130
Jonathan De Vydt catalysts of the next generation of cooperative heterogeneous Green aldol condensation: synthesis, testing and design Academic year 2014-2015 Faculty of Engineering and Architecture Chairman: Prof. dr. Isabel Van Driessche Vakgroep Anorganische en Fysische Chemie Chairman: Prof. dr. ir. Guy Marin Department of Chemical Engineering and Technical Chemistry Master of Science in Chemical Engineering Master's dissertation submitted in order to obtain the academic degree of Counsellor: Ir. Jeroen Lauwaert Supervisors: Prof. dr. ir. Joris Thybaut, Prof. Pascal Van Der Voort

Green aldol condensation: synthesis, testing and design of

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Green aldol condensation: synthesis, testing and design of

Jonathan De Vydt

catalystsof the next generation of cooperative heterogeneousGreen aldol condensation: synthesis, testing and design

Academic year 2014-2015Faculty of Engineering and Architecture

Chairman: Prof. dr. Isabel Van DriesscheVakgroep Anorganische en Fysische Chemie

Chairman: Prof. dr. ir. Guy MarinDepartment of Chemical Engineering and Technical Chemistry

Master of Science in Chemical EngineeringMaster's dissertation submitted in order to obtain the academic degree of

Counsellor: Ir. Jeroen LauwaertSupervisors: Prof. dr. ir. Joris Thybaut, Prof. Pascal Van Der Voort

Page 2: Green aldol condensation: synthesis, testing and design of

..

Page 3: Green aldol condensation: synthesis, testing and design of

Jonathan De Vydt

catalystsof the next generation of cooperative heterogeneousGreen aldol condensation: synthesis, testing and design

Academic year 2014-2015Faculty of Engineering and Architecture

Chairman: Prof. dr. Isabel Van DriesscheVakgroep Anorganische en Fysische Chemie

Chairman: Prof. dr. ir. Guy MarinDepartment of Chemical Engineering and Technical Chemistry

Master of Science in Chemical EngineeringMaster's dissertation submitted in order to obtain the academic degree of

Counsellor: Ir. Jeroen LauwaertSupervisors: Prof. dr. ir. Joris Thybaut, Prof. Pascal Van Der Voort

Page 4: Green aldol condensation: synthesis, testing and design of

Acknowledgement Nu deze opleiding op zijn einde loopt, wil ik graag even de tijd nemen om enkele mensen te

bedanken die me geholpen hebben tijdens deze periode. Deze thesis was een ware uitdaging.

Voornamelijk de vele arbeidsintensieve momenten in het labo, werden in het tweede semester

steeds langer. Maar wat eens zo veraf bleek, is nu binnen handbereik. Het studentenleven kan

met een goed, toch wel trots, gevoel afgesloten worden.

In de eerste plaats bedank ik prof. dr. ir. Guy B. Marin voor de mogelijkheid die ik heb gekregen

om mijn masterproef op het Laboratory for Chemical Technology uit te voeren.

Mijn promotor prof. dr. ir. Joris W. Thybaut wil ik bedanken voor de raad en de opvolging tijdens

deze thesis. Het was verrijkend om tijdelijk deel uit te maken van de CaRE-groep. Via de CaRE

lunch meetings kwam ik in contact met de ruime wereld van Catalytic Reaction Engineering. Een

interessante ervaring!

Daarnaast wens ik ook prof. dr. Pascal Van Der Voort te danken voor de gekregen kans om me

verder te verdiepen in het onderzoeksdomein van periodieke mesoporeuze organosilica’s.

Mijn begeleider, Jeroen Lauwaert, verdient een heel grote “dankuwel”. Hij was steeds

beschikbaar indien ik vragen had. Ook als er problemen waren in het labo, stond hij steeds

paraat. Onder zijn begeleiding heb ik veel bijgeleerd. Tevens wil ik hem bedanken voor het

geduld bij het nalezen van dit werk.

Er zijn nog zovelen die direct of indirect betrokken zijn bij de ontwikkeling van deze thesis.

Zo ook een woord van dank aan Els De Canck, Judith Ouwehand en Dolores Esquivel voor hun

hulp tijdens de functionalisatie en karakterisatie van de gesynthetiseerde materialen. Tom

Planckaert wil ik bedanken voor de vele elementaire analyses en XRD metingen. Ook de

technische staf van het LCT, in het bijzonder Erwin Turtelboom, bedank ik voor het in orde

brengen van de waterkoeling in het labo.

Ik wil ook mijn klasgenoten, en dan speciaal de mensen uit het thesislokaal, bedanken voor de

vele verhalen en discussies. Het was fijn deze momenten met hen te kunnen delen.

Ten slotte gaat een oprecht woord van dank uit naar mijn ouders voor de onophoudelijke steun

die ze geven. Dat ik deze masterproef en de daarbijhorende opleiding tot een goed einde heb

kunnen brengen, is zeker ook aan hen te danken!

Page 5: Green aldol condensation: synthesis, testing and design of

FACULTY OF ENGINEERING AND ARCHITECTURE

Department of Chemical Engineering and Technical Chemistry

Laboratory for Chemical Technology

Director: Prof. Dr. Ir. Guy B. Marin

Laboratory for Chemical Technology • Technologiepark 914, B-9052 Gent • www.lct.ugent.be

Secretariat : T +32 (0)9 33 11 756 • F +32 (0)9 33 11 759 • [email protected]

Laboratory for Chemical Technology

Declaration concerning the accessibility of the master thesis Undersigned, Jonathan De Vydt

Graduated from Ghent University, academic year 2014-2015 and is author of the

master thesis with title:

Green aldol condensation: synthesis, testing and design of the next generation of

cooperative heterogeneous catalysts

The author gives permission to make this master dissertation available for consultation

and to copy parts of this master dissertation for personal use.

In the case of any other use, the copyright terms have to be respected, in particular with

regard to the obligation to state expressly the source when quoting results from this master

dissertation.

(Date) 26/05/15

(Signature)

Page 6: Green aldol condensation: synthesis, testing and design of

Summary Green aldol condensation: synthesis, testing and design of the next generation of cooperative

heterogeneous catalysts

Author: Jonathan De Vydt

Supervisors: Prof. dr. ir. Joris Thybaut, Prof. Pascal Van Der Voort

Coach: Ir. Jeroen Lauwaert

Faculty of Engineering and Architecture

Academic year 2014-2015

Abstract: In this work, the synthesis of a new type of Periodic Mesoporous Organosilica (PMO)

materials is described. These structured organosilicas have been synthesized by the acid-

catalyzed hydrolysis and condensation of bridged precursors. These materials are functionalized

with cysteine using a thiol-ene click reaction. Afterwards, the cooperative catalysts are tested in

the aldol condensation of 4-nitrobenzaldehyde with acetone. It was found that the materials

with ethane bridges are more active than the ones with benzene groups in the support. In the

second section of this work, the effect of silanol groups on the hydrophobicity of the material is

tuned by adding tetraethyl orthosilicate (TEOS) to the precursor mixture. Materials with both

hydrophobic and hydrophilic blocks are obtained. The effect of the support’s hydrophobicity is

assessed by determining the catalyst activity in the presence of water. The outcome of these

experiments is that the TE-materials have promising results. It is an interesting result to see

that, even in the presence of water, the most hydrophobic material (TE1-C) is able to preserve

its activity. This indicates that the hydrophobic pore walls impede the water molecules to enter

the pores.

Keywords: aldol condensation, cysteine, furfural, thiol-ene click reaction, effect of hydrophobicity

Page 7: Green aldol condensation: synthesis, testing and design of
Page 8: Green aldol condensation: synthesis, testing and design of

Green aldol condensation: synthesis, testing and

design of the next generation of cooperative

heterogeneous catalysts

Jonathan De Vydt

Coach: Ir. Jeroen Lauwaert

Supervisors: Prof. dr. Pascal Van Der Voort, Prof. dr. ir. Joris W. Thybaut

Abstract: In this work, the synthesis of a new type of Periodic

Mesoporous Organosilica (PMO) materials is described. These

structured organosilicas have been synthesized by the acid-

catalyzed hydrolysis and condensation of bridged precursors.

These materials are functionalized with cysteine using a thiol-ene

click reaction. Afterwards, the cooperative catalysts are tested in

the aldol condensation of 4-nitrobenzaldehyde with acetone. It

was found that the materials with ethane bridges are more active

than the ones with benzene groups in the support. In the second

section of this work, the effect of silanol groups on the

hydrophobicity of the material is tuned by adding tetraethyl

orthosilicate (TEOS) to the precursor mixture. Materials with

both hydrophobic and hydrophilic blocks are obtained. The

effect of the support’s hydrophobicity is assessed by determining

the catalyst activity in the presence of water. The outcome of

these experiments is that the TE-materials have promising

results. It is an interesting result to see that, even in the presence

of water, the most hydrophobic material (TE1-C) is able to

preserve its activity. This indicates that the hydrophobic pore

walls impede the water molecules to enter the pores.

Keywords: aldol condensation, cysteine, furfural, thiol-ene click

reaction, effect of hydrophobicity

I. INTRODUCTION

The aldol condensation is an important reaction to create

new C-C bonds and yield larger and more complex molecules.

The aldol condensation might have a bright future in the

discipline of green, renewable and sustainable energy, e.g. in

the valorization of biocomponents such as glycerol and

furfural.

Figure 1: Aqueous-phase conversion of sugars and derivatives into

liquid hydrocarbon fuels [1]

Even with the shale gas boom in recent years, one cannot

deny that the diminishing reserves of these non-renewable

resources cause a lot of concerns. In this regard, biomass can

serve as a sustainable source for our industrialized society.

Furfural is an important intermediate during the conversion of

biomass and can be used as reactant in the aldol condensation

with acetone, see Figure 1, to produce C8 and C13 alkanes

which can be employed as biofuel [1, 2].

In industry, the product stream of furfural contains water.

Therefore for an industrial application of the aldol

condensation with furfural, it would be beneficial for the

feasibility if the reaction can proceed at a significant rate even

in the presence of water.

In this regard, Periodic Mesoporous Organosilicas can be a

good solution. Due to their large pore sizes and surface areas,

these mesoporous silica materials provide sufficient space to

incorporate multiple functional groups while the stability of

the matrix is ideal for many types of reaction. Thus, if one is

able to synthesize a material with hydrophobic pores, the

water could be excluded from entering, leading to a situation

in which only the reactants are able to physisorb in the pores.

II. PROCEDURES

A. Catalyst synthesis

An overview of the synthesis and functionalization of the

Periodic Mesoporous Organosilica (PMO) materials is shown

in Figure 2.

The co-condensation of 1,2-bis(triethoxysilyl)ethane

(BTEE) and 1,4-bis(triethoxysilyl)benzene (BTEB) with

vinyltriethoxysilane (VTES) is performed in the presence of a

structure directing agent, Brij 76. The synthesis mixture is

stirred at 50°C during 24 hours and subsequently aged at 90°C

for 24 hours. After filtration of the precipitate, the material is

refluxed three times with acidified ethanol at 80°C during 24

hours to remove the surfactant as much as possible.

B. Functionalization

After the synthesis of these PMO materials, the amino acid

is introduced into the pores via a post-functionalization. In a

UV reactor, cysteine is grafted onto the vinyl group of the

VTES precursor via a thiol-ene click reaction.

