively charged surfactant
molecules S [15,59]. In general, one considers the
mesostructured silica network obtained under alkaline conditions to contain signicant amounts of
negative charges.

3.1.1. MCM-41 (Grun synthesis) [53]


The standard reference material is pure SiMCM-41 synthesized in the presence of cetyltrimethylammonium bromide (CTAB, n 16) as
the template, denoted C16-MCM-41. The d(1 0 0)spacing of the well-resolved hexagonal p6m phase,
measured ex situ prior to calcination, is about 4.05
nm, giving a unit cell constant aas-made 4:68 nm.
After the removal of the template by thermal
treatment, the hexagonal phase is retained (Fig.
1a). The d-spacing of the (1 0 0) reection is however shifted to about 3.55 nm, resulting in
acalcined 4:1 nm, which corresponds to a lattice
shrinkage of about 12%. The unit cell constant
aas-made for a C16-MCM-41 aged at 90 C for
7 days increases slightly with d1 0 0 4:2 nm
and a 4:85 nm. In contrast, however, the lattice
shrinkage upon thermal treatment is substantially
reduced. The d(1 0 0) of the calcined sample aged
at 90 C is 4.15 nm with acalcined 4:79 nm, indicating a lattice shrinkage of only 12%. This fact
suggests that materials obtained after an aging
period have a higher thermal stability. Furthermore, the diraction pattern obtained on a material aged for several days seems to exhibit a better
resolution with a higher signal-to-noise ratio
compared to MCM-41 synthesized at room temperature. Assuming a same volume of matter for
the X-ray preparation, one may suggest a higher
degree of order and/or a larger coherent scattering
domain size. The relative I(1 0 0):I(1 1 0) and
I(1 1 0):I(2 0 0) ratios measured ex situ after calcination are 21 and 1.5 for C16-MCM-41-aged, and
23.5 and 1.3 for C16-MCM-41 synthesized at
room temperature, respectively. The N2 sorption
isotherms in Fig. 1b show that a capillary condensation step is present in both cases. The position of the capillary condensation step is shifted to
higher relative pressures for the aged MCM-41
material, indicating a larger pore size, in agreement with the XRD data. Furthermore, the aged
sample presents a steeper increase during the
capillary condensation suggesting a narrower pore
size distribution. The adsorption capacity and
surface area are comparable.
We have shown in a previous study concerned
with in situ XRD investigations during calcination
of Si-MCM-41 that an initial intensity change

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

4.00

Intensity
Intensity

3.50
I(100) conventional
conventional sample 3.25
aged sample

Si-MCM-41 aged
x8
RT 150 250 350

450 550

550

d-spacing [nm]

3.75

3.00
150 RT

Temperature [C]
Si-MCM-41 conventional

2.4

10

2 theta []

Intensity

(a)

2.2
2.0
1.8

Volume adsorbed [cm3/g]

600

I(110)
I(200)

1.6

500
RT 150 250 350
400

450 550

550

150 RT

Temperature [C]

300
Si-MCM-41 conventional

200

Si-MCM-41 aged
100
0
0.0

(b)

d(110)
d(200)

d-spacing [nm]

x8

0.2

0.4

0.6

0.8

1.0

P/P0

Fig. 1. (a) Comparative X-ray diraction patterns of a C16MCM-41 sample synthesized at room temperature (bottom)
and C16-MCM-41 after aging at 90 C for a week (top). Both
samples are calcined. The patterns were obtained ex situ on a
diractometer in transmission geometry. (b) N2 sorption isotherms at 77 K obtained on a C16-MCM-41 sample synthesized
at room temperature (open symbols) and C16-MCM-41 after
aging at 90 C for a week (solid symbols).

occurs up to 250 C with an increase of all reection intensities [19]. Fig. 2 shows the details of
the evolution of the maximum intensities of the
reections at low angles recorded in situ for C16MCM-41 synthesized at room temperature (conventional) and illustrates the evolution of the
d-spacings during the calcination. The graph clearly
shows a stepwise evolution of the reections (room
temperature200 C, 200350 C, 350 Ccooling

Fig. 2. Evolution of the reection intensities (solid symbols) of


C16-MCM-41 (CTAB/MCM-41) as a function of temperature
(calcination at 550 C for 5 h). Also plotted are the d-spacing
values (open symbols) of the respective reections. The triangle
symbols, in the top gure, show the evolution of the reections
of MCM-41 aged at 90 C. Top: evolution of the (1 0 0) reection. Bottom: evolution of the (1 1 0) and (2 0 0) reections.
The connecting black solid lines are used as guide for the eye.

process), with the d-spacing values following a


similar temperature dependence. The d(1 0 0) is
reduced by about 0.45 nm, the largest d-spacing
shift occurring between 150 and 250 C. A subsequent less pronounced shrinkage takes place
mostly above 400 C up to the maximum temperature applied. During the cooling process, a
slight loss in scattering intensity is observed, indicating a loss of scattering contrast. This eect is
reversible, since heating the sample leads to the
recovery of the intensity, and it can be attributed
to physisorbed water condensing in the pores.
The faster increase of the (1 0 0) intensity below
200 C relative to the (1 1 0) and (2 0 0) reections
is clearly observed. In addition, the scattering intensities of the (1 1 0) and (2 0 0) reections reach
a maximum at 350400 C, before decreasing

10

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

linearly (bottom graph). The general strong increase in intensity appearing in XRD with increasing temperature is due to the higher scattering
contrast between the pore walls and the inside of
the pores, caused by the burning-out of the templating organic species. As the relative intensities
of the low angle reections are very dependent on
the distribution of matter in the pores [6063], we
assume that selective removal and redistribution of
surfactant fragments in the pores may be responsible for the dierent growth rates of the individual
reections. In particular, the (1 0 0) reection intensity is shown to be very sensitive to the density
contrast between framework walls and mesopores.
Recent works on simulations of XRD pattern
allow precise investigation of the inuence of matter distribution in the unit cell on the diraction
pattern. Hammond et al. [61] could explain the
dramatic change in the X-ray scattering that occurs upon removal of the template from MCM-41
by using a lattice model with hexagonal channels
and a dierent scattering form factor for the wall
and the matter in the channel. They proved that
the (1 0 0) intensity increase upon calcination arises
from dierent phase cancellation between scattering from the wall and the pores. Therefore, once
the template is removed from the pore region, the
eect of this phase cancellation is reduced, leading
to an enhanced scattering intensity. In a recent
series of studies, Solovyov et al. [64,65] used
Rietvelds method in combination with continuous
electron density representations to achieve structural modeling of MCM-41 materials, before and
after removal of the n-alkylammonium template.
These authors compared experimental and calculated XRD patterns and modeled the averaged
density distribution in the materials. They could
show by electron density distribution maps the
decrease of the density within the mesopores that
accompanies the diraction peak increase during
calcination. Most interestingly, they also showed
that the surfactant distribution in the mesopores of
as-synthesized samples obtained under alkaline
conditions is not uniform, with a distinct minimum
in the pore center. This fact might play a signicant role in the contrast evolution of the diffraction peaks during the template removal from
Si-MCM-41. In addition to this, the surface

roughness and the presence of molecules bonded


to the inorganic surface within the pores may also
induce strong variation in scattering contrast at
low angle [63].
The sample aged at 90 C shows a similar
evolution of the reection intensities observed in
general for all reections during thermal treatment
according to the three step process as described for
the conventional material. From the graph, the
remarkable stability of the mesophase treated at 90
C is evidenced. The sample aged at 90 C shows
only a small d-spacing shift (0.1 nm) to lower
values, mostly between room temperature and 150
C, as physisorbed water is removed. It is reasonable to propose that the lower shrinkage of the
unit cell of an aged MCM-41 is a result of a better
wall condensation [66]. The increased thermal
stability and structural order are achieved since
hydrolysis of the silicon source and further condensation of the inorganic network are enhanced
during the aging period performed in the mother
liquor, leading to higher cross-linking of the inorganic species making up the walls [67]. On the
other hand, the expulsion of electrolytes and
surfactant molecules may play an additional role
in the packing of the surfactant and the condensation of the mesophase [68,69]. One has to keep in
mind, however, that depending on the synthesis
conditions, undesired structural degradation and
loss of pore uniformity could occur upon prolonged hydrothermal treatment at high temperatures, due to re-hydrolysis and dissolution of the
framework [70].
The TGDTA results showed that three main
processes take place upon heating C16-MCM-41,
in agreement with Zhao et al. [71]. At temperatures
below 150 C, physically adsorbed water is removed. The following stage occurs between 150
and 350 C and corresponds to the decomposition
of the organics. Finally, an additional weight loss
is measured at higher temperatures up to 600 C,
often assigned to dehydroxylation of silanol and
residual coke combustion. Our previous TG
DTA/MS studies revealed that the mechanism of
removal of the organic template involves three
precise steps [19]. The initial endothermic eect
between 150 and 250 C is caused by the elimination of the trimethylamine head group, via