First 0.50 g photo-initiator (Irgacure 2959) is dissolved in

10 mL demineralized water and the long vertical Schlenk

flask is placed into a supersonic bath. Meanwhile 0.53 g

cysteine is added to a separate glass bottle with 10 mL

demineralized water. The mixture is heated using a heat gun

Page 9: Green aldol condensation: synthesis, testing and design of

until a clear solution is obtained. Next, the cysteine and 0.50 g

of the organosilica are brought together in the Schlenk flask.

The stopcock is used to place the mixture under an inert

atmosphere with helium. After that, the mixture is placed in

the UV reactor and stirred during 24 hours for the thiol-ene

click reaction to take place.

Figure 2: Synthesis and functionalization of Periodic Mesoporous

Organosilicas

C. Catalyst characterization

Nitrogen adsorption-desorption measurements are carried

out at 77K using a Tristar 3000 gas analyser of Micromeritics.

The specific surface area and pore volume are determined

using the Brunauer–Emmett–Teller (BET) method. The

average pore size of the organosilica is obtained using the

Broekhoff and de Boer (BdB-FHH) method, with the

modification of Frenkel, Halsey and Hill.

The amine loading is determined using elemental (CHNS)

analysis. These experiments are performed on a Thermo Flash

2000 elemental analyser using V2O5 as catalyst.

The structure of the mesoporous materials is determined

with the use of X-ray diffraction. The X-ray diffraction

patterns were recorded on a ARL X'TRA Diffractometer of

Thermo Scientific which is equipped with a Cu Kα tube and a

Peltier cooled lithium drifted silicon solid stage detector.

The presence of amine groups after grafting is demonstrated

by means of Diffuse Reflectance Infrared Fourier Transform

(DRIFT) spectroscopy. These measurements are performed on

a Nicolet 6700 of Thermo Scientific with a nitrogen cooled

MCT-A detector. The spectra are obtained using a Graseby

Specac diffuse reflective cell, operating in vacuo at 120°C.

D. Catalytic experiments

The experiments were performed in a 25 mL two-neck

round-bottom flask equipped with a condenser and a septum.

The reaction mixture was prepared separately by mixing the

desired amounts of acetone (50 vol%), n-hexane (co-solvent,

50 vol%), 4-nitrobenzaldehyde (0.03 mmol/mL) and methyl

4-nitrobenzoate (internal standard, 0.022 mmol/mL).

Afterwards 8 mL of the reaction mixture was injected into the

flask which contains the catalyst (4 mol% with respect to the

4-nitrobenzaldehyde concentration) and the flask was

immediately placed in an oil bath at 45°C. The moment the

flask was placed in the oil bath was taken as the start of the

reaction. The reaction was monitored for 4 hours by taking

a sample of the reaction mixture (about 100 µL) every 30

minutes. Typically 0.9 mL of acetone was used to wash the

syringe needle and to transfer the sample to a vial. Afterwards

the catalyst was separated from the sample by means of

centrifugation. Finally, the samples were analyzed using a

reversed-phase high-performance liquid chromatograph (RP-

HPLC). The components were identified using a UV-detector

with a variable wavelength. Quantification of the different

components in the reaction mixture was performed by relating

the peak surface areas to the amount of internal standard,

methyl 4-nitrobenzoate [3].

E. Vapour-Liquid equilibrium

Because the molar ratio of the reactants has an influence on

the reaction rate, it is important to know the exact composition

of the liquid phase after the vapour-liquid equilibrium has

been established. Thus, it is necessary to check how much of

each component has to be added to the Parr reactor, in order to

obtain the desired concentrations in the liquid phase. An

existing code based on the Non-Random Two-Liquid (NRTL)

and Hayden-O’Connell (HOC) methods has been modified

which allows for an initial estimation of the vapour-liquid

equilibrium. This code accounts for the thermodynamic non-

ideality of both the liquid as well as the gas phase via so-

called activity coefficients and fugacity coefficients.

III. EXPERIMENTAL RESULTS

A. Valorization of furfural via the aldol condensation

In order to obtain higher conversions for the aldol

condensation of furfural with acetone, higher reaction

temperatures were investigated. When higher temperatures

(>60°C) are applied, one has to consider the vapour-liquid

equilibria of the reactants. In particular for acetone, with a

boiling point of 56°C, one has to take into account that a

fraction of this reactant will be in the vapour phase.

The results from a simulation with the NRTL-HOC code

indicate that this system deviates a few percentages from an

ideal mixture. As was expected, only a fraction of the acetone

evaporates at higher temperatures. For toluene and furfural it

is correct to assume that everything remains in the liquid

phase. If a reaction temperature of 90°C is considered, about

1.55 % of the acetone will evaporate, which corresponds with

a pressure increase up to a total pressure of 3.7 bar.

For the aldol condensation of furfural with acetone at 90°C,

the decline of the turnover frequency as a function of the

volume percentage of water in the reaction mixture indicates

that a small amount of 1 vol% or 2 vol% water is already

sufficient to lower the catalytic activity with 23.43 %,

respectively 57.17 %.

Figure 3: TOF ratio as a function of the volume % of water in the

reaction mixture

0.0

0.2

0.4

0.6

0.8

1.0

0 2 4 6 8 10 12

TO

F r

ati

o

Water (vol %)

Page 10: Green aldol condensation: synthesis, testing and design of

B. Catalyst characterization

Characterization of the organosilicas affirmed the successful

synthesis and functionalization of the materials. With nitrogen

physisorption a type IV isotherm was obtained, indicating that

the material is mesoporous. After functionalization, a small

decrease in surface area and total pore volume - resulting from

the loss in free volume - was observed.

X-ray diffraction patterns confirm that for most materials a

hexagonal ordered mesoporous structure is maintained before

and after the functionalization. However, from the X-ray

diffractograms, it was found that the washing procedure at

120°C is too harsh for the "modified SBA-15" materials (T1).

Adapting the washing procedure to a milder temperature of

40°C and a shorter time (2h - 3h) results in a material which

properties are better preserved. Although, the mesoscopic

ordering is still lost and an amorphous solid was obtained.

Elemental analysis determined the amine loading to be in

the range of 0.05 – 0.15 mmol/g.

The DRIFT spectra demonstrated the presence of a primary

amine after functionalization with cysteine.

C. Aldol condensation of 4-nitrobenzaldehyde with acetone

The functionalized materials were used as catalyst in the

aldol condensation of 4-nitrobenzaldehyde with acetone at

45°C. To investigate the influence of the hydrophobicity of

the catalyst support on the catalytic activity, the reaction was

performed both in the absence and presence of water.

Although both catalyst have similar amine loadings, E1-C

(0.0592 mmol/g) and B10-C (0.0692 mmol/g) have very

different reaction rates. The turnover frequency of E1-C is

about one order of magnitude higher than B10-C.

For a possible explanation it is necessary to take into

consideration that the concentration of reactants in the pores

may be different from the concentration of the reaction

mixture. The properties of the pores play an important role on

the physisorption of the reactants into the pores. It is not

unlikely that for the material B10-C the environment of the

pores is more in favour of 4-nitrobenzaldehyde to enter due to

some interactions with the benzene groups in the support. This

can be justified with the "like likes like" principle.

As a consequence of the relative higher concentration of

4-nitrobenzaldehyde in the pores, this reactant reacts with the

primary amine and may lead to the formation of a stable

Schiff base, which eliminates active sites. This inhibition of

the reaction kinetics was also observed by Kandel et al. [4]

An overview of the turnover frequencies for the catalysts

comprising of both hydrophobic and hydrophilic blocks (TE-

and TB-materials) is given in Figure 4.

The lower turnover frequencies obtained from the catalytic

experiments indicate that no linear relationship applies for the

activity of the catalysts. Adding three different precursors to

the precursor mixture complicates the synthesis because every

precursor has its own hydrolysis and condensation rates.

When the difference between these rates is too large, the

tendency towards homocondensation reactions increases. This

will be a problem during the co-condensation because the

homogeneous distribution of the organic functionalities in the

framework cannot be guaranteed.

As a consequence, it is possible that the different precursors

are clustered in the support. And after functionalization, the

amine functionalities are too close to each other and are not

fully promoted by neighbouring silanol groups. Leading to

lower TOF-values than would be expected from linear

interpolation.

Figure 4: Turnover frequency as a function of the molar % of TEOS

in the catalysts. Blue: TE-materials; Red: TB-materials. Dashed lines

indicate the theoretical TOF values in case of linear interpolation.

D. Effect of water in the reaction mixture

The influence of the hydrophobicity of the catalyst support

on the catalytic activity in the presence of water is

investigated. This is an important property because for a

possible application in industry, e.g. in the valorization of

biocomponents such as glycerol and furfural, the reactants

will most likely be in an aqueous solution. Thus the activity of

these bifunctional catalysts in the presence of water will have

a major impact on the potential feasibility of these materials.

The turnover frequencies of the TE-materials both for the

reference reaction and the reaction in presence of 1 vol%

water are compared in Table 1. For the most hydrophobic

material, TE1-C, the catalytic activity remains at the same

level. It is an interesting result to see that, even in the presence

of water, TE1-C is able to preserve its activity. This indicates

that the hydrophobic pore walls impede the water molecules

to enter the pores. Thus the reaction continues as it would in

the situation without water. Interestingly, while in the

reference reaction the turnover frequency of TE1-C was still

lower than the one of T1-C-S2, this catalyst now becomes

more active in the presence of water. Materials with more

TEOS in the pore wall are more susceptible to the water and

therefore the effect of water is more pronounced.

The turnover frequencies of the TB-materials in the absence

or presence of water are shown in Table 2. In comparison to

the TE-materials, the presence of water has a much larger

effect on the catalytic activity. The TOF values for the aldol

condensation in the presence of 1 vol% water drop

significantly. All TB-materials exhibit a turnover frequency

around 9 x 10-6

s-1

.

An explanation why the ratio of the TB-materials is much

lower than the TE-materials has to do with the Snyder polarity

index. Water is one the most polar solvents, and has a Snyder

polarity index of 9, while benzene has a lower value of 3. The

polarity of an ethane is in the range of 0.1 to 0.3. Comparing

the polarity index of ethane and benzene, it can be expected

that the ethane groups in the silica support have a stronger

repulsive effect on water. Thus, the TE-materials are capable

to exclude water to a greater extent from the pores than the

TB-materials.

0.0E+00

2.0E-05

4.0E-05

6.0E-05

8.0E-05

1.0E-04

1.2E-04

1.4E-04

1.6E-04

1.8E-04

2.0E-04

0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0 100.0

TO

F (

1/

s)

TEOS (mol %)

Page 11: Green aldol condensation: synthesis, testing and design of

Table 1: Comparison of TOF values for the TE-materials

Material TOFreference reaction (s-1

) TOFreaction with water (s-1

) Ratio

E1-C 1.77 x 10-4

1.86 x 10-4

1.049

TE1-C 7.16 x 10-5

7.70 x 10-5

1.075

TE2-C 4.41 x 10-5

3.73 x 10-5

0.846

T1-C-S2 8.67 x 10-5

5.15 x 10-5

0.594

Table 2: Comparison of TOF values for the TB-materials

Material TOFreference reaction (s-1

) TOFreaction with water (s-1

) Ratio

B10-C 1.77 x 10-5

1.30 x 10-5

0.721

TB1-C 2.00 x 10-5

9.71 x 10-6

0.486

TB2-C 3.22 x 10-5

9.04 x 10-6

0.281

TB3-C 4.45 x 10-5

9.54 x 10-6

0.214

T1-C-S2 8.67 x 10-5

5.15 x 10-5

0.594

E. Regeneration

Finally, the reusability of the TE- and TB-materials was

investigated. First, assuming that no inactive intermediates are

formed and thus no important changes occur to the

heterogeneous catalyst, the material is simply washed with a

small amount of deionized water and afterwards with a larger

amount of acetone to improve the drying of the sample. When

the regenerated materials were tested in the aldol

condensation of 4-nitrobenzaldehyde with acetone at 45°C,

the turnover frequencies dropped significantly.

A second regeneration method was tested in which the spent

catalyst was thoroughly rinsed with ethyl acetate, deionized

water, and ethanol for several times and dried in vacuum at

40°C. This method was described by An et al. [5] and worked

fine for materials with alternating hydrophobic and

hydrophilic blocks in the pore wall. However, when these

regenerated materials were tested in the aldol condensation,

similar TOF values were obtained indicating that the inactive

intermediates were still present on the catalytic surface.

DRIFT spectra of the TE1-C material confirmed the presence

of nitro groups on the catalyst after reaction. Nonetheless, no

indication of an increased amount of imine was measured.

It has to be mentioned that the catalyst only loses its activity

during the first catalytic run. After that, the material retains a

constant catalytic activity, see Figure 5.

Figure 5: Conversion as a function of time for the TE1-C material.

Blue: reference reaction; Red: regeneration 1; Green: regeneration 2

IV. CONCLUSIONS & FUTURE WORK

The synthesis of Periodic Mesoporous Organosilicas (PMO)

containing vinyltriethoxysilane (VTES) was performed. These

materials were functionalized with cysteine via a thiol-ene

click reaction. Characterization of the organosilicas affirmed

the successful synthesis and functionalization of the materials.

Afterwards, the cooperative catalysts were tested for the aldol

condensation of 4-nitrobenzaldehyde with acetone. It was

found that the materials with ethane bridges are more active

than the ones with benzene groups in the support.

To investigate the influence of the catalyst support's

hydrophobicity on the catalytic activity, the reaction was also

performed in the presence of water. Especially, the catalysts

with ethane groups in the pore wall show promising results.

For the most hydrophobic material, TE1-C with 64.7 mol%

ethane, the catalytic activity remains at the same level. It is

interesting to see that, even in the presence of water,

TE1-C is able to preserve its activity. This indicates that the

hydrophobic pore walls impede the water molecules to enter

the pores.

The regenerated catalysts exhibit lower turnover frequencies

after the first catalytic run. However, in a second and third

catalytic run, the TOF-value doesn’t change anymore. There

are indications that the initial drop in activity is a consequence

of one-time only pore blocking. Afterwards, the active sites

that aren't blocked, remain accessible for subsequent

reactions, resulting in a steady catalyst's activity. It could be

useful to investigate the exact cause of this deactivation.

Different regeneration procedures have to be tested in order to

achieve again the original activities.

In this work, differences in surface arrangements due to

clustering of the precursors may have interfered with a correct

comparison between the catalysts’ activity. It might be useful

for further investigations on the effect of the catalyst support’s

hydrophobicity to carefully control the arrangement of the

active sites with respect to each other. A material with

alternating hydrophobic and hydrophilic blocks could provide

such possibility.

Finally, it is recommended to examine the functionalization

of these organosilicas with secondary amines containing an

alcohol group close to the amine. Because literature suggests

that intermolecular amine-silanol or intramolecular amine-

alcohol cooperativity is better than intramolecular amine-

carboxylic acid cooperativity [6].

0,0

0,5

1,0

1,5

2,0

2,5

3,0

3,5

4,0

4,5

5,0

0 50 100 150 200 250 300

Co

nve

rsio

n (

%)

Time (min)

TOF = 7.16 10-5 s-1

TOF = 1.90 10-5 s-1

TOF = 1.95 10-5 s-1

Page 12: Green aldol condensation: synthesis, testing and design of

ACKNOWLEDGEMENTS

The author would like to acknowledge the helpful hand and

suggestions of Els De Canck, Dolores Esquivel and Judith

Ouwehand of the Center for Ordered Materials, Organo-

metallics & Catalysis. The author also would like to thank

Tom Planckaert for the characterization of many samples via

elemental analysis and X-ray diffraction.

REFERENCES

1. Serrano-Ruiz, J.C. and J.A. Dumesic, Catalytic

routes for the conversion of biomass into liquid

hydrocarbon transportation fuels. Energy &

Environmental Science, 2011. 4(1): p. 83-99.

2. Chheda, J.N. and J.A. Dumesic, An overview of

dehydration, aldol-condensation and hydrogenation

processes for production of liquid alkanes from

biomass-derived carbohydrates. Catalysis Today,

2007. 123(1-4): p. 59-70.

3. Ouwehand, J. and J. Lauwaert, Internal discussion.

2015.

4. Kandel, K., et al., Substrate inhibition in the

heterogeneous catalyzed aldol condensation: A

mechanistic study of supported organocatalysts.

Journal of Catalysis, 2012. 291: p. 63-68.

5. An, Z., et al., L-Proline-Grafted Mesoporous Silica

with Alternating Hydrophobic and Hydrophilic

Blocks to Promote Direct Asymmetric Aldol and

Knoevenagel-Michael Cascade Reactions. Acs

Catalysis, 2014. 4(8): p. 2566-2576.

6. Lauwaert, J., et al., Spatial arrangement and acid

strength effects on acid–base cooperatively catalyzed

aldol condensation on aminosilica materials. Journal

of Catalysis, 2015. 325(0): p. 19-25.

Page 13: Green aldol condensation: synthesis, testing and design of
Page 14: Green aldol condensation: synthesis, testing and design of

Table of Contents i

Table of Contents

Table of Contents i

List of Figures iii

List of Tables v

Nomenclature vi

Chapter 1 Introduction 1

1.1 Aldol condensation 4

1.1.1 General mechanisms 4

1.1.1.1 Base-catalyzed aldol condensation 5

1.1.1.2 Acid-catalyzed aldol condensation 6

1.1.2 Types of aldol condensation 8

1.1.3 Applications 10

1.2 Catalysis: homogeneous vs heterogeneous 12

1.3 Mesoporous materials 14

1.3.1 Silica materials 14

1.3.2 Functionalized silica materials 14

1.3.2.1 Post-synthesis functionalization 16

1.3.2.2 One-pot synthesis 17

1.3.2.3 Periodic mesoporous organosilicas – PMOs 18

1.3.2.4 PMO synthesis parameters 21

1.4 Cooperative catalysis 24

1.4.1 Effect of the acidity of the acid group 25

1.4.2 Effect of the base functionality type 32

1.4.3 Effect of the distance between the base and acid functionality 38

1.4.4 Effect of the reaction medium 42

1.4.5 Effect of the hydrophobicity of the support 44

1.5 Scope of the thesis 45

Chapter 2 Materials and methods 47

2.1 Materials 47

2.2 Elemental Analysis 47

2.3 Nitrogen physisorption 48

2.4 X-Ray Diffraction 50

2.5 Diffuse Reflective Infrared Fourier Transform (DRIFT) Spectroscopy 52

2.6 Aldol condensation of furfural with acetone 53

Page 15: Green aldol condensation: synthesis, testing and design of

Table of Contents ii

2.7 Aldol condensation of 4-nitrobenzaldehyde with acetone 53

2.8 High Performance Liquid Chromatography (HPLC) 54

2.9 NRTL-HOC code 55

Chapter 3 Valorization of furfural via the aldol condensation 59

3.1 Aldol condensation of furfural with acetone 60

3.2 Optimization of the reaction conditions 61

3.3 Effect of water in the reaction mixture 63

Chapter 4 Synthesis and characterization of cooperative acid-base PMOs materials 65

4.1 Synthesis procedure 65

4.2 Functionalization 67

4.2.1 Thiol-ene “click” chemistry 67

4.2.2 Recipe 68

4.3 First set of materials 70

4.3.1 Molar composition of the PMO materials 70

4.3.2 Characterization 71

4.3.2.1 DRIFT 71

4.3.2.2 Nitrogen physisorption and elemental analysis 72

4.3.2.3 XRD 73

4.4 Reference work 75

4.5 Second set of materials 76

4.5.1 Determination of the theoretical amine loading 76

4.5.2 Molar composition of the PMO materials 77

4.5.3 Characterization 77

4.5.3.1 DRIFT 78

4.5.3.2 Nitrogen physisorption and elemental analysis 78

4.5.3.3 XRD 81

Chapter 5 Catalytic experiments 85

5.1 Comparison of ethane and benzene bridged organosilica supports 85

5.1.1 Catalysts containing both hydrophobic and hydrophilic blocks 86

5.2 Effect of water in the reaction mixture 87

5.2.1 Catalysts containing both hydrophobic and hydrophilic blocks 88

5.3 Regeneration 90

Chapter 6 Conclusions and Future Work 93

Chapter 7 References 97

Appendix A 101

Appendix B 104

Page 16: Green aldol condensation: synthesis, testing and design of

List of Figures iii

List of Figures

Figure 1: Conversion of biomass to refined products [1] ................................................................................. 2

Figure 2: Aldol condensation of 5-hydroxymethylfurfural (HMF) with acetone [3] ............................. 3

Figure 3: The aldol condensation [6] ......................................................................................................................... 4

Figure 4: Base-catalyzed aldol condensation [6] .................................................................................................. 5

Figure 5: Base-catalyzed dehydration of an aldol [6] ......................................................................................... 6

Figure 6: Acid-catalyzed keto-enol tautomerism [6] .......................................................................................... 7

Figure 7: Acid-catalyzed aldol condensation [6] .................................................................................................. 7

Figure 8: Acid-catalyzed dehydration of an aldol [6] .......................................................................................... 8

Figure 9: Aldol condensation between ethanal and propanal [6] ................................................................. 9

Figure 10: Formation of a cyclopentenone by an aldol cyclization [6] ....................................................... 9

Figure 11: Formation of a cyclohexenone by an aldol cyclization [6] ...................................................... 10

Figure 12: Robinson annulation reaction [6] ...................................................................................................... 10

Figure 13: Four possible stereoisomers of the aldol product [5] ............................................................... 11

Figure 14: Structures of mesoporous silicas: a) MCM-41, b) MCM-48 and c) MCM-50 [27] ........... 14

Figure 15: Overview of the different synthesis methods for obtaining hybrid materials [27] ...... 15

Figure 16: Post-functionalization of a silica material by anchoring R-Si-(OR')3 to the silanol

groups. R is an organic functionality, R' is an alkyl group [27] ................................................................... 16

Figure 17: Functionalization via the co-condensation of an organosilane with TEOS [27] ............. 17

Figure 18: Synthesis of a PMO material via the condensation of a bissilane in the presence of a

surfactant and an acidic or base environment [27] .......................................................................................... 19

Figure 19: Overview of precursors that have been used in PMOs. Terminal Si = Si(OR)3 with R

equal to –CH3 or –C2H5 [27]......................................................................................................................................... 20

Figure 20: Structure of the different PMO precursors used during the synthesis [43] 1:

bis(triethoxysilyl)ethane, 2: bis(triethoxysilyl)methane, 3: bis(triethoxysilyl)ethene, 4: 1,4-

bis(triethoxysilyl)benzene .......................................................................................................................................... 21

Figure 21: Interactions between the inorganic species (I) and the head group of the surfactant

(S) with consideration of the possible synthetic pathway in acidic, basic or neutral media [27] 23

Figure 22: Aldol condensation of 4-nitrobenzaldehyde with acetone [45] ............................................ 25

Figure 23: Proposed catalytic cycle [46] ............................................................................................................... 26