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

Hofmann degradation, which leads to a hydrocarbon chain (m=z 26, 41, 42, 55, 69). The second eect in the TGDTA is exothermic and takes
place in the temperature range of 250300 C.
Several fragments assigned to shorter chain
lengths (m=z 26, 41, 42) appear in this interval
accompanied with early oxidation processes producing CO2 (m=z 44), NO2 (m=z 30, 46), and
H2 O (m=z 18). We proposed that this step results from a successive carbon chain fragmentation
or decomposition, with early oxidation reactions.
Finally, the major part of the oxidation occurs
between 300 and 350 C and converts the remaining organic components (18%) to carbon dioxide, water, and probably residual carbonaceous
species. After the oxidation processes up to 350 C
have been completed, about 15% of the organics
remain in the material up to higher temperature.
These template residues are probably carbonaceous species since only small quantities of water
are produced from SiOH condensation beyond
350 C. The removal of the template from mesostructured samples of MCM-41 aged 7 days at 90
C proceeds via the same reaction scheme. The
thermal stability of the aged materials is, however,
higher. Furthermore, a lower amount of surfactant is contained in the as-synthesized mesophase
(Table 3). Total weight losses of 46% and 40% are
measured for C16-MCM-41 and C16-MCM-41
aged, respectively.
All materials synthesized with nc 1218 exhibit XRD patterns (not shown) showing a wellresolved hexagonal mesophase indexed to the p6m

11

symmetry. As expected, the interplanar distance


d(1 0 0) increases with increasing alkyl chain
length. The unit cell size and pore size of the
calcined materials are found to be determined by
the alkyl chain length of the cationic surfactant
used (Tables 1 and 2). The in situ XRD results
obtained for materials synthesized with alkylammonium bromide surfactants having alkyl chains
with nc 12, 14 and 18 are obviously similar to
the ones obtained with nc 16 (Fig. 3), suggesting
a similar stepwise process of template removal. A
decrease of the distance d is observed for all
samples predominantly before 300 C, similar to
that observed with C16-MCM-41. The scattering
intensities of all reections increase progressively
at dierent growth rates as the template species are
burnt out of the channels. The variations in the
growth rates with dependence on the surfactant
chain length, observed at the early stages of the
heating process (temperatures below 300 C), seem
to suggest that the growth rate in scattering intensity for materials synthesized with shorter
surfactant chains could be slightly lower (Fig. 3b).
The direct comparison of the relative intensities at
low 2h angles for materials having dierent unit
cell size is critical due to the dependence of the
scattering form factors with diraction angle. In
general, higher scattering intensities are expected
with decreasing 2h angles. However, since the unit
cell size at the dierent temperature steps are
comparable for a material, and the changes with
angle in the range of interest are small [63], we may
relate this eect to the respective ratio of wall

Table 3
Mass losses recorded by thermogravimetry for MCM-41 samples synthesized with surfactants having dierent chain lengths (mass
losses below 110 C attributed to physisorbed water)
Samples

25110 C
(%)

110265 C
(%)

265305 C
(%)

305395 C
(%) (lV/mg)a

3951000
C (%)

Total mass
loss (%)

C12-MCM-41
C14-MCM-41
C16-MCM-41
C18-MCM-41
C14-MCM-41 aged
C16-MCM-41 aged
C18-MCM-41 aged

3
3
2
3
4
2
3

24
24
21
19
21
19
17

5
6
9
10
6
7
10

3
5
8
9
4
7
9

4
5
6
7
4
5
6

39
43
46
48
39
40
45

(0.5)
(0.6)
(0.8)
(1.10)
(0.7)
(0.9)
(1.2)

a
We provide values obtained for energy released associated with the main exothermic DTA peak to highlight qualitatively the
trends. Precise quantitative calorimetric data can be obtained from additional DSC experiments.

12

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

d-spacing [nm]

4.6

C18
C16

C14
C12

4.2

3.8

3.4

3.0
RT 150

250

350

(a)

450

550

550

150 RT

Log10 intensity

Temperature [C]

C12
C14
C16
C18
RT

(b)

150

200

250

300

350

400

450

500 550

Temperature [C]

Fig. 3. (a) d(1 0 0) values recorded as a function of temperature


for the materials synthesized in the presence of alkylammonium
surfactants with nc 1218. (b) Log10 plot of the intensities of
the low angle reections during the heating ramp for the same
materials (top: plots for (1 0 0), bottom: plots for (1 1 0)).

thickness to pore size. The size of the pore is


shown to decrease with shorter carbon chains,
with the wall size bd remaining relatively constant
around 1 nm (Table 2), in agreement with the
published data [70]. To gain further evidence, a
detailed comparative study of experimental XRD
data and series of simulated XRD patterns would
be highly informative here. The in situ data obtained on aged samples proved the higher thermal
stability of all Cnc -MCM-41 aged samples compared to their conventional counterparts.
The TGDTA/MS measurements performed on
these materials show that the weight loss is found
to increase as the alkyl chain length of the surfactant increases (Table 3). This fact could be attributed to an increasing size and molecular weight
[72] of the micellar aggregates relative to the
amount of inorganic matter that is embedding the

templating species. The lower amount of organics


contained in the material after aging may result
from re-dissolution of some of the template in the
mother liquor and/or expulsion of template molecules upon the course of condensation. For all
samples, the TGDTA/MS indicates a similar
mechanism of decomposition of the organic template involving three steps. A general trend is that
the proportion of organics removed at lower
temperature (between 150 and 260 C) seems to be
reduced with increasing surfactant chain length.
On the other side, the relative fraction of organics
that undergo subsequent decomposition and oxidation reactions at higher temperatures increases.
This eect is particularly marked for C18-MCM41. However, it remains unclear whether it is
caused by a lower yield of the Hofmann degradation in favor to oxidation processes, or by mass
transfer limitations of the larger organic species
produced by the elimination. In addition, if masstransport is a rate-limiting step, it is expected that
diusion control could result in much broader
DTA peaks, such as observed at high temperatures
(above 400 C) during the calcination of smaller
pore metal oxophosphates [19].
To facilitate the comparison, the weight change
derivatives can be calculated from the weight
change curves measured in TG (Fig. 4). The decomposition/desorption of the cationic surfactant
gives rise to pronounced peaks in the weight
change derivatives at 150400 C. In all cases, the
peaks indicate three temperature ranges for the
decomposition of the template, corresponding to
the stepwise process. An additional broader peak
appears at about 550 C, increasing in intensity
with the alkyl chain length increasing.
In Fig. 5 are plotted various molecular species
recorded during TG/MS, that are attributed to the
carbon chain of the surfactant (m=z 26, 42, 55)
and the trimethylammonium head group (m=z
59) of the dierent MCM-41 samples. This gure
shows the decrease of the fraction of the organic
chain removed at lower temperatures relative to
the fraction removed during decomposition and
oxidations at higher temperatures. This eect is
particularly marked for smaller molecular fragments. Furthermore, the fragments m=z 59 attributed to the trimethylammonium head group of

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

C12-MCM-41

C14-MCM-41

C16-MCM-41

-0.6
-0.8
-1.0

DTG [%/C]

C18-MCM-41 -0.2
-0.4

-1.2
-1.4

100

200

300

400

500

600

700

800

900

Temperature [C]
Fig. 4. Weight change derivatives of MCM-41 synthesized with
alkyltrimethylammonium surfactants with increasing carbon
chain length. C12, C14 and C16-MCM-41 are shifted for clarity
by 1.8, 1.2 and 0.6%/C, respectively. The dotted line indicates
the limit temperature (250 C) between Hofmann elimination
and oxidation.

the surfactants are detected at lower temperatures


for the longer chains surfactants. Interestingly, the
plots obtained for the C18-MCM-41 show an additional second step or shoulder in the lower
temperatures range in the MS traces. This new step
is probably caused by surfactant molecules involved in dierent types of surface interactions
with the inorganic framework.
The nature of the interaction between the head
group of the surfactant and the silica surface seems
to play a crucial role in directing the processes
involved in the removal of the surfactant. The inuence of the surfactant head group might be
probed by changing the nature of the polar head
group. Khushalani et al. [73] demonstrated rst
that MCM-41 could also be synthesized in the
presence of cetylpyridinium chloride (CPCl) as the
template. It is believed that the CPCl/MCM-41
mesophase formation is possible according to the
S I route because this surfactant is cationic, and
has a similar aggregation number and CMC value
to CTACl, which is often used. However, CPCl
allows also for variation in the charge density at