Figure 24: Possible pathway of proton transfer assisted by silanol groups [47] ................................ 28

Figure 25: TOFs of the catalysts (type A, rhombus) and the HMDS-treated catalysts (type B,

square) as a function of the molar silanol-to-amine ratio [55] ................................................................... 29

Figure 26: Proposed reaction mechanism for the aldol condensation of 4-nitrobenzaldehyde

with acetone in the presence of the promoting silanol groups [55].......................................................... 30

Figure 27: Schematic representation of the synthesis of SBA-X with X referring to the functional

group, i.e. 'I' for the iodo-group, 'A' for the secondary amine, 'AL' for the alcohol group, 'CA' for

the carboxylic acid and 'PA' for the phosphoric acid [56] ............................................................................. 31

Figure 28: Formation of a Schiff base between 4-nitrobenzaldehyde and the aminopropyl group

of AP-MSN [47] ................................................................................................................................................................ 34

Figure 29: Proposed reaction mechanism for the aldol condensation of 4-nitrobenzaldehyde

with acetone [58] ............................................................................................................................................................ 34

Figure 30: Nucleophilic addition of acetone to a primary amine [58] ..................................................... 35

Page 17: Green aldol condensation: synthesis, testing and design of

List of Figures iv

Figure 31: Percentage of amines promoted as a function of the silanol-to-amine ratio; Diamonds:

catalysts functionalized with primary amine; Squares: catalysts functionalized with secondary

amine [58] .......................................................................................................................................................................... 37

Figure 32: Proximal-C-A-SBA-15 and maximum-C-A-SBA-15 [49] ............................................................ 38

Figure 33: Proposed reaction mechanism for the aldol condensation reaction on the solid

support [49] ....................................................................................................................................................................... 39

Figure 34: Aminosilanes grafted onto SBA-15 with controlled linker lengths [53] ............................ 40

Figure 35: Conversion of 4-nitrobenzaldehyde in the aldol condensation with acetone at 50 °C

for the catalysts with different alkyl linkers [53] .............................................................................................. 40

Figure 36: Overall reaction pathways for 4-nitrobenzaldehyde and acetone in the presence of

AP-MSN: inhibition (red) and aldol reaction (black) [48] .............................................................................. 43

Figure 37: Presentation of a mesoporous silica with alternating hydrophobic and hydrophilic

blocks in the pore wall and the different interfaces inside the channels [61] ....................................... 44

Figure 38: Different types of adsorption isotherms [63] ................................................................................ 48

Figure 39: Type IV isotherm of SBA-15 [32] ........................................................................................................ 49

Figure 40: Diffraction of X-rays [65] ....................................................................................................................... 51

Figure 41: Scheme of the geometry of the XRD device [66] .......................................................................... 51

Figure 42: Representation of a hexagonal (P6mm) structure ...................................................................... 52

Figure 43: Aqueous-phase conversion of sugars and derivatives into liquid hydrocarbon fuels

[73] ........................................................................................................................................................................................ 59

Figure 44: Aldol condensation of furfural with acetone [26]........................................................................ 60

Figure 45: Conversion as a function of time for the aldol condensation of furfural with acetone

(blue) and 4-nitrobenzaldehyde with acetone (red) ........................................................................................ 61

Figure 46: Turnover frequencies for the aldol condensation of furfural with acetone at 90 °C as a

function of the molar ratio of the reactants (furfural/acetone) .................................................................. 62

Figure 47: TOF ratio as a function of the volume % of water in the reaction mixture ....................... 63

Figure 48: General overview of the synthesis ..................................................................................................... 65

Figure 49: Precursors added during synthesis. a) VTES: vinyltriethoxysilane b) BTEE: 1,4-

bis(triethoxysilyl)ethane c) BTEB: 1,4-bis(triethoxysilyl)benzene d) TEOS: tetraethyl

orthosilicate ....................................................................................................................................................................... 66

Figure 50: Reaction mechanism for the hydrothiolation of a C=C bond in the presence of a photo-

initiator and the appropriate UV radiation. In the propagation circle R stands for cysteine .......... 67

Figure 51: Structure of the components used during the post-functionalization a) Irgacure 2959

b) Cysteine ............................................................................................................................................ 68

Figure 52: Functionalization of the synthesized PMO materials ................................................................. 69

Figure 53: Structure of the functionalized PMO material ............................................................................... 69

Figure 54: Co-condensation of BTEE with VTES or BTEB with VTES ....................................................... 70

Figure 55: DRIFT spectrum of E10 (blue) and E10-C (red) ........................................................................... 72

Figure 56: XRD diffractogram of E10 (blue) and E10-C (red) ...................................................................... 73

Figure 57: Conversion as a function of time ......................................................................................................... 75

Figure 58: Comparison of the turnover frequencies at the same conditions ......................................... 75

Figure 59: DRIFT spectrum of TB1 (blue) and TB1-C (red) .......................................................................... 78

Figure 60: Comparison between the pore size distribution before and after functionalization and

washing procedures. Red: 120 °C, Green: 40 °C ................................................................................................. 80

Figure 61: XRD diffractogram of E1 (blue) and E1-C (red) ............................................................................ 81

Figure 62: XRD diffractogram of T1 (blue) and T1-C (red) ........................................................................... 82

Page 18: Green aldol condensation: synthesis, testing and design of

List of Tables v

Figure 63: Results of the aldol condensation of 4-nitrobenzaldehyde and acetone at 45 °C, Blue:

E1-C; Green: B10-C ......................................................................................................................................................... 85

Figure 64: Turnover frequency as function of the molar percentage of TEOS in the catalysts.

Blue: TE-materials; Red: TB-materials. Dashed lines indicate the theoretical TOF values in case of

linear interpolation ........................................................................................................................................................ 87

Figure 65: Turnover frequency as function of the molar percentage of TEOS in the catalysts.

Blue: TE-materials; Red: TB-materials. Dashed lines indicate the theoretical TOF values in case of

linear interpolation ........................................................................................................................................................ 88

Figure 66: Drift spectrum of TE1-C (red) and TE1-C-REGEN (green) ...................................................... 91

Figure 67: Conversion as a function of time for the TE1-C material. Blue: reference reaction; Red:

regeneration 1; Green: regeneration 2 .................................................................................................................. 92

Figure 68: Results of the aldol condensation of 4-nitrobenzaldehyde with acetone at 45 °C, Blue:

TE1-C; Green: TE2-C ....................................................................................................................................................106

Figure 69: Results of the aldol condensation of 4-nitrobenzaldehyde with acetone at 45 °C, Blue:

TB1-C; Red: TB2-C; Green: TB3-C ..........................................................................................................................107

List of Tables

Table 1: List of frequently used surfactants [44] ............................................................................................... 22

Table 2: Results of the catalytic tests [46] ............................................................................................................ 26

Table 3: Effect of SBA-15 as an additive on the amine-catalyzed aldol reaction [57] ........................ 32

Table 4: Turnover frequencies obtained with both the unpromoted base as the corresponding

cooperative acid-base catalysts [56, 58] ............................................................................................................... 36

Table 5: Catalytic properties [49] ............................................................................................................................ 38

Table 6: Solvent effect on the catalysis [45] ........................................................................................................ 42

Table 7: Components from the reaction mixture with their retention times, wavelength for

maximal absorption and calibration factors ....................................................................................................... 55

Table 8: Molar composition of the PMO materials ............................................................................................ 71

Table 9: Properties of the PMO materials ............................................................................................................. 72

Table 10: Physicochemical properties of the synthesized PMO materials ............................................. 74

Table 11: Molar composition of the organosilicas ............................................................................................ 77

Table 12: Properties of the organosilicas before and after functionalization ....................................... 79

Table 13: Properties of the T1 material before and after functionalization .......................................... 80

Table 14: Physicochemical properties of the synthesized organosilicas ................................................ 82

Table 15: Comparison of the turnover frequencies for the TE-materials ............................................... 89

Table 16: Comparison of the turnover frequencies for the TB-materials ............................................... 90

Table 17: Parameters of the NRTL obtained from Aspen Plus ..................................................................101

Table 18: Added amount of precursors ...............................................................................................................104

Table 19: Evolution of the physisorpted amine concentrations ...............................................................105

Page 19: Green aldol condensation: synthesis, testing and design of

Nomenclature vi

Nomenclature Roman Symbols

a Mass of the adsorbent [g]

A Cross-section of one adsorbate molecule [m²]

a0 Width of the unit cell [nm]

AP-MSN Amine-functionalized mesoporous silica nanoparticle

APTES 3-aminopropyltriethoxysilane

BdB-FHH Method of Broekhoff and de Boer, with the modification of Frenkel, Halsey

and Hill

BET Theory of Brunauer, Emmett and Teller

BJH Method of Barrett, Joyner and Halenda

BTEB 1,4-bis(triethoxysilyl)benzene

BTEE Bis(triethoxysilyl)ethane

BTEENE Bis(triethoxysilyl)ethene

BTME Bis(trimethoxysilyl)ethane

c BET constant

CTAB Cetyltrimethylammonium bromide

d Interplanar distance between two lattice planes [nm]

dp Average pore diameter [nm]

DRIFT Diffuse Reflective Infrared Fourier Transform

e Wall thickness of the material [nm]

F127 Pluronic F127

HMDS 1,1,1,3,3,3-hexamethyldisilazane

HMF 5-hydroxymethylfurfural

HOC Hayden-O'Connell model

Irgacure 2959 2-Hydroxy-4-(2-hydroxyethoxy)-2-methylpropiophenone

IUPAC International Union of Pure and Applied Chemistry

MAPTMS N-methylaminopropyltrimethoxysilane

MCM Mobil Composition of Matter

MPTMS 3-mercaptopropyltrimethoxysilane

MSN Mesoporous silica nanoparticle

MVK Methyl vinyl ketone

Page 20: Green aldol condensation: synthesis, testing and design of

Nomenclature vii

n Order of the reflection

NA Avogadro constant

NMR Nuclear Magnetic Resonance

NRTL Non-Random Two-Liquid model

OTAC Octadecyltrimethylammonium chloride

P123 Pluronic P123

PMO Periodic Mesoporous Organosilicas

PMS Periodic Mesoporous Silicas

RP-HPLC Reversed-phase high-performance liquid chromatograph

SBA Santa Barbara Amorphous

SDA Structure determining agent

SBET Specific surface area [m²/g]

TB Organosilica containing VTES, TEOS and BTEB

TE Organosilica containing VTES, TEOS and BTEE

TEOS Tetraethyl orthosilicate

TMOS Tetramethyl orthosilicate

TOF Turnover frequency [s-1]

V Adsorbed volume [cm³]

Vmon Volume corresponding to monolayer adsorption [cm³]

Vp Pore volume [cm³/g]

VTES Vinyltriethoxysilane

XRD X-ray diffraction

z Relative pressure (p/p°)

Greek symbols

ε Molar absorptivity at a given wavelength

θ Scattering angle [°]

λ Wavelength of the X-rays [nm]

Page 21: Green aldol condensation: synthesis, testing and design of
Page 22: Green aldol condensation: synthesis, testing and design of

Introduction 1

Chapter 1

Introduction

The aldol condensation may have good prospects in the discipline of green, renewable and

sustainable energy, e.g. in the valorization of glycerol and furfural. Fossil fuels, such as coal, oil

and natural gas provide more than 75 % of the world's energy. Even with the shale gas boom in

recent years, one cannot deny that the diminishing reserves of these non-renewable resources

cause a lot of concerns (e.g. rising oil prices, the increasing energy demand, limited supply of

fuels, etc.) [1]. In this regard, biomass can serve as a sustainable source for both the organic

carbon as for the energy in our industrialized society. Moreover, the production of energy from

biomass has the potential to generate lower greenhouse gas emissions, because the CO2 released

during energy conversion can be utilized for biomass regrowth. The Roadmap for Biomass

Technologies has predicted that by 2050, biofuels could provide 27 % of total transportation fuel

and 25 % of chemicals will be produced from biomass. At the same time reducing the total

amount of greenhouse gas emission in the energy sector by 80 % in 2050 (compared to the

levels in 1990) [2].