13

the head group. In addition, it is suggested that the


micellar aggregates formed during the synthesis
exhibit an increased rigidity compared to other
alkylammonium surfactants and that the interaction between the pyridyl ring of the surfactant and
silica is stronger. Therefore, it is likely that the
removal of the surfactant species upon thermal
treatment could proceed dierently.
The XRD patterns stack plot of MCM-41 obtained from the Gr
un synthesis with CPCl is
shown in Fig. 6a. The d(1 0 0) value measured for
the as-synthesized materials is about 3.9 nm,
comparable to CTAB/MCM-41 (C16-MCM-41).
The XRD patterns indicate that the reections
observed at low angles are narrower and with a
higher signal-to-noise ratio than CTAB/MCM-41,
in agreement with the literature statements [54,73].
From the developing XRD patterns, it is seen that
the scattering intensities of all reections at low
angles remain constant up to 250 C, which suggests no drastic changes in the distribution of
matter in the pores and/or low contrast variation
between pore walls and the inside of the channels.
From 250 C, all intensities increase drastically,
with the (1 0 0) reection increasing at a slightly
higher growth rate, reaching their maxima at the
calcination plateau of 550 C. From 550 C, the
intensities remain constant until the cooling starts
where variations in contrast are again observed,
due to water adsorption. The (1 1 0), (2 0 0), and
(2 1 0) reections are retained after calcination. In
graph 6b are illustrated the evolution of the reection intensities and the respective values of the
d-spacings. The simultaneous increase in intensity
of all reections is observed after 250 C. The
comparison with CTAB/MCM-41 emphasizes the
dierent behavior. The rst noticeable decrease in
d-spacing occurs up to 250300 C. Subsequently,
a second observable shrinkage is measured between 500 C and the rst hour at 550 C.
The TGDTA/MS data recorded on CPCl/
MCM-41 are depicted in Fig. 7. The rst stage is
the removal of the physisorbed water. It is followed by the main conversion of the template
species between 180 and 450 C which is a twostep process (200300 and 300450 C). During
this stage, only exothermic peaks are detected, suggesting oxidative decomposition and burning-out

14

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

C2H2
m/z = 26

100 200 300 400

500 600 700

C3H6
m/z = 42

100 200 300 400

N(CH3)3
m/z = 59

C4H7
m/z = 55

100 200 300 400

500 600 700

Temperature [C]

500 600 700

100 200 300 400

500 600 700

Temperature [C]

Fig. 5. Plots of various molecular species recorded with MS on C12, C14, C16 and C18-MCM-41 (from top to bottom) and their
evolution with temperature.

of the organic species. The major exothermic


process takes place between 290 and 420 C (centered at 340 2 C), similar to CTAB/MCM-41.

The total mass loss recorded on the mesophase


synthesized with CPCl is comparable to that of
CTAB/MCM-41. However, the removal of the

15

Intensity

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

2.0

3.0

4.0

5.0

6.0

7.0

RT
550 150C
C

45 550
RT 250C 0C C

2 theta []

(a)

4.0

3.0
100
110
200

2.5

d-spacing [nm]

Intensity

3.5

2.0

(b)

RT 150 250 350 450 550

550

1.5
150 RT

Temperature [C]

Fig. 6. (a) XRD patterns stack plot of MCM-41 obtained from the Gr
un synthesis with CPCl as template. Shown are subsequent
XRD patterns as the material is calcined up to a temperature of 550 C, held at this temperature for 5 h and cooled to room temperature. (b) Evolution of the reection intensities of CPCl/MCM-41 as a function of temperature (calcination at 550 C for 5 h). Also
plotted are the d-spacing values of the respective reections (open symbols). Represented with a cross-dotted line is the intensity of
(1 0 0) of CTAB/MCM-41 as reference.

template occurs dierently as highlighted by the


comparative curves in Figs. 6b and 7. The presence
of the fragment assigned to the pyridyl head group
(m=z 79) appearing around 250 C suggests that
the surfactant molecule is rst decomposed into
two species, probably via an elimination-type reaction or cracking. However, signicantly lower
amounts of the carbon chain species are detected
at this temperature range in relation to the high
fraction evolved at higher temperatures. At this
step, only 24% of the surfactant species are removed (11% in weight loss). This may explain the
modest growth rate of the XRD reections inten-

sity. Subsequently, 44% of the organic template


(20% in weight loss) is removed during the main
oxidation step between 300 and 450 C with a
relatively high energy release. This process is accompanied by the simultaneous increase of all reection intensities in the XRD patterns (Fig. 6).
It seems that the major part of the surfactant is
removed by exothermic decomposition and oxidation processes, likely increasing the production
of coke during the calcination. The CPCl/MCM41 sample studied here contains about 15% of
carbonaceous species after the main oxidation reaction (>450 C), which is substantially higher

16

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

Intensity of mass signal

255C

C5H5N m/z = 79

CO2 m/z = 44

H2O m/z = 18
CxHy m/z = 42
C16-MCM-41
-1.5 %
-0.2

-11%
90
exo
80

-0.6

-20%
-1.0

70
-1.4
60

-15%

341C

-1.8

100 200 300 400 500 600 700 800

DTA signal [uV/mg]

Weight loss [%]

100

CxHy m/z = 42,55

900

Temperature [C]
Fig. 7. TGDTA/MS measurements performed on a CPCl/
MCM-41 mesophase. Bottom: TGA data with a black dashed
line with its rst derivative curve (gray curve), and the DTA
curve with a solid line. The added grey dotted line corresponds
to the TGA data of a MCM-41 sample synthesized with CTAB
(C16-MCM-41). Top: molecular species recorded from the MS
and their evolution with temperature.

than the amount estimated for CTAB/MCM-41


(Table 3). These carbonaceous species represent
approximately 33% of the organics. The residual
species are converted to CO2 at temperatures between 450 and 550 C, with an additional shrinkage of the hexagonal structure (Fig. 6), and an
additional X-ray scattering intensity increase. The
major dierence compared to CTAB/MCM-41 is
the lower weight loss occurring during elimination
and larger amounts of coke produced. The carbonaceous residues are however readily removed
to yield complete template removal. It seems that

the pyridinium head group surfactant undergoes


Hofmann elimination less readily than the trimethylammonium head group surfactant does.
The dierent diraction peak intensity ratios
I(1 0 0):I(1 1 0) and I(1 1 0):I(2 0 0) calculated for all
MCM-41 materials, as-synthesized and calcined,
are listed in Table 1. These ratios were obtained in
automatic slit divergence conguration, which
usually results in an increase of the reections at
higher angles compared to the (1 0 0) reection, and
therefore smaller values than those measured ex situ
in transmission geometry. Marked dierences in
the relative intensities between the reections at
low angles are usually expected for materials with
dierent lattice size, having dierent pore size or/
and wall thickness, depending largely on the ratio
of the wall thickness to the pore diameter. However, within the experimental error, the ratios
I(1 0 0):I(1 1 0) measured before and after calcination are comparable. Conversely, the I(1 1 0):
I(2 0 0) ratio seems to increase for all MCM-41
samples after removal of the template. The eect is
more pronounced for materials that have not been
aged at 90 C, and might therefore originate from
increasing condensation features [69]. The stronger
increase of both ratios seen for the C12-MCM-41
sample is likely caused by a decrease in the sample
ordering and the loss of the reection at higher
angles.
Except for CPCl/MCM-41, the results described
herein seem to conrm our previous statement
that two specic thermal behaviors can be distinguished during the calcination of samples synthesized with n-alkylammonium surfactants under
alkaline conditions, implying the presence of
surfactant molecules involved in two types of interactions within the mesopores and with the silica
surface [19]. The presence of non-uniform distribution of surfactant molecules in the mesopores,
as revealed by Solovyov et al. [64,65], is in agreement with dierent locations of the surfactant
species, and possible dierences in the inorganic
surfactant interactions.
3.1.2. MCM-48 (cubic Ia3d phase)
In comparison to MCM-41, the synthesis of
MCM-48 is more dicult. However, high quality
MCM-48 is obtained following the method de-

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

material. The contraction of the unit cell upon


calcination is about 15%, and the well-ordered
structure is retained.
The in situ XRD patterns stack plot obtained
on MCM-48 (Fig. 8a) shows the same scattering
intensity behavior as that of MCM-41. All intensities increase drastically as the calcination
proceeds, resulting from changes in scattering
contrasts. The (2 1 1) and (2 2 0) reections increase
at a higher growth rate than the other reections,
probably due to a dierent distribution of matter
inside the pores, similarly to Cn-MCM-41. From
300 C, changes in intensity are less pronounced.
The result shows the good thermal stability of the