Figure 1 illustrates the conversion of biomass, and particularly lignocellulosic biomass, into

bioproducts and energy. In general, a variety of fuels and chemical intermediates can be

produced from carbohydrates by employing various types of reactions including hydrolysis,

dehydration, aldol condensation, reforming and oxidation. Which can be combined to produce

an even wider range of products [1].

Unlike petroleum which contains limited functionality, biomass-derived carbohydrates contain

an excess of functional groups (especially groups containing oxygen) which can be used as fuels

and chemicals. The challenge in this field is to develop methods to control the functionality in

the final product [1].

Page 23: Green aldol condensation: synthesis, testing and design of

Introduction 2

Figure 1: Conversion of biomass to refined products [1] The aldol condensation and subsequent hydrogenation are key intermediate steps to form larger

organic molecules using biomass-derived molecules. The production of heavier liquid-phase

alkanes from carbohydrates involves a series of reaction steps starting with an acid hydrolysis of

polysaccharides such as (hemi)cellulose and starch to produce monosaccharides such as

glucose, fructose and xylose. These carbohydrates can further undergo an acid-catalyzed

dehydration to form carbonyl-containing furan compounds such as 5-hydroxymethylfurfural

(HMF) and furfural. It is important to note that furfural and HMF cannot undergo self-

condensation, because they do not have an α-H atom that is required for the reaction. However,

both furfural and HMF have aldehyde groups that can condense with other molecules that can

form carbanion species. As shown in Figure 2, acetone forms an intermediate carbanion species

that can cross-condense with HMF in the presence of a base catalyst to form C9 species, which

can subsequently react with a second HMF molecule to form a C15 species. As a result of their

non-polar structure, these aldol adducts have a low solubility in water and thus precipitate out

of the aqueous phase. Next, the C=C and C=O bonds are saturated by hydrogenation in the

presence of a metal catalyst (usually Pd). Finally, these molecules are converted into liquid

alkanes (C9 - C15) by aqueous-phase dehydration/hydrogenation over a bifunctional catalyst

containing an acid and metal site [1, 3].

Page 24: Green aldol condensation: synthesis, testing and design of

Introduction 3

Figure 2: Aldol condensation of 5-hydroxymethylfurfural (HMF) with acetone [3]

Chapter 1 describes the different mechanisms of the aldol condensation and its applications.

The advantages of heterogeneous catalysts are briefly mentioned. Next, the properties of

mesoporous materials and the synthesis of periodic mesoporous organosilicas are discussed.

The last section of the introduction is about recent developments in cooperative catalysis.

An overview of the different effects on the catalyst’s activity is given.

The techniques and procedures which were employed for the characterization of the

synthesized periodic mesoporous organosilicas and the analysis of the catalytic experiments are

described in chapter 2.

In chapter 3, the results of the valorization of furfural via the aldol condensation are presented.

A comparison with the aldol condensation of 4-nitrobenzaldehyde with acetone is made.

The synthesis, functionalization and characterization of cooperative heterogeneous catalysts are

extensively discussed in chapter 4.

In chapter 5, catalytic experiments are performed with the synthesized organosilicas. The

influence of the hydrophobicity of the catalyst support on the catalytic activity in the presence of

water is investigated.

Chapter 6 summarizes the conclusions of this work. Also future work is proposed to continue

the research on this topic.

Page 25: Green aldol condensation: synthesis, testing and design of

Introduction 4

1.1 Aldol condensation

The aldol addition reaction is an acid- or base-catalyzed coupling reaction of a ketone or

aldehyde. The product of the aldol addition reaction, a β-hydroxy carbonyl compound, is called

an aldol because it contains both the functional group of an aldehyde and an alcohol.

Under certain conditions, the reaction product may undergo dehydration leading to an

α,β-unsaturated aldehyde or ketone. This variant is called the aldol condensation [4]. Even if the

dehydration is not spontaneous, it can usually be easily performed by heating the mixture,

because the new double bond is connected to the C=O bond. Condensation reactions combine

two (or more) molecules, often with the loss of a small molecule such as water or an alcohol.

This leads to a method which can be used for preparing α,β-unsaturated aldehydes (enal) and

ketones (enone) [4-6]. The reaction is shown in Figure 3.

The aldol condensation is a frequently used reaction in organic chemistry to connect two smaller

carbonyl groups to each other and is thus a convenient way to increase the carbon chain length.

It is an important reaction to create new C-C bonds and yield larger and more complex

molecules.

Figure 3: The aldol condensation [6] Because the product of an aldol condensation is again an aldehyde or ketone, a possible

troublesome side reaction is the further condensation of the aldol, enal or enone.

1.1.1 General mechanisms

Up to a few years ago, most aldol condensations have been homogenously catalyzed. These

homogenous catalysts exhibit an excellent selectivity and, due to a good understanding of the

reaction mechanism, allow an easy identification of the intermediates.

Although acid-catalyzed aldol reactions (e.g. by means of sulfuric acid) are known [7], the most

common form of the reaction uses a base. During the first years, sodium hydroxide and sodium

carbonate were commonly used in this reaction. However recently, stronger bases, such as

alkoxides (RO-) or amides (R2N-) are also often used. Also weakly base amines have been used as

Page 26: Green aldol condensation: synthesis, testing and design of

Introduction 5

catalysts in the aldol condensation. Depending on the nature of the catalyst, different reaction

mechanisms are obtained.

1.1.1.1 Base-catalyzed aldol condensation

The mechanism of a strong base-catalyzed reaction involves the formation of an enolate ion

(equilibrium between two ions), followed by a nucleophilic addition of this enolate to the

carbonyl group of another aldehyde or ketone. In a third step, protonation of the formed

alkoxide anion leads to the formation of the aldol product and the recovery of the basic catalyst

(see Figure 4). Finally, dependent on the acidity of the α proton of the aldol product, dehydration

can occur. Abstraction of the α proton gives an enolate that can eliminate hydroxide ion to give a

more stable, conjugated product [4, 6]. The dehydration mechanism is shown in Figure 5.

Figure 4: Base-catalyzed aldol condensation [6]

Page 27: Green aldol condensation: synthesis, testing and design of

Introduction 6

Figure 5: Base-catalyzed dehydration of an aldol [6] The hydroxide ion is not strong enough to convert substantially all of an aldehyde or ketone

molecule to the corresponding enolate ion. This means that the equilibrium of step 1 in Figure 4

is shifted to the left. Nevertheless, enough enolate ion is present for the reaction to proceed [5].

The location of this equilibrium can be influenced by the solvent. Protic solvents, such as water

or ethanol, are acidic enough to donate protons and react with the enolate anion, thus shifting

the equilibrium to the left. In an aprotic solvent, such as acetone or acetonitrile, the equilibrium

lies more to the right. But in the next steps of the mechanism the enolate ion reacts with another

carbonyl to form an alkoxide, followed by a protonation to give the aldol product. By removal of

the enolate ion, the equilibrium of step 1 is slightly shifted to the right [5].

1.1.1.2 Acid-catalyzed aldol condensation

It is also possible to carry out the aldol condensation under acidic conditions. In a first step, the

enol is formed by an acid-catalyzed keto-enol tautomerism as is shown in Figure 6. Hereby, a

proton is moved from the α carbon to oxygen by first protonating oxygen and then removing a

proton from carbon.

Page 28: Green aldol condensation: synthesis, testing and design of

Introduction 7

Figure 6: Acid-catalyzed keto-enol tautomerism [6] Next, the enol form of the aldehyde or ketone serves as a weak nucleophile to attack a

protonated carbonyl group. Deprotonation of the resonance-stabilized intermediate gives the

aldol product (see Figure 7).

Figure 7: Acid-catalyzed aldol condensation [6] Finally, this aldol product can be transformed into a conjugated α,β-unsaturated aldehyde or

ketone. The mechanism is similar to the dehydration reaction of alcohols. Dehydration results

from E1 elimination of the protonated alcohol functionality as can be seen in Figure 8. First,

protonation converts the hydroxyl group of the aldol product to a good leaving group. After the

water leaves, a carbocation is formed. Followed by the removal of a proton, gives the new C=C

double bond.

Page 29: Green aldol condensation: synthesis, testing and design of

Introduction 8

Figure 8: Acid-catalyzed dehydration of an aldol [6]

1.1.2 Types of aldol condensation

The aldol reaction can be divided into two categories. First there is the selfcondensation, this is a

reaction between two molecules of the same aldehyde or the same ketone. The second group are

bundled under the name crossed aldol condensation. These reactions involve two different

aldehydes or two different ketones, or a combination of an aldehyde and a ketone.

The reaction rate of the selfcondensation of two ketones is lower than that of two aldehydes due

to sterical hindrance. When the enolate of one aldehyde (or ketone) adds to the carbonyl group

of a different aldehyde or ketone, the result is called a crossed aldol condensation. If the

compounds used in the reaction are not selected carefully, a mixture of four products will be

formed. However, if one of the reactants does not have an α hydrogen, only two aldols are

possible, and in many cases the crossed product is the main one. This crossed-aldol reaction is

often called the Claisen-Schmidt reaction [6, 8]. As an example, the aldol condensation between

ethanol and propanal is shown in Figure 9.

Page 30: Green aldol condensation: synthesis, testing and design of

Introduction 9

Figure 9: Aldol condensation between ethanal and propanal [6] In general, the reactions in the addition phase are reversible. The equilibrium constant for the

dehydration phase is usually favourable because of the conjugated α,β-unsaturated carbonyl

system that is formed. When the reaction conditions are sufficient to cause dehydration, the

overall reaction can go to completion, even if the equilibrium constant for the addition step is

unfavourable (e.g. selfcondensation of ketones) [4].

Aldol reactions are often used to close five- and six-membered rings. This intramolecular aldol

condensation is well known under both basic [6] or acid [9] conditions. Because of the favorable

entropy, such aldol cyclization generally take place without much difficulty. Aldol cyclizations of

rings larger than six and smaller than five are less common because larger and smaller rings are

less favored due to their larger ring strain (mixture of angle strain and torsional strain). Figure

10 shows how a 1,4-diketone can condense and dehydrate to give a cyclopentenone. While

Figure 11 shows how a 1,5-diketone gives a cyclohexenone [5, 6].

Figure 10: Formation of a cyclopentenone by an aldol cyclization [6]

Page 31: Green aldol condensation: synthesis, testing and design of

Introduction 10

Figure 11: Formation of a cyclohexenone by an aldol cyclization [6] An important extension of the intramolecular aldol condensation is the Robinson annulation

reaction, which has been often used in the synthesis of steroids and terpenes [10]. The

mechanism is shown in Figure 12 and begins with the Michael addition of the cyclohexanone

enolate to methyl vinyl ketone (MVK), forming a δ-diketone. The enolate of the methyl ketone

attacks the carbonyl of the cyclohexanone. Finally, the aldol product dehydrates to give a

cyclohexenone .