420
332
431 422
521
611
541
543

220

321
400

Intensity

211

scribed by Fr
oba et al. [55] using TEOS as the
silicon source, and CTAB as the template in the
presence of KOH and water. As-synthesized materials were isolated after 35 days of hydrothermal treatment at 115 C. The high quality of all
as-made MCM-48 samples obtained (Fig. 8a inset)
is indicated by the presence of at least eight reections in the XRD patterns [74]. A pure cubic
phase is obtained. The average d(2 1 1) value is
3.97 nm for the as-synthesized material (aas-made
9:73 nm). After calcination, the reections are
shifted to higher 2h angles, d2 1 1 3:37 nm
giving acalc: 8:25 nm, and their intensities are
increased in comparison with the as-synthesized

17

intensity

x8
2

10

2 theta []

RT
550150C
C

450 550C
RT 250C C

2.0 3.0 4.0 5.0 6.0 7.0 8.0

(a)

2 theta []
4.0
3.8

intensity

(211)
(220)
(332)

3.6
3.4

d(211) [nm]

d(211)

3.2

RT 150 250 350

(b)

450 550

550

3.0
150 RT

Temperature [C]

Fig. 8. (a) XRD patterns stack plot of CTAB/MCM-48 obtained by hydrothermal synthesis with CTAB as template and TEOS as
silicon source. Shown are subsequent XRD patterns as the material is calcined up to temperature of 550 C, held at this temperature for
5 h and cooled to room temperature. Inset shows ex situ XRD pattern of as-synthesized MCM-48. (b) Graph showing the evolution of
the reection intensities of MCM-48 as a function of temperature (calcination at 550 C for 5 h). Also plotted is the evolution of
d(2 1 1) (open symbols).

CO2 m/z = 44

H2O m/z = 18

CxHy m/z = 26,42

m/z = 30
N(CH3)3m/z = 59
100
90
80

-3%

-10%
-0.2
exo

-16%

70

-0.4
-0.6

-11%
-0.8

60
-9 %
50

-1.0
-7%

325C

DTA signal [uV/mg]

3-D cubic structure during calcination. The graph


depicted in Fig. 8b shows a marked decrease in dspacing observed predominantly below 250 C
during the heating ramp. The evolution of the interplanar distance d(2 1 1) and the intensity of the
reections approximately follow a three-step process. With respect to the XRD results, one can
assume that the mechanism of the removal of the
surfactant from MCM-48 is identical to that described for CTAB/MCM-41. The intensity ratio
I(2 1 1):I(3 3 2) is shown to increase signicantly
after calcination (Table 1), and a fairly marked
lattice shrinkage occurs (13%). This may point
towards a slightly lower thermal stability of the
MCM-48 synthesized under the conditions described, compared to the Cn-MCM-41 samples.
Fig. 9 shows the TGDTA/MS measurements
performed on C16TAB/MCM-48. The results are
very similar to those of MCM-41. The stepwise
decomposition of the template occurs from 150 to
400 C. This stage corresponding to the decomposition processes of the templating species and
their thermodesorption represents about 45% in
mass loss, which is comparable to the data reported by other authors (4045%) [75,76]. The
step attributed to the Hofmann degradation of the
surfactant (150250 C) represents a weight loss of
26%, which is similar to that of the MCM-41
mesophases synthesized with short chain C12TAB
and C14TAB. Conversely, the successive oxidation
steps involve weight losses more comparable to
those of C16TAB/MCM-41. The main oxidation
peak is centered at 325 2 C. From 400 C,
changes provoked in the structure and scattering
contrast are due mostly to the removal of the residual carbonaceous species (about 7%), and the
release of a low amount of water due to condensation of silanols. The organic content is shown to
be higher than that estimated previously for the
hexagonally ordered mesostructured silica mesophases, with a total weight loss of about 56%,
consistent with a higher void volume for the 3-D
structure. One can conclude that the 3-D open
structure has no strong eects on the processes
responsible for the removal of the template by
thermal treatment. Moreover, an additional step is
observed at low temperatures around 130 C, and
likely assigned to surfactant species. The weight

Intensity of mass signal

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

Weight loss [%]

18

100 200 300 400 500 600 700 800 900

Temperature [C]
Fig. 9. TGDTA/MS measurements performed on an assynthesized MCM-48 mesophase. Bottom: TGA data with a
dashed line with its rst derivative curve (grey curve), and the
DTA curve with a solid line. Top: various molecular species
recorded from the MS with temperature.

loss measured in TG between 120 and 140 C is


about 1.62%. One may speculate that a certain
amount of surfactants molecules remains adsorbed
in layers at the external surface. These molecules
could be desorbed at low temperature as usually
observed for lamellar phase. The presence of a
mixture of cubic and lamellar silicasurfactant
phase is rather unlikely on the basis of the high
quality XRD pattern of as-synthesized samples
(Fig. 8a inset), and the physisorption data obtained after calcination (not shown). The nitrogen
sorption isotherm obtained for MCM-48 is a
typical type IV isotherm. No evidence of triangular hysteresis at high relative pressures, originating
from the presence of a pore structure with narrow

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

constrictions, could be observed [77]. Such a pore


structure is usually expected when the lamellar
phase collapses. In addition, no microporosity
could be detected from the t-plot method. All
MCM-48 samples synthesized within 35 days at
115 C showed similar features. However, contamination by a very low amount of silicasurfactant lamellar phase is still possible, but if so, not
critical for the quality of the nal material. Another explanation is the possible lower temperature thermodesorption of surfactant contained in
the pores due to high diusion properties of the
bicontinuous 3-D interconnected porous network.
The similitude of the thermal behaviors observed for MCM-41 and MCM-48 suggests that
the surfactant molecules are organized, in both
case, rather similarly within the mesopore channels, and that their respective interactions with
the silicate surface are of identical nature. In addition, Solovyov et al. revealed that a non-uniform
surfactant density is also observed for MCM-48
samples [65], conrming the resemblance with
MCM-41.
3.2. Materials synthesized under acidic conditions
(S X I )
3.2.1. SBA-3 [15]
The synthesis of SBA-3 is carried out in
strongly acidic aqueous solutions below the isoelectric point of silica. Under these conditions,
halide ions X mediate the interaction between the
surfactant and positively charged oligomeric inorganic species (S X I ) through weak hydrogen
bonding forces, which ensure the assembly of the
mesophase. Silica mesophases synthesized under
acidic conditions have dierent composition, pore
structure, wall thicknesses, and adsorption properties compared with samples obtained by alkaline
routes [15,78]. During the polymerization process,
the protons associated with the silica species are
excluded until a neutral inorganic framework remains. Hence, the framework charge is neutral or
slightly positive.
The ex situ XRD pattern shows the low angle
reections indexed to a hexagonal p6m mesophase
(see Fig. 10a inset) with d1 0 0 3:94 nm
(aas-made 4:55 nm). After calcination at 550 C

19

for 5 h, d(1 0 0) decreased to 3.18 nm (acalc: 3:67


nm), indicating a large lattice shrinkage of 19%.
Fig. 10a shows the development of the XRD
pattern during calcination of SBA-3. All the intensities of the low angle reections start increasing simultaneously at temperatures above 250 C.
From 250 C, the growth rate is high, with strong
increase in the scattering intensity. The intensities
reach their maxima at temperatures close to the
calcination plateau. During the remaining part of
the process of the calcination, no drastic changes
occur. The decrease in intensity observed during
the cooling stage is attributed to the adsorption of
water, which slightly decreases the scattering contrast between walls and pores. Another feature is
the higher I(1 0 0):I(1 1 0) intensity ratio compared
to that observed for MCM-41. Considering that
the unit cell sizes of SBA-3 and C14 or C16-MCM41 are comparable (Table 1), this diraction effect is likely caused by the dierences in wall
thickness observed when comparing both type of
materials [63,78]. In addition, a signicant increase
is clearly observed for the I(1 0 0):I(1 1 0) intensity
ratio of SBA-3 after calcination. The graph shown
in Fig. 10b illustrates the changes of the intensities
of the low angle reections combined with the
evolution of the interplanar distances d(1 0 0),
d(1 1 0) and d(2 0 0) as a function of temperature.
The d-spacings of the hexagonal mesophase decrease slightly below 150 C upon removal of
physisorbed water. The system remains then
mostly unchanged until 300 C where all intensities
increase drastically, accompanied by a marked
decrease in d-spacing. The d-spacing shift to lower
values occurs mostly between 300 and 400 C.
Following this, no further lattice shrinkage is detected, suggesting a relatively rapid process of decomposition or thermodesorption of the template
with no further structural rearrangement. The
overall unit cell contraction is about 18% and is
shown to be signicantly larger than that measured for MCM-41 or MCM-48. Moreover, the
comparison with the evolution of the (1 0 0)
reection intensity of C16-MCM-41 illustrates the
dierent nature of the decomposition process responsible for the removal of the template, and
the resulting dierences in X-ray scattering contrasts.

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

in situ
ex situ

Intensity

Relative Intensity

20

1.0

calcined
as-synthesized
3.0

5.0

7.0

9.0

2 theta []

2.1

3.1

4.1

5.1

6.1

40 550
RT 200C 0C C

RT
550 150C
C

2 theta []

(a)

4.0

Intensity

3.0
I(100)
I(110)
I(200)

(100)
(110)
(200)

2.5

d-spacing [nm]

3.5

2.0

RT150 250 350 450 550

(b)

550 150

1.5
RT

Temperature [C]

Fig. 10. (a) XRD patterns stack plot of SBA-3. Shown are subsequent XRD patterns as the material is calcined up to 550 C, held for 5
h and cooled to room temperature. Inset shows a comparison of the XRD patterns recorded before and after calcination for SBA-3
under ex situ and in situ conditions. (b) Graph showing the evolution of the reection intensities of SBA-3 as a function of temperature
(calcination at 550 C for 5 h). Also plotted are the d-spacing values of the respective reections (open symbols). Represented with a
cross-dotted line is the intensity of the (1 0 0) reection of CTAB/MCM-41 as reference.

The TGDTA/MS data (Fig. 11) show that the


template decomposition is a three-step process in a
narrow temperature range. After the desorption of
the physically adsorbed water (1%), an endothermic process is recognized between 180 and 260 C
(centered at 239 2 C), which, however, has not
been discussed before. A marked heat eect and
a large weight loss (25%) are observed. During
this step, evaporation of HCl and H2 O from the
mesophase initially takes place. This is followed by
elimination of the surfactant head group fragment
(m=z 59) and some carbon chain fragments
(m=z 42, 55). It has to be noted that about half
of the mass loss (25%) is measured during this

endothermic step, resulting in a surprisingly


modest increase in X-ray scattering contrast and
no d-spacing shift. It is followed by an exothermic
process (260300 C), corresponding to about 10%
in weight loss, likely attributed to surfactant decomposition and early oxidation step with elimination of some CO2 . The rst derivative curve
calculated for the TG prole shows one broad
peak maximum between 150 and 300 C, suggesting that the two rst steps might be somehow
temperature-dependent competing eects. Above
250 C, endothermic evaporation and thermodesorption, and oxidative decomposition could occur
simultaneously and overlap in the DTA prole.

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

CO2 m/z = 44

Intensity of mass signal

HCl m/z = 37

H2O m/z = 18
342C

Cx Hy m/z = 26,42
,
N(CH3)3 m/z = 59
-1%

90
80

exo

-25%

0
-0.2

239 C

-0.4
70

-10%

60
50

-0.6
-0.8

-11%
dm/dT
d

100

337C

- 9%

-1.0

DTA signal [uV/mg]

Weight loss [%]

100

200 300 400 500 600 700 800 900

Temperature [C]

Fig. 11. TGDTA/MS measurements performed on as-synthesized SBA-3. Bottom: TGA data with a dashed line with its
rst derivative curve (grey line), and the DTA curve with a solid
line. Top: various molecular species recorded from the MS and
their evolution with temperature.

Although the mass loss reached 35%, these combined processes do not seem to result in any
apparent change in the scattering contrasts or dspacing values. Following this, the main exothermic process with the higher energy release takes
place between 300 and 400 C (centered at 337 2
C), where the weight loss is 11%. Here, the strong
increase of all intensities at the same growth rate is
very likely caused by the rapid removal of carbonrich species from the inside of the mesopores. The
major exothermic oxidation process occurs with
release of a large amount of CO2 and smaller
fragments of the carbon chain. The oxidation is
then completed by combustion of residual carbonaceous species (coke) and water release at higher

21

temperatures. Above 400 C, the weight loss of 9%


corresponds to the removal of carbonaceous species at 450550 C and water losses via condensation of silanol groups, and, at 600 C, to high
temperature condensation of remaining silanol
groups. The total weight loss measured for SBA-3
is about 55%.
Two main processes are evidenced: (1) low
temperature endothermic removal of HCl, water
and parts of the surfactant species with no apparent structural change, and (2) oxidation and
combustion at higher temperature, leading to rapid scattering contrast variations and contraction
of the hexagonal mesophase. It is, however, questionable whether the rst process could be assigned to the Hofmann degradation observed for
Cn-MCM-41 since the in situ XRD results do not
show the same drastic contrast variation and differences in growth rate. Furthermore, the neutral
silica framework is less favorable for the basecatalyzed Hofmann elimination of the trimethylamine head group. Terminal silanol groups are
protonated so that the bulk composition of SBA-3
and MCM-41 made with the same surfactant are
distinctly dierent in hydrogen and halide ion
content, as supported by the TG/MS results. In
addition to the dierent chemical composition, one
has to consider at least two other factors that
might inuence the thermal behavior of the hexagonal mesophase synthesized under acidic conditions: (1) thicker silica walls are suggested for
SBA-3 materials compared to those of MCM-41,
(2) the possible presence of disordered micropores
in the inorganic framework walls. Albouy et al.
[79] reported recently the presence of microporosity inside the walls of mesoporous hexagonal
phase silica synthesized under acidic conditions
with CTAB as a template. In another recent report,
Lee et al. [80] emphasized the dierence in porosity
observed between SBA-3 type materials and
MCM-41. They showed that in the case of a similar
hexagonal phase mesoporous silica, synthesized
with CTAB, replication of the structure into a
CMK-type carbon materials is possible, signifying
thus the presence of complementary pores in the
silica walls. The mesoporous channels are therefore
somewhat interconnected and the resulting CMK
replica retained an ordered structure, which is in

22

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

strong contrast with replication of MCM-41 leading to nanober-like carbons. The evolution of the
scattering intensity of the low angle reections is
governed by the contrast between the walls and the
inside of the pores, whereas the TG proles depend
on the compositions and the interaction between
the matrix and the included species. The removal of
the template from as-synthesized SBA-3 is governed by the size of the surfactant species relative to
the size of the honeycomb mesopores and possible
framework micropores, and the strength of the
interaction between the template and the solid. The
presence of microporosity within the walls may
induce perturbation in the scattering contrast and a
dierent phase cancellation behavior is expected
since the scattering density of the walls is not
constant. The origin of the pore connectivities or
wall microporosity is, however, still unclear. Nevertheless, one could propose that it is related to the
inherent nature of the silicate oligomeres resulting
from hydrolysis and condensation of TEOS at very
low pH, and the extent of cross-linking and density
of the silica network formed.
To reduce the damage caused by the removal of
the template by thermal treatment and reduce the
cost of the synthesis of mesoporous materials,
non-destructive solvent extraction techniques have
been developed. For mesoporous material obtained according the S I route, the templating
species interact strongly with the inorganic
framework via charge-balancing ionic interactions.
The destruction of this kind of interaction is rather
dicult to achieve by solvent extraction alone.
However, in the acid synthesized mesophase the
surfactant cationic charge is balanced by a halide
ion, which allows the template to be removed by
solvent extraction without providing any exchangeable cations [15].
Extraction was therefore performed in pure
boiling ethanol according to the method proposed
by Tanev et al. [43]. The solvent extraction was
carried out twice, with a sample-to-extraction
media ratio of 1 g/150 ml, and subsequent washing
with ethanol. The TGDTA measurements performed on the extracted sample show a remaining
total weight loss of about 1013%, indicating a
removal of 7580% of the template by the extraction. Therefore, the removal of the residual

template species still requires subsequent calcination.


The d(1 0 0) of the hexagonal mesophase of the
extracted sample is 3.81 nm, with aextracted 4:40
nm, indicating a lattice contraction of 3%. Fig. 12a
shows the development of the XRD patterns
during calcination at 550 C for 5 h of SBA-3 after
extraction in pure boiling ethanol. The constant
intensities of the low angle reections support the
notion that a large amount of organics has been
removed from the pores. This conrms also that
the changes in intensity observed previously are
due to contrast variations. Fig. 12b shows that the
hexagonal lattice of the material after extraction
undergoes linear shrinkage from 250 C up to the
end of the calcination plateau, whereas the dspacings of the as-synthesized SBA-3 mesophase
without extraction decrease strongly in a rapid
step at 250300 C. The d(1 0 0) of the extracted
mesophase is 3.3 nm after calcination, indicating a
lower shrinkage of about 14%, compared to a nonextracted sample. The continuous large shrinkage
observed for the extracted sample suggests that the
exothermic reactions with high energy release are
not necessarily responsible for the lattice shrinkage. Nevertheless, high heat eects may enhance
this contraction eect since the eective temperature of the sample might be substantially higher
than the oven temperature. Consequently, the
shrinkage of the hexagonal lattice is likely caused
by successive condensations of the framework and
temperature-dependent structural rearrangements
of the framework.
3.2.2. SBA-15
Important progress in the preparation of mesoporous silicas was made by Zhao et al. [16], who
used triblock polyoxoalkylene copolymers (pluronic type) for the synthesis under acidic conditions of large-pore materials called SBA-15 and
other related materials. Here the mechanism of
formation is proposed to be S X I or (S0 H )
(X I ) since the block-copolymer is positively
charged under the reaction conditions. In addition,
the shape of the mesopores is apparently more
cylindrical than the pores of MCM-41-type materials that are generally considered to be hexagonal
[65].