Figure 12: Robinson annulation reaction [6]

1.1.3 Applications

The aldol condensation is used for the synthesis of organic compounds, particularly specialty

chemicals such as perfumes and synthetic flavours. Many aldehydes have very typical odours.

For example, benzaldehyde has a characteristic almond-like odour and is at low levels used as an

artificial flavouring in many nutrients [11]. Perfumers have an almost unlimited number of raw

materials available to create fragrances. These raw materials can be grouped into three classes:

natural essential oils, synthetic chemicals and semi-synthetic chemicals derived by chemical

processing of natural products. A number of important perfume chemicals are manufactured on

a large scale by a variety of condensation reactions, including the aldol condensation. Five

widely used products made via aldol condensation are pseudomethylionone, α-amylcinnamic

aldehyde, α-hexylcinnamic aldehyde, alkyl-α-methyldihydrocinnamic aldehyde and cinnamic

aldehyde. The reactions are catalyzed by alkali under condition which minimize self-

condensation of the more reactive aldehyde, i.e. propionaldehyde in the preparation of the

unsaturated intermediates for Cyclamen aldehyde and acetaldehyde for cinnamic aldehyde [12].

Page 32: Green aldol condensation: synthesis, testing and design of

Introduction 11

The aldol condensation can be used to synthesize a larger molecule with a specific

functionalization from smaller reagents. Therefore, the reaction is also widely applied in the

pharmaceutical industry, as well as for the preparation of coatings, plastics and many other fine

chemicals [13].

The wide synthetic applicability of the aldol reaction can be attributed to the ability to achieve

both versatility in reactants and control of regiochemistry and stereochemistry. The term

directed aldol addition is applied to reaction conditions that are designed to achieve specific

regio- and stereochemical outcomes. Control of the product structure requires that one reactant

acts exclusively as the electrophile and the other exclusively as the nucleophile [4].

The reaction can be made regioselective by preforming the reactive nucleophilic enolate and

ensuring that the addition step is fast. The aldol reaction creates two new stereogenic centers,

and, in the most general case, there are four stereoisomers of the aldol product, which are

presented in Figure 13 [4, 5].

Figure 13: Four possible stereoisomers of the aldol product [5] Aldol condensations can also be made enantioselective by using chiral enol derivatives, chiral

aldehydes or ketones, or both. In this case only one of the four isomers of the products will be

dominant. Chiral amino acids, such as proline, can be used for this reaction [5].

Page 33: Green aldol condensation: synthesis, testing and design of

Introduction 12

1.2 Catalysis: homogeneous vs heterogeneous

Catalysis is an important field in chemistry. Today over 90 % of all industrial chemicals are

produced with the aid of catalysts [14]. Because of environmental issues, energy supply and

pollution control, catalysis appears to be even more important than before. The IUPAC

(International Union of Pure and Applied Chemistry) has defined a catalyst as "a substance that

increases the rate of a reaction without modifying the overall standard Gibss energy change in

the reaction" [15]. The catalyst is both a reactant and a product of the reaction, meaning that the

catalyst is restored after each catalytic cycle. Also, the catalyst does not influence the

thermodynamical equilibrium composition after the reaction has terminated [14, 16].

In the 1990s the 12 principles of green chemistry have been introduced, a comprehensive

set of guidelines including the atom economy, the use of reusable catalysts and the minimization

of waste [17]. Within green chemistry, there is an increasing interest in developing novel

materials that can be applied as heterogeneous catalysts for chemical processes. Thus replacing

effective but hazardous chemicals as H2SO4, HF and AlCl3. The essence behind green chemistry is

that instead of using every possible chemical process, only those that are environmentally

benign should be used. The E-factor, a quantitative parameter that can be used to evaluate the

environmental acceptability of a process, was proposed by Sheldon in 1992 and reports the

amount of unit waste produced per unit product. Ideally this factor should be equal to zero,

but an analysis of the waste generation for each chemical industry sector, highlights the

need for improvements especially in fine chemicals (E-factor: 5-50) and pharmaceuticals

(E-factor: 25-100) [16-19].

The concern about renewables, CO2 levels and sustainability lead to a new chemical field

oriented at biomass feedstocks which is still very dependent upon new catalysts to make it

competitive with petroleum based feedstocks [14]. More efficient catalytic processes require

improvements in the catalytic activity and selectivity. Both aspects can be improved by a

tailored design of catalytic materials with the desired structures and the desired distribution of

the active sites. Materials such as zeolites, metal organic frameworks and mesoporous materials

offer such possibilities [16].

The first zeolite, stilbite, was discovered by Axel Fredrik Cronstedt in 1756 who found that the

mineral loses water rapidly on heating and thus seems to boil. The name zeolite comes from the

Greek words ζέω (zéō = to boil) and λίθος (líthos = stone). But it is only since 1955 that the

development of synthetic zeolites lead to the discovery of many new and shape selective

catalytic processes [14].

Page 34: Green aldol condensation: synthesis, testing and design of

Introduction 13

Since the end of the twentieth century, societal concerns have become an important motivation

for intensive research to convert the current by-products to useful products and to treat all

kinds of wastes in order to preserve and protect the environment. A challenge for our century is

the control of technological processes in the chemical, petrochemical and pharmaceutical

industries to develop atom-economical, environmental friendly processes with as little as

possible by-products [16].

Nowadays, the substitution of the homogeneous by the more environmentally friendly

heterogeneous process is highly desirable. Because of environmental concerns, heterogeneous

counterparts for the homogeneous catalysts are being investigated. The use of insoluble solid

catalysts presents several advantages. For one, the reaction temperature can be raised because

solvent is not necessarily required and consequently, aldol condensations in the gas phase can

be performed. Secondly, solid catalysts result in less energy intensive separation processes and

problems related to the disposal of strong acid or basic solutions can be avoided. The

regeneration and recycling of used catalyst enhance the overall efficiency of the process. Also,

the reactions can be carried out without the reactor being exposed to concentrated solutions of

corrosive acids or bases [20].

In the past two decades, a variety of heterogeneous catalysts - including zeolites [21, 22], anion-

exchange resins [23], Mg-Al layered double hydroxides [24, 25] and metal organic frameworks

[26] - have already been tested for the aldol condensation. But most advantages can perhaps be

achieved with mesoporous materials. By combining the properties of organic and inorganic

building blocks within a single material it is possible to combine the enormous functional

variation of organic chemistry with the advantages of a thermally stable and robust inorganic

substrate. The symbiosis of organic and inorganic components can lead to materials whose

properties differ considerably from those of their individual, isolated components. Through the

incorporation of the desired functional groups in a mesoporous structure, a solid heterogeneous

catalyst is created [27]. These materials will be discussed in the next chapter.

Page 35: Green aldol condensation: synthesis, testing and design of

Introduction 14

1.3 Mesoporous materials

It is only since 1980 that serious attempts have been made to create "zeolite-like" materials with

ordered and uniform pores which are greater than two nanometers. These larger pores would

allow the host-guest chemistry of large molecules and polymers (= guest) to realize specific

interactions with the support (= host), but also increase the rate of diffusion of these molecules

in the pores. Larger pores would be a huge advantage in for example the fine chemical and

pharmaceutical industries as bulky products do not fit in smaller pores [28, 29].

1.3.1 Silica materials

During the last decades, different types of mesoporous materials (pore diameter between 2 and

50 nm) are synthesized, but one of the most important discoveries was the development of

ordered periodic mesoporous silicas (PMS) with a particular arrangement (MCM-41: hexagonal,

MCM-48: cubic, MCM-50: laminar, see Figure 14) of the mesopores. In 1992, scientists at Mobil

Oil Corporation discovered the direct synthesis of mesoporous materials, the Mobil Catalytic

Materials or MCM [30]. In the same year, researchers at the University of Santa Barbara also

developed mesoporous silica materials (SBA-15) by using a triblock copolymer (P123 or F127)

as surfactant [31, 32].

Figure 14: Structures of mesoporous silicas: a) MCM-41, b) MCM-48 and c) MCM-50 [27]

1.3.2 Functionalized silica materials

After the development of the MCM and SBA-materials, the interest to incorporate organic groups

in the periodic mesoporous materials has increased in order to enrich them with different

properties. These periodic mesoporous silica materials, with high internal surface areas up

to about 1000 m²/g and pore diameters up to 6 nm, are the ideal starting point for the

incorporation of many and large organic groups. The integration of functional organic groups in

Page 36: Green aldol condensation: synthesis, testing and design of

Introduction 15

the inorganic structures has led to the discovery of organic-inorganic hybrid materials with well-

defined porous structures and unique properties [33-35].

The synthesis of these hybrid materials can proceed via three routes and an overview of the

three different synthesis methods is shown in Figure 15. Every synthesis method will be

discussed in more detail.

1. A post-synthesis functionalization takes place through the anchoring of terminal organic

groups on the pore walls by making use of the reactivity of the silanol groups in the

material. (grafting method; paragraph 1.3.2.1)

2. The condensation of a tetraalkoxysilane with an terminal trialkoxyorganosilane of the

type R-Si-(OR’)3 is a one-pot synthesis, in which R is an organic functionality, and R’ is an

alkyl group (co-condensation method; paragraph 1.3.2.2)

3. The direct condensation of a poly(trialkoxysilyl) bridged organic component, eg.

(R’O)3-Si-R-Si-(OR’)3. (Periodic Mesoporous Organosilicas – PMOs; paragraph 1.3.2.3)

Figure 15: Overview of the different synthesis methods

for obtaining hybrid materials [27]

Page 37: Green aldol condensation: synthesis, testing and design of

Introduction 16

1.3.2.1 Post-synthesis functionalization

The term grafting refers to a post-

synthesis modification of the interior

surface with organic groups. This

procedure involves the condensation

of an organosilane, typically of the type

R-Si-(OR’)3, with the free silanol groups

present on the pore walls (Figure 16).

This procedure is also called post-

functionalization. These materials are

eventually made from an inorganic

surface on which an organic layer is

grafted [27, 33].

Figure 16: Post-functionalization of a silica

material by anchoring R-Si-(OR')3 to the

silanol groups. R is an organic functionality,

R' is an alkyl group [27]

Because of the significant quantity of possible organosilanes one can, by variation of the organic

group R, prepare a large set of organic functionalized silica materials with different chemical and

physical properties. The amount of organic groups is limited by the number of free silanol

groups on the pore walls.

A second advantage to this method is that the original mesostructure of the silica material

almost remains unchanged when a controlled functionalization of the organic groups occurs.

As a third advantage one can, by using bulky R-groups, selectively anchor these organosilanes to

the pore-opening. This allows them to block or completely shut-off the pores in such a way that

the air is trapped inside the pore. This property is of great importance for obtaining materials

with a low dielectric constant or also known as low-k materials (i.e. an important application of

PMO materials) [32, 33].

If during the beginning of the synthesis, the organosilanes bind mainly to the pore mouth, the

further spreading of the following molecules can be obstructed. This leads to a non-

homogeneous distribution of the organic groups in the pores and a lower degree of occupation.

A second disadvantage of this post-functionalization is that by introducing organic groups at the

surface, the pore volume will logically decrease. Moreover, this post-synthesis functionalization

requires of course an additional synthesis step [27, 32].