23

Intensity

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

2.0

3.0

4.0

5.0

40 550
RT 200C 0C C

6.0

RT
550150C
C

2 theta []

(a)
4.0

d-spacing [nm]

3.5
3.0
(100)
(110)
(200)

2.5

(100)
(110)
(200)

2.0
1.5
RT 150 250

(b)

350

450

550

550

150 RT

Temperature [C]

Fig. 12. (a) XRD patterns stack plot of SBA-3 obtained after extracted in pure ethanol. Shown are subsequent XRD patterns during
the calcination. (b) d-spacings of SBA-3 during calcination of an as-synthesized sample (open symbols) and an extracted sample (solid
symbols).

SBA-15 was synthesized with a hydrothermal


treatment applied at 90 C during 24 h. As-synthesized SBA-15 exhibits reections indexed to the
p6m hexagonal symmetry at very low angles,
d1 0 0 10:49 nm indicative of a very large unit
cell with a 12:11 nm. The position of the reections of the hexagonal phase at very low angle
renders the in situ XRD measurement dicult to
realize due to the conguration of the sample
holder set up, and may induce large experimental
errors. To allow a measurement, the set up has
been changed from automatic divergence slit, as
was used so far, to a xed slit conguration.
Furthermore, due to the large signal caused by the
primary beam at very low angle, the developing

in situ XRD patterns are corrected by subtraction


of the background measured with the empty Pt/Rh
band. The XRD powder diraction pattern obtained for as-synthesized SBA-15 is shown in Fig.
13. One can note that the intensity of the (1 1 0)
diraction peak is lower than the intensity of
(2 0 0), with the I(1 1 0):I(2 0 0) intensity ratio being
about 0.8 (Fig. 13 inset). Fig. 13 illustrates the
development of the XRD patterns during calcination. Such in situ experiments are presented for
the rst time in the case of a SBA-15 mesophase.
The in situ XRD patterns stack plot reveals very
interesting and new features compared to the previously described silica-based materials. A rapid
strong increase of all reection intensities is

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

Intensity

24

x 8
calcined
x 8
as-made
2.0

2 theta []

3.0

Intensity

1.0

45
150
550
550
550
500
400
300
200
RT

1.7

2.7

2 theta []
Fig. 13. XRD patterns stack plot obtained for SBA-15. Shown are subsequent XRD patterns as the material is calcined up to 550 C,
held at this temperature for 5 h and cooled to room temperature. Inset shows in detail the XRD patterns of SBA-15 before and after
calcination, respectively.

observed rst up to 200 C. Surprisingly, upon


further increase of the temperature up to 250 C, a
relative decrease in scattering contrast is then observed. Following this, continuous increase in intensity is occurring at a slow growth rate, the
intensities reaching a maximum around 550 C.
This slow linear increase in intensity at higher
temperatures is accompanied with an observable
d-spacing shift. After calcination, the I(1 1 0):
I(2 0 0) intensity ratio increases to about 1.4, which
is likely a result of increased framework wall
condensation [69]. On the other hand, the I(1 0 0):
I(1 1 0) intensity ratio remains constant.
The graphs depicted in Fig. 14 show the detailed
evolution of the intensities of the low angle reections during the calcination process, and the associated changes in d-spacings. From the graphs, it is
clearly seen that all reections follow an identical
evolution. A rst increase is observed up to 200 C,
with no apparent shrinkage of the hexagonal lattice. From 200 to 250 C, a substantial decrease in
intensity occurs without structural change in lattice

size. Above 250 C, all the d-spacings shift linearly


towards lower values, up to at 550 C. In parallel to
the lattice shrinkage, a progressive increase in
scattering contrast is observed up to 550 C. From
this, changes are then less pronounced, the reection intensities and d-spacings remaining relatively
constant. Upon cooling to room temperature, water
is physically adsorbed resulting in a slight lattice
expansion, as observed previously for other silicabased mesostructures, and an observable new increase in the reection intensities. The unit cell
constant of calcined SBA-15 is 10.52 nm, indicating
a lattice shrinkage of about 13%.
From the TGDTA/MS (Fig. 15) one can
conclude that the template removal from an assynthesized SBA-15 sample consists of a single
main step occurring at relatively low temperatures
(below 280 C), followed by higher temperature
elimination of residual carbonaceous species. After
the removal of physisorbed water (4%) below 150
C in an endothermic process, the exothermic decomposition of the organic template takes place in

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

25

11.0

Intensity

10.5

9.5
9.0

RT 150

250

350

(a)

450

550

550

8.5
150 RT

Temperature [C]

Intensity of mass signal

10.0
I(100)
d(100)

CO2 m/z = 44

C3H6O2 m/z = 74
C3H6O m/z = 58
HCl m/z = 37

6.0

I(110)
I(200)

250

350

450

550

d(110)
d(200)

550

5.0

4.5
150 RT

Weight loss [%]

100

RT 150

(b)

H2O m/z = 18

5.5
-4%

0.2
exo

90
80

-0.2
-42%

-0.4

70

-0.6

60

-0.8

Temperature [C]
50

Fig. 14. Graph showing the evolution of the reection intensities of SBA-15 as a function of temperature (calcination at 550
C for 5 h). Also plotted are the d-spacing values of the respective reections (open symbols). Top: evolution of the (1 0 0)
reection. Bottom: evolution of the (1 1 0) and (2 0 0) reections. The connecting black solid lines are used as guide for the
eye.

one step between 150 and 280 C with a weight loss


of 42%. During this exothermic step (DTA peak
centered at 169 2 C), all fragments assigned to
block-copolymer fragments (m=z 42, 58, 74),
HCl (m=z 37) and water are detected simultaneously in the MS. The marked exothermic eect
and important weight loss below 200 C (32%) are
to be associated with the strong increase in scattering contrast observed in the in situ XRD experiments. Furthermore, one can note that all MS
peaks are relatively narrow between 150 and 200
C. However, a broadening of the peaks attributed
to CO2 and water is clearly seen between 200 and
280 C. During this temperature range, a weight
loss of about 10% is measured. This observable
delayed removal of CO2 and water is related to the

-1.0
-10%

169C

DTA signal [uV/mg]

Intensity

317C

-1.2

100 200 300 400 500 600 700 800 900

Temperature [C]
Fig. 15. TGDTA/MS measurements performed on as-made
P123/SBA-15. Bottom: TGA data with a dashed line and the
DTA curve with a solid line. Top: various molecular species
recorded from the MS measurements and their evolution with
temperature.

range of temperature where a decrease in intensity


is detected. Subsequently, between 280 and 400 C,
another weight loss of 10% is measured, with an
exothermic sharp peak maximum at 317 2 C,
assigned to the removal of water and residual
carbonaceous species (m=z 18 and 44). The total
weight loss measured is about 56%, as reported
before [16,30,47].
The rst major step is the complete decomposition and combined combustion of the organics.
Large amounts of organic template are removed
from the inside of the pores leading to a strong
increase in scattering contrast between the walls
and the mesopores. During this step, part of the
organic template is converted into carbonaceous