Page 38: Green aldol condensation: synthesis, testing and design of

Introduction 17

1.3.2.2 One-pot synthesis

The co-condensation method is, in contrast to the anchoring of organosilanes (see paragraph

3.2.1), a functionalization method in which all the precursors are already together at the

beginning of the reaction, as shown in Figure 17. The hybrid material is prepared by the co-

condensation of tetra-alkoxysilanes, Si(OR)4, usually tetraethyl orthosilicate (TEOS) or

tetramethyl orthosilicate (TMOS), with terminal tri-alkoxyorganosilanes R-Si-(OR’)3 in the

presence of a structure determining agent (SDA) [32, 33].

Figure 17: Functionalization via the co-condensation of an organosilane with TEOS [27] Because the organic groups are added during the beginning of the synthesis and are thus already

part of the pore walls, the chance of pore-blocking is smaller. One obtains a more homogeneous

distribution of the organic functions in comparison with the materials via post-synthesis

functionalization [36, 37]. However, it should be mentioned that the homogeneous distribution

of the organic functions is highly dependent on the hydrolysis and condensation rates of the

silica and the organosilica precursors. If the difference in relative velocities is too large, it may

give rise to the separate condensing of the two components. This process is also referred to as

self-condensation and prevents the homogeneous distribution of the organic groups [32, 33].

A second disadvantage to this method is the effect of the organic groups on the degree of

mesoscopic ordering of these materials. In general, the ordering decreases with an increasing

concentration of organosilanes and in the worst case will lead to completely unstructured

materials. Therefore, no more than 40 mole percent of the organic functions is usually added.

Typical values are in the majority of cases even lower, around 15 – 25 mole percent [27, 34].

Page 39: Green aldol condensation: synthesis, testing and design of

Introduction 18

This limitation may be due to a number of aspects: the organic functions are likely to interfere

with the formation of the silica structure and reduce the formation of micelles, especially when it

concerns bulky R-groups [38].

An increasing concentration of organosilanes will also have a decrease of the pore diameter,

pore volume and specific surface area as a result. Finally, a last disadvantage which must be

taken into account is that during the removal of the surfactant, the organic groups functions may

not be damaged or destroyed. For these reasons, a calcination step is not advisable and generally

only extraction methods are used [27, 32].

1.3.2.3 Periodic mesoporous organosilicas – PMOs

Although the previous synthesis methods are widely used in order to obtain some important

hybrid materials, both the post-synthesis and the one-pot synthesis have advantages as well as

disadvantages. As a solution to these drawbacks, a third alternative route was developed in

order to obtain organic-inorganic ordered mesoporous materials with a high concentration of

organic groups, wherein the organic and inorganic units are uniformly distributed in the

structure. These Periodic Mesoporous Organosilica materials, often abbreviated as PMO,

may be formed through the direct condensation of bridged organosilanes, for example

(R’O)3-Si-R-Si-(OR’)3, in the presence of a structure determining agent [27, 32, 33].

With this synthesis method, the organic units are directly incorporated into the three-

dimensional silica structure via two covalent bonds. This is not the case with the organic

functionalized silica materials which have been obtained via post-functionalization or direct

synthesis. In these last two methods, a combination of a tetra-alkoxysilane and a terminal tri-

alkoxysilane of the type R-Si-(OR’)3 is used. During the synthesis the alkyl group (OR’)3 co-

condense with each other, while the organic functionality R is oriented into the pore.

The first materials synthesized in this way (see Figure 18) have a homogeneous distribution

of the organic groups, a very large internal surface area and an improved thermal stability.

PMOs are characterized by periodically ordered pores and a very small pore size distribution.

These materials are also characterized by both the large amount and the uniform distribution of

the organic groups in the pore walls; large internal surface areas; thick pore walls; large pores

and high pore volumes [27, 39].

Page 40: Green aldol condensation: synthesis, testing and design of

Introduction 19

Figure 18: Synthesis of a PMO material via the condensation of a bissilane

in the presence of a surfactant and an acidic or base environment [27]

Because the organic functions are embedded in the walls of the channels, the pores remain fully

open and they are easily accessible for further manipulations (see paragraph 4.2.1) of the

chemical and physical properties without losing the stability of the porous structure [33].

With the help of this third route, one can circumvent problems such as the non-uniform

distribution and the influence of the organic groups on the mesoscopic ordering, the blocking of

the pores, different hydrolysis and condensation rates of the precursors and possible other

disadvantages of the grafting and co-condensation procedure [36].

The first PMO materials were synthesized in 1999 by three independent research groups.

Stein et al. [40] used two precursors with different bissilanes, namely bis(triethoxysilyl)ethane

(BTEE) and bis(triethoxysilyl)ethene (BTEENE). As a surfactant, cetyltrimethylammonium

bromide (CTAB) was used. In the same year, Inagaki et al. [41] succeeded to convert

bis(trimethoxysilyl)ethane (BTME) to an organic-anorganic hybrid material in base conditions

and in the presence of octadecyltrimethylammonium chloride (OTAC) as a structure

determining agent. Finally, Ozin et al. [42] used a combination of bis(triethoxysilyl)ethane

with TEOS, which was dissolved in a solution of CTAB, ammoniumhydroxide and water to obtain

a periodic mesoporous organosilica.

During the first years, only a few precursors (methane, ethane, ethene and benzene bissilanes)

were used frequently used. But nowadays, there is a lot of research for introducing all kinds of

bridged organic components. Figure 19 gives a small overview of organic precursors that have

been used in PMO materials [27, 32, 33].

Page 41: Green aldol condensation: synthesis, testing and design of

Introduction 20

Figure 19: Overview of precursors that have been used in PMOs.

Terminal Si = Si(OR)3 with R equal to –CH3 or –C2H5 [27]

Starting the PMO synthesis from these bissilanes (precursors) one can synthesize materials with

specific functions [32, 33]. Although the list of organic precursors is not exhaustive, only with a

limited number of these precursors an ordered mesoporous hybrid material can be prepared

which only consists of that precursor. This is because the pore walls of the PMO material must

possess a certain rigidity. Therefore, short chains or conjugated systems give better results

If the material does not have the desired structure, one usually chooses a co-condensation with a

rigid organic precursor with an ethane or benzene function as organic bridge. These are called

bifunctional PMO materials [27, 33, 34]. The bridged organic components should not be too

bulky either. If the distance between two inorganic units (Si-R-Si) in the pore wall becomes too

large, there is a risk that the pore walls will collapse. A final constraint is the interaction between

the surfactant and the precursor. To avoid a possible phase separation with the surfactant, one

can adjust one of the following parameters: acidity, temperature, concentration of both

components, type of surfactant, environmental conditions and additives [33, 36].

Page 42: Green aldol condensation: synthesis, testing and design of

Introduction 21

1.3.2.4 PMO synthesis parameters

- Precursor:

In the last years many syntheses have been discussed in the literature. The great majority of

reports describe the synthesis and characterization of one specific PMO. Often, the reported

synthetic procedure does not produce ordered organosilicas when different precursors are

substituted. This can be explained by the kinetics of the precursor since every precursor has its

own specific hydrolysis and condensation rates.

A general procedure to prepare a wide variation of PMO materials with different organic

functionalities was described by Burleigh et al. [43] in 2004. A typical synthesis procedure starts

with adding concentrated HCl to deionized water to obtain the initial aqueous solutions. Brij-76

surfactant was added to the solution upon stirring, and the mixtures were heated at 50 °C for 12

hours. The precursors (Figure 20) were then added to the resulting clear solution. The synthesis

mixtures were covered and stirred at 50 °C for 12 hours, followed by heating at 90 °C under

static conditions for an additional 24 hours.

Figure 20: Structure of the different PMO precursors used during the synthesis [43]

1: bis(triethoxysilyl)ethane, 2: bis(triethoxysilyl)methane,

3: bis(triethoxysilyl)ethene, 4: 1,4- bis(triethoxysilyl)benzene

The versatility of this synthesis can be attributed to a combination of key factors that affect the

surfactant template approach to synthesizing porous materials. These include the precursor

molecular structure and reactivity, interactions between the surfactant and the hydrolyzed

precursors, the overall stability of Brij-76 micelles and the acidic reaction conditions [43].

- Surfactant:

An overview of some frequently used surfactants during the synthesis of PMO materials is given

in Table 1. PMOs synthesized using an ionic surfactant as a structure determining agent often

have pore diameters which are restricted to the range between 2 to 5 nm. This limitation was

overcome by using non-ionic triblock copolymers of the type EOxPOyEOx (EO = poly(ethylene

oxide) and PO = poly(propylene oxide)). Characteristic to using P123 or F127 as template is that

mesoporous structures can be obtained under various pH values. This allows for the synthesis of

PMO materials with pore diameters bigger than 5 nm [44].

Page 43: Green aldol condensation: synthesis, testing and design of

Introduction 22

Table 1: List of frequently used surfactants [44]

Abbreviation Name Type interaction Structure

Ionic: pore diameter: 2 - 5 nm

CTAC/CTAB Cetyltrimethylammonium

chloride/bromide S+I-

of

S+X-I+

OTAC Octadecyltrimethylammonium

chloride

Non-ionic: pore diameter: > 5 nm

Brij-56 Polyoxyethylene(10) cetyl ether

S°I°

of

S°H+X-I°

Brij-76 Polyoxyethylene(10) stearyl ether

P123 Pluronic P123

F127 Pluronic F127

The alkylethylene oxide surfactant Brij-76 is an ideal choice for a PMO template. In addition,

Brij-76 is inexpensive, non-toxic and biodegradable. It also has a better template stability during

the synthesis owing to the relatively low solubility of Brij-76 in ethanol [43].

- Acidity

The acidity of the reaction mixture influences the interactions between the precursor and the

surfactant. It also has an impact on the hydrolysis and condensation rate of the precursor. This

parameter is very important for the ordering in the PMO material. Figure 21 shows the possible

interactions between the head groups of the surfactant and the inorganic species, taking into

account whether these interactions take place in a basic (21a and 21c), acid (21b and 21d) or

neutral (21e and 21f) media [27].

Page 44: Green aldol condensation: synthesis, testing and design of

Introduction 23

Figure 21: Interactions between the inorganic species (I) and the head group of the surfactant (S) with consideration of the possible synthetic pathway

in acidic, basic or neutral media [27]

- Temperature

The reaction temperature can play an important role. Each synthesis will feature an optimal

temperature which promotes the formation of larger pores. Within a certain procedure, the

various synthesis steps can take place at different temperatures. For example, the synthesis can

start at room temperature, but in a second step require 80- 100 °C. Another important factor is

the stirring time, which determines how long the temperature is maintained. Thanks to stirring,

a better distribution of heat is obtained, this also has an effect on the pore size distribution.

Page 45: Green aldol condensation: synthesis, testing and design of

Introduction 24

1.4 Cooperative catalysis

In the last decade, incorporating multiple types of active sites into solid materials has become an

area of interest in the synthesis of catalysts. The design of multifunctional materials capable of

carrying out heterogeneous catalysis through the incorporation of organic functional groups has

been growing in attention. In organisms, the aldol reaction is performed by aldolases, which

activate donor ketones with the amino group of lysine, to give enamines. Next, the enamines

attack aldehyde acceptors and are then hydrolysed to release the product. Enzymes, such as

aldolases, are interesting examples of multifunctional catalysts as they are capable of handling

multistep reactions. Enzymes can immobilize functional groups, which are mutually

incompatible and would not be tolerated together in solution, in a certain way that maintains

their independent functionality [45-47].