26

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

species. The temperature of this decomposition


process is lower than the temperature at which the
pure P123 decomposes (about 210 C) [30], and
substantially lower than the oxidative decomposition temperature observed for the cationic surfactants. The delayed removal of CO2 and water
might be caused by the decomposition or oxidation of species located in the complementary pores
known to be present in SBA-15 framework walls.
The behavior of the scattering contrast is strongly
dependent on the structure of the walls and the size
of the pores. In other terms, the scattering power
depends on the electron density contrast between
the dierent moieties of the unit cell. It has been
recently demonstrated that the large structural
mesopores of SBA-15 are accompanied by disordered micropores or mesopores located within the
silica pore walls, likely providing connectivity between the ordered large-pore channels [48,8184].
Interestingly, Galarneau et al. [85] showed precisely in a very recent report that the nature of the
complementary porosity is strongly depending on
the synthesis conditions. They elucidated the porous topology of SBA-15 samples synthesized at
dierent temperatures, using TEM images of Pt
replicas and low pressure Ar adsorption. They
concluded the presence of a secondary porosity
with pore diameter larger than 1.5 nm bridging
the structural mesopore channels only in the
case of materials synthesized at temperatures
>80 C. Below this temperature, ultramicroporosity (<1 nm) is usually observed as a corona
around the mesopores, with no clear evidence of
porous bridges connecting the channels. In our
case (90 C), the presence of such complementary
bridging pores in the walls may inuence the
scattering density of the framework. As the distribution of matter in the pores in the walls and the
mesopores is temperature dependent, unexpected
contrast behavior might take place with possible
phase cancellation. Consequently, the change in
scattering contrast occurring when CO2 and water
are removed does not seem to be caused by the
elimination of species located within the large pore
channels, but more likely to species that were
found in the bridging complementary pores of the
framework walls. The subsequent broader step
corresponds to the removal of coke by combus-

tion, and to the dehydroxylation of silanols. In this


second step, a peak appears at about 315 C where
more water is released, with traces of HCl. Of note
is that the condensation of silanol to form siloxane
has been generally suggested to occur at higher
temperatures in the case of mesophases synthesized with cationic surfactants (400600 C)
[22,71]. During this broad step, the scattering intensities reach progressively their maximal values.
The nal scattering increase observed is probably
caused by re-adsorption of water, which might
undergo preferential condensation in the smaller
pores (micropores and small mesopores) located in
the thick walls of SBA-15.
The removal of the template is shown to be very
dierent from that of MCM-41 or SBA-3 synthesized with low molecular weight surfactants.
Most of high molecular weight non-ionic template is removed from mesostructured SBA-15
at lower temperatures, between 150 and 250 C,
in a single oxidation step. No contraction of the
unit cell is observed during this process. Solovyov
et al. [65] showed recently that the block-copolymer surfactant is distributed more uniformly
within the large cylindrical mesopores of SBA-15,
than CTAB in MCM-41 or MCM-48. These observations suggest that the block-copolymer template is likely involved in a dierent type of
interaction with the silica surface. The inorganic
framework seems to catalyze the thermal decomposition and oxidation of the block copolymer at
low temperatures in the presence of oxygen, since
the decomposition is strongly delayed under nitrogen [47]. Residual carbonaceous species and
water are then removed from the structure upon
heating at higher temperatures, from 300 C up to
550 C. During this subsequent process a large
contraction of the hexagonal unit cell is observed, possibly due to further framework condensation. The presence of additional framework
porosity or connectivities between the mesopore
channels seems to be responsible for the strong
scattering contrast variations observed during the
thermal treatment. In addition, dierences in surface roughness have to be taken into account
when comparing SBA-15 with the other types
of mesoporous materials described in this study
[63].

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

27

4. Conclusions

Acknowledgements

The XRD studies showed the changes in scattering contrast, observed for the low angle reections, occurring when the template is removed.
Dierences in scattering contrast variations and
chemical reactions involved are observed for mesoporous silicas depending on the synthesis conditions
and type of surfactant, which highlight the role of
the silicasurfactant interfaces. For all samples, a
strong increase in scattering contrast is evidenced,
thus resulting in increasing reection intensities
upon removal of the templating agent. The removal
of the surfactant, in the case of Si-MCM-41 or
Si-MCM-48, occurs by a stepwise mechanism. The
rst step of the template decomposition via Hofmann degradation is conrmed for all MCM-41
samples synthesized with n-alkyltrimethylammonium surfactants and MCM-48, with however, different proportions of the organics involved in
the temperature-dependent processes. The use of
surfactants with dierent chain lengths underlines
the eects of the surfactant-surface interactions
and to a lower extent, the probable role of the pore
size on the thermal desorption of the decomposed
organics. Possible mass transfer limitations for
the diusion of larger hydrocarbon species may be
suggested. The temperature at which the dierent
alkylammonium surfactants are removed may serve
to probe the strength of the interactions considered.
Furthermore, the exchange of the surfactant trimethylammonium head group for a pyridinium group
stresses the determining inuence of the interactions
of the polar head group and the inorganic surface
during thermal treatment. Materials synthesized
following the acidic route show dierent behaviors
depending on the type of template employed. Despite having thicker walls, the materials obtained via
the acidic synthesis route show the highest lattice
shrinkage, which may be related to the nature of the
framework walls. Moreover, the presence of complementary wall microporosity or porous bridges
connecting the mesopore channels seems to greatly
inuence the XRD scattering contrast behavior
and the processes of the thermal desorption of the
organics. Finally, the SBA-15 framework catalyzes
the oxidation of the block copolymer template species at low temperatures.

The European Community (project HPRN-CT99-00025) is gratefully acknowledged for nancial


supports. F.K. wishes to thank Dr. C. Weidenthaler, Dr. F. Marlow (MPI M
ulheim), and Dr. M.
bo Akademi, Finland) for helpful and
Linden (A
stimulating discussions.

References
[1] C.T. Kresge, M.E. Leonowicz, W.J. Roth, J.C. Vartuli, J.S.
Beck, Nature 359 (1992) 710.
[2] T. Yanagisawa, T. Shimizu, K. Kuroda, C. Kato, Bull.
Chem. Soc. Jpn. 63 (1990) 988.
[3] A. Corma, Chem. Rev. 97 (1997) 2373.
[4] M. Linden, S. Schacht, F. Sch
uth, A. Steel, K.K. Unger,
J. Porous Mater. 5 (1998) 177.
[5] U. Ciesla, F. Sch
uth, Micropor. Mesopor. Mater. 27 (1999)
131.
[6] J.Y. Ying, C.P. Mehnert, M.S. Wong, Angew. Chem. Int.
Ed. Engl. 38 (1999) 56.
[7] F. Sch
uth, A. Wingen, J. Sauer, Micropor. Mesopor.
Mater. 4445 (2001) 465.
[8] K.K. Unger, D. Kumar, M. Gr
un, G. B
uchel, S. L
udtke,
Th. Adam, K. Schumacher, S. Renker, J. Chromatogr. A
892 (2000) 47.
[9] M. Vallet-Regi, A. Ramila, R.P. del Real, J. PerezPariente, Chem. Mater. 13 (2001) 308.
[10] R.C. Hayward, P. Alberius-Henning, B.F. Chmelka,
G.D. Stucky, Micropor. Mesopor. Mater. 4445 (2001)
619.
[11] F. Sch
uth, W. Schmidt, Adv. Mater. 14 (2002) 629.
[12] G. Wirnsberger, P. Yang, B.J. Scott, B.F. Chmelka, G.D.
Stucky, Spectrochim. Acta A 57 (2001) 2049.
[13] B.J. Scott, G. Wirnsberger, G.D. Stucky, Chem. Mater. 13
(2001) 3140.
[14] J.S. Beck, J.C. Vartuli, W.J. Roth, M.E. Leonowicz, C.T.
Kresge, K.D. Schmitt, C.T.-W. Chu, D.H. Olson, E.W.
Sheppard, S.B. McCullen, J.B. Higgins, J.L. Schlenker, J.
Am. Chem. Soc. 114 (1992) 10834.
[15] Q. Huo, D.I. Margolese, U. Ciesla, D.G. Demuth, P.
Feng, T.E. Gier, P. Sieger, A. Firouzi, B.F. Chmelka, F.
Sch
uth, G.D. Stucky, Chem. Mater. 6 (1994) 1176.
[16] D. Zhao, J. Feng, Q. Huo, N. Melosh, G.H. Frederickson,
B.F. Chmelka, G.D. Stucky, Science 279 (1998) 548.
[17] J.C. Vartuli, K.D. Schmitt, C.T. Kresge, W.J. Roth, M.E.
Leonowicz, S.B. McCullen, S.D. Hellring, J.S. Beck, J.L.
Schlenker, D.H. Olson, E.W. Sheppard, Chem. Mater. 6
(1994) 2317.
[18] N.K. Raman, M.T. Anderson, C.J. Brinker, Chem. Mater.
8 (1996) 1682.
[19] F. Kleitz, W. Schmidt, F. Sch
uth, Micropor. Mesopor.
Mater. 4445 (2001) 95.

28

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129

[20] J. Blanchard, P. Trens, M. Hudson, F. Sch


uth, Micropor.
Mesopor. Mater. 39 (2000) 163.
[21] F. Sch
uth, U. Ciesla, S. Schacht, M. Thieme, Q. Huo, G.
Stucky, Mater. Res. Bull. 34 (1999) 483.
[22] C.-Y. Chen, H.-X. Li, M.E. Davis, Micropor. Mater. 2
(1993) 17.
[23] M. Soulard, S. Bilger, H. Kessler, J.L. Guth, Zeolites 11
(1991) 107.
[24] A. Fonseca, J.B. Nagy, J. El Hage-Al Asswad, R.
Mostowicz, F. Crea, F. Testa, Zeolites 15 (1995) 259.
[25] L.M. Parker, D.M. Bibby, J.E. Patterson, Zeolites 4 (1984)
168.
[26] O. Kresnawahjuesa, D.H. Olson, R.J. Gorte, G.H. K
uhl,
Micropor. Mesopor. Mater. 51 (2002) 175.
[27] K.O. Kongshaug, H. Fjellv
ag, B. Klewe, K.P. Lillerud,
Micropor. Mesopor. Mater. 39 (2000) 333.
[28] A. Corma, V. Fornes, M.T. Navarro, J. Perez-Pariente, J.
Catal. 148 (1994) 569.
[29] M.T.J. Keene, R.D.M. Gougeon, R. Denoyel, R.H.
Harris, J. Rouquerol, P.L. Llewellyn, J. Mater. Chem. 9
(1999) 2843.
[30] D. Zhao, Q. Huo, J. Feng, B.F. Chmelka, G.D. Stucky, J.
Am. Chem. Soc. 120 (1998) 6024.
[31] F. Sch
uth, Chem. Mater. 13 (2001) 3184.
[32] U. Ciesla, M. Fr
oba, G.D. Stucky, F. Sch
uth, Chem.
Mater. 11 (1999) 227.
[33] D. Trong On, Langmuir 15 (1999) 8561.
[34] D. Khushalani, G.A. Ozin, A. Kuperman, J. Mater. Chem.
9 (1999) 1491.
[35] F. Kleitz, S.J. Thomson, Z. Liu, O. Terasaki, F. Sch
uth,
Chem. Mater. 14 (2002) 4134.
[36] C. Serre, M. Hervieu, C. Magnier, F. Taulelle, G. Ferey,
Chem. Mater. 14 (2002) 180.
[37] G.J. de A.A. Solar-Illia, A. Louis, C. Sanchez, Chem.
Mater. 14 (2002) 750.
[38] P. Yang, D. Zhao, D.I. Margolese, B.F. Chmelka, G.D.
Stucky, Chem. Mater. 11 (1999) 2813.
[39] P.T. Tanev, T.J. Pinnavaia, Science 267 (1995) 865.
[40] S. Hitz, R. Prins, J. Catal. 168 (1997) 194.
[41] F. Sch
uth, Ber. Bunsen.-Ges. Phys. Chem. 99 (1995) 1306.
[42] S. Kawi, M.W. Lai, Chem. Commun. (1998) 1407.
[43] P.T. Tanev, T.J. Pinnavaia, Chem. Mater. 8 (1996) 2068.
[44] D.M. Antonelli, J.Y. Ying, Angew. Chem. Int. Ed. Engl.
35 (1996) 426.
[45] D.M. Antonelli, A. Nakahira, J.Y. Ying, Inorg. Chem. 35
(1996) 3126.
[46] D.M. Antonelli, J.Y. Ying, Chem. Mater. 8 (1996) 874.
[47] M. Kruk, M. Jaroniec, C.H. Ko, R. Ryoo, Chem. Mater.
12 (2000) 1961.
[48] Z. Liu, O. Terasaki, T. Ohsuna, K. Hiraga, H.J. Shin, R.
Ryoo, Chem. Phys. Chem. 4 (2001) 229.
[49] M.T.J. Keene, R. Denoyel, P.L. Llewellyn, Chem. Commun. (1998) 2203.
[50] G. B
uchel, R. Denoyel, P.L. Llewellyn, J. Rouquerol, J.
Mater. Chem. 11 (2001) 589.
[51] T. Clarck Jr., J.D. Ruiz, H. Fan, C.J. Brinker, B.I.
Swanson, A.N. Parikh, Chem. Mater. 12 (2000) 3879.

[52] G. B
uchel, P. Llewelyn, Private communication.
[53] M. Gr
un, I. Lauer, K.K. Unger, Adv. Mater. 9 (1997) 254.
[54] M. Gr
un, K.K. Unger, A. Matsumoto, K. Tsutsumi,
Micropor. Mesopor. Mater. 27 (1999) 207.
[55] M. Fr
oba, R. K
ohn, G. Bouaud, O. Richard, G. van
Tendeloo, Chem. Mater. 11 (1999) 2858.
[56] A. Hahn, T. Ressler, R.E. Jentoft, F.C. Jentoft, Chem.
Commun. (2001) 537.
[57] F. Kleitz, Ph.D. Thesis, Ruhr-Universitat, Bochum.
[58] F. Kleitz, W. Schmidt, in: E. Kapsch, M. Hollering (Eds.),
Gekoppelte Techniken in der Thermischen Analyse, SKT,
2001, pp. 159168.
[59] R. Zana, J. Frasch, M. Soulard, B. Lebeau, J. Patarin,
Langmuir 15 (1999) 2603.
[60] B. Marler, U. Oberhagemann, S. Vortmann, H. Gies,
Micropor. Mater. 6 (1996) 375.
[61] W. Hammond, E. Prouzet, S.D. Manhati, T.J. Pinnavaia,
Micropor. Mesopor. Mater. 27 (1999) 19.
[62] Z. Tun, P.C. Mason, Acta Cryst. A 56 (2000) 536.
[63] J. Sauer, F. Marlow, F. Sch
uth, Phys. Chem. Chem. Phys.
3 (2001) 1.
[64] L.A. Solovyov, S.D. Kirik, A.N. Shmakov, V.N. Romannikov, Micropor. Mesopor. Mater. 4445 (2001) 17.
[65] L.A. Solovyov, O.V. Belousov, A.N. Shmakov, V.I.
Zaikovskii, S.H. Joo, R. Ryoo, E. Haddad, A. Gedeon,
S.D. Kirik, in: Proceedings of IMMS2002, Stud. Surf. Sci.
Catal. 146 (2003) 299.
[66] L. Chen, T. Horiuchi, T. Mori, K. Maeda, J. Phys. Chem.
B 103 (1999) 1216.
[67] S. Biz, M.G. White, Micropor. Mesopor. Mater. 40 (2000)
159.
[68] Q. Huo, D.I. Margolese, G.D. Stucky, Chem. Mater. 8
(1996) 1147.
[69] M. Linden, J. Blanchard, S. Schacht, S. Schunk, F. Sch
uth,
Chem. Mater. 11 (1999) 3002.
[70] M. Kruk, M. Jaroniec, Y. Sakamoto, O. Terasaki, R.
Ryoo, C.H. Ko, J. Phys. Chem. B 104 (2000) 292.
[71] X.S. Zhao, C.Q. Lu, A.K. Whittaker, G.J. Millar, H.Y.
Zhu, J. Phys. Chem. B 101 (1997) 6525.
[72] M. Jaroniec, M. Kruk, H.J. Shin, R. Ryoo, Y. Sakamoto, O. Terasaki, Micropor. Mesopor. Mater. 48 (2001)
127.
[73] D. Khushalani, A. Kuperman, N. Coombs, G.A. Ozin,
Chem. Mater. 8 (1996) 2188.
[74] M. Morey, A. Davidson, G.D. Stucky, Micropor. Mater. 6
(1996) 99.
[75] M. Kruk, M. Jaroniec, R. Ryoo, S.H. Joo, Chem. Mater.
12 (2000) 1414.
[76] A.A. Romero, M.D. Alba, J. Klinowski, J. Phys. Chem. B
102 (1998) 123.
[77] M. Kruk, M. Jaroniec, M.L. Pena, F. Rey, Chem. Mater.
14 (2002) 4434.
[78] S. Schacht, Q. Huo, I.G. Voigt-Martin, G.D. Stucky, F.
Sch
uth, Science 273 (1996) 768.
[79] P.-A. Albouy, A. Ayral, Chem. Mater. 14 (2002) 3391.
[80] J.-S. Lee, S.H. Joo, R. Ryoo, J. Am. Chem. Soc. 124 (2002)
1156.

F. Kleitz et al. / Microporous and Mesoporous Materials 65 (2003) 129


[81] C.H. Ko, R. Ryoo, M. Kruk, V. Antochshuk, M. Jaroniec,
J. Phys. Chem. B 104 (2000) 11465.
[82] M. Imperior-Clerc, P. Davidson, A. Davidson, J. Am.
Chem. Soc. 122 (2000) 11925.
[83] P.I. Ravikovitch, A.V. Neimark, J. Phys. Chem. B 105
(2001) 6817.

29

[84] S.H. Joo, R. Ryoo, M. Kruk, M. Jaroniec, J. Phys. Chem.


B 106 (2002) 4640.
[85] A. Galarneau, H. Cambon, F. DiRenzo, R. Ryoo, M.
Choi, F. Fajula, New J. Chem. 27 (2003) 73.
[86] M. Kruk, M. Jaroniec, A. Sayari, Chem. Mater. 11 (1999)
492.

You might also like