Surface-modified mesoporous materials with various active sites have been widely investigated

in recent years. Monofunctionalized mesoporous silica catalysts have been demonstrated to

have unique properties. But for many applications, bifunctional mesoporous catalysts are much

more wanted because the combined functionalities may act in a cooperative way to improve the

reactivity of the catalyst. Unravelling of reaction mechanisms in heterogeneous catalysis poses

additional challenges due to the different environments that the active sites can encounter on a

solid support, interfacial phenomena, competitive adsorption and kinetics of mass transfer [47-

49].

Many research groups have been devoting their time to this cooperative catalysis [45-56]. They

found out that the catalytic activity of a heterogeneous catalyst is dependent on many factors,

such as the acidity of the acid group, the type of the amine, the distance between both the base

and acid functionality. Also the reaction medium plays an important role during the reaction.

During the last years, different reaction mechanisms have been proposed. These aspects will be

discussed in the following chapter.

Most of them use SBA-15 (Santa-Barbara Amorphous) as a heterogeneous support. SBA-15 is

a well-ordered 2d-hexagonal mesoporous silica. Due to their large pore sizes and surface areas,

these mesoporous silica materials provide sufficient space to incorporate multiple functional

groups while the stability of the matrix is ideal for many types of reaction. Thus, making these

materials very useful as model catalyst for research purposes. Attachment of an organosilane

onto a mesoporous silica material is commonly achieved through grafting or via co-

condensation. Both methods have been discussed in the previous chapter.

Page 46: Green aldol condensation: synthesis, testing and design of

Introduction 25

Coupling reactions that are typically acid or base-catalyzed, such as the aldol reaction, can be

accelerated through the addition of a second component that acts cooperatively. The two

catalytic components can separately activate the two different substrates, yielding a lower

energy reaction pathway [50]. In the context of catalysis, the term cooperativity refers to a

system where at least two different catalytic entities act together to increase the rate of a

reaction beyond the sum of the rates achievable from the individual entities alone [51].

1.4.1 Effect of the acidity of the acid group

Last years, attention has been focused on combinations of organic functional groups (amines,

thiols and sulfonic acid groups). In 2006, Zeidan et al. [45] reported the synthesis of a solid

material that contains both organic acid and base groups. A sulfonic acid and amine were

incorporated into SBA-15. The immobilized catalyst was tested for activity in the aldol

condensation of 4-nitrobenzaldehyde with acetone at 50 °C (see Figure 22).

Figure 22: Aldol condensation of 4-nitrobenzaldehyde with acetone [45] The material exhibited a remarkable reactivity that is not achievable with the same groups in

solution. The bifunctional catalyst gave a total conversion of 62 % after 20 hours. While the

individual SBA-15 materials, functionalized with either only a sulfonic acid or amine group, gave

significantly lower conversions of 16 %, respectively 33 %. Interesting, a physical mixture of

acid-functionalized SBA-15 and amine-functionalized SBA-15 showed an intermediate level of

conversion, i.e. 44 %, indicating the importance of both functionalities being present at the same

time. With the homogeneous analogues being added separately, only conversions lower than

8 % were achieved as no cooperative effects are possible. When the homogeneous sulfonic acid

and amine were simultaneously added, the reaction showed no conversion because these

functionalities neutralize each other, forming a salt that has no catalytic reactivity [45].

The observed enhancement of reactivity is likely due to interactions between neighbouring acid

and base sites [45]. The equilibrium between acids and bases when immobilized together is not

well defined but likely is very important in understanding the catalysis properties of such

materials. The proposed mechanism for acid/base cooperative catalysis is shown in Figure 23

and demonstrates the potential importance of acid/base interactions [46].

Page 47: Green aldol condensation: synthesis, testing and design of

Introduction 26

Figure 23: Proposed catalytic cycle [46] According to the mechanism in Figure 23, regeneration of the active catalytic species relies on

the rate of proton exchange between the acid and base groups, which would be facilitated by

surface silanols and water when immobilized on mesoporous supports.

Zeidan et al. [46] also investigated the effect of the equilibrium between acid and base groups by

varying the pKa of the acid group. When examining this equilibrium, they reasoned that a careful

choice of acids to pair with the amino group could allow for a change in the equilibrium toward

an increased number of free acid/base sites and potentially lead to higher catalytic activity if less

strong acids with a higher pKa were chosen. Three different acids were immobilized on the SBA-

15 and tested in the aldol condensation of 4-nitrobenzaldehyde with acetone. The results of

these catalytic tests are given in Table 2. In the case of sulfonic acids and amines, the equilibrium

would be expected to be heavily shifted towards the inactive, neutralized ion pair.

Table 2: Results of the catalytic tests [46]

Acid catalyst pKa A [%] B [%] Total conversion [%]

Sulfonic acid -2 45 17 62

Phosphoric acid 3 62 16 78

Carboxylic acid 5 75 24 >99

The conversion increases significantly with decreasing acid strength. The carboxylic acid/amine

material converts the 4-nitrobenzaldehyde nearly quantitatively and gives a constant selectivity

Page 48: Green aldol condensation: synthesis, testing and design of

Introduction 27

throughout the course of the reaction. By proper pairing of weaker acid groups with amine

groups, it is possible to synthesize a catalytic material with excellent conversion toward the

aldol reaction, indicating that the equilibrium is shifted towards the non-neutralized material

and gives rise to enhanced reactivity [46].

In 2012, Brunelli et al. [50] prepared a set of functionalized mesoporous silica materials to

investigate the role of the synthesis method and the impact of the interaction of amines, silanols,

and carboxylic acid groups on the catalytic rate during the aldol condensation of 4-

nitrobenzaldehyde and acetone. They concluded that the synthesis of a carboxylic acid

organosilane is best performed with a protecting group, such as a tert-butyl group that can be

thermally cleaved, to prevent hydrolysis of the ethoxysilyl groups during the synthesis and to

permit grafting onto the surface in the presence of an organic base. The grafting was carried out

in both toluene and methanol. There are indications that methanol as grafting solvent would

promote better spacing upon surface functionalization than toluene [52]. Thus obtaining a more

active catalyst because of an increase in the cooperativity achieved by preventing amine

aggregation on the surface of the silica support.

The reference material (SBA-15 functionalized with 0.5 mmol/g aminopropyl) reached 80 %

conversion after 5 hours. Introduction of the protected carboxylic acid in the grafted materials

decreased the activity of the catalyst to only 50 % conversion after 5 hours. It is assumed that

the second organic group occupied silanol groups adjacent to the amine sites, inhibiting the

beneficial cooperative interaction between amines and silanols. After the material was heated,

the thermal deprotection step introduces a more acidic carboxylic acid site than the silanol it

removes. The activity further decreased to only 35 % conversion after 5 hours. These results

suggest that acidic protons in very close proximity to amines lead to protonated amines that are

significantly less active for the aldol condensation. Brunelli and co-workers concluded that the

acid strength and the proximity determine the equilibrium concentration of inactive amines, and

thus the corresponding activity of the catalyst [50].

Removal of the silanols and the introduction of a carboxylic acid has a negative impact on the

activity of the catalyst. A consensus is forming that weaker acid sites result in the greatest

enhancement of the reaction rate in base-catalyzed coupling reactions such as the aldol

condensation [53]. Of particular importance is the acid-base cooperative catalytic interactions of

amines and silanols that was first demonstrated by Kubota et al. [54] in 2003 for a soluble amine

mixed with a mesoporous silica.

The role of silanols in the reaction can be explained by the fact that carbonyl compounds adsorb

on the surface of the support via hydrogen bonding. NMR measurements suggest that silanol

groups play a key role in bringing all reactants and the catalytic group together for the reaction

Page 49: Green aldol condensation: synthesis, testing and design of

Introduction 28

to take place. Silanol groups assist the catalytic activity of immobilized amines by offering

binding sites for the reactants in close proximity to the amines, providing pathways for proton

transfer throughout all the steps of the reaction, and facilitating the departure of water during

the formation of intermediates (see Figure 24). The formation of the intermediate enamine

involves a series of proton transfers. The mildly acidic silanol groups could assist theses

transfers by aligning with acetone and amine groups in six-membered ring-like arrangements

[47].

Figure 24: Possible pathway of proton transfer assisted by silanol groups [47] More recently, Lauwaert et al. [55] investigated the effect of the silanol-to-amine ratio on the

aldol condensation. As already mentioned, the catalytic activity of amines in aldol condensations

can be improved by incorporating weak acid sites. This promoting effect is in particular

interesting because the silica materials already intrinsically possess weakly acidic silanol

groups. In this work, (3-aminopropyl)triethoxysilane (APTES) was chosen as a precursor for the

active amine groups. Two types of materials were synthesized. Via the grafting of APTES, a

material with both amine and free silanol groups is obtained, leading to a bifunctional catalyst

(type A). For the second type, the bifunctional material is treated with 1,1,1,3,3,3-

hexamethyldisilazane (HMDS). It is assumed that through this treatment with HMDS all the

surface silanols are removed, resulting in a monofunctional catalyst (type B) [55].

The catalysts' performance was tested for the aldol condensation of 4-nitrobenzaldehyde with

acetone at 55 °C. The evolution of the turnover frequencies (TOFs) obtained with the different

catalysts as a function of their silanol-to-amine ratio is shown in Figure 25.

Page 50: Green aldol condensation: synthesis, testing and design of

Introduction 29

Figure 25: TOFs of the catalysts (type A, rhombus) and the HMDS-treated catalysts (type B, square) as a function of the molar silanol-to-amine ratio [55]

Silanol groups are not necessary to catalyse the aldol condensation because they are not an

actual, catalytically active site, but they do act as a promoter to the reaction. This can be seen

from the HMDS-treated catalysts (type B) all exhibiting a similar catalytic activity, while the

TOFs obtained with the bifunctional catalysts (type A) are higher and exhibit an S-shaped

relation with the molar silanol-to-amine ratio. The observed TOF increases rapidly for a molar

silanol-to-amine ratio in the range from 0.5 to 1.1 and reaches a plateau around 1.7. A further

increase in the silanol-to-amine ratio does not lead to increase in the catalytic activity. This

suggests that when an excess of silanol groups are present, no amines are left which are not

promoted [55].

According to the reaction mechanism in Figure 26, the activity of all amines is promoted if all

amines have one silanol as neighbour. This should result in the catalytic activity reaching a

plateau at a molar silanol-to-amine ratio of one. But as can be seen in Figure 25, the plateau is

only reached for a slightly higher silanol-to-amine ratio. This indicates that at a molar silanol-to-

amine ratio of one, not every amine has a silanol next to its location and the actual arrangement

on the surface deviates from an ideal draughtboard pattern. This is explained because hydrogen

bonding can already occur between two amines in the catalyst synthesis mixture. If sufficient

amines are present in the synthesis mixture, they can be considered to move in a clustered

manner during the grafting procedure and will be positioned next to each other on the surface.

This model was confirmed by experimental data [55].

To obtain a better understanding of the heterogeneously catalyzed reaction and in particular,

the surface intermediates that are potentially formed, Lauwaert et al. [55] investigated the

homogeneous reaction between acetone and n-propylamine by using in situ Raman

spectroscopy. Based upon the experimental observations, a reaction mechanism is proposed for

Page 51: Green aldol condensation: synthesis, testing and design of

Introduction 30

the primary-amine-catalyzed aldol condensation in the presence of silanol groups. The

promoting effect can be assigned to the hydrogen-bridge interactions between the carbonyl

moieties in the reactants and the silanol groups. As a result, the reactants become more

susceptible to nucleophilic reactions and the formation of intermediates (1b) and (3b) is

facilitated [55]. The promoting effects of the silanol groups are indicated by (1a) and (3a) in

Figure 26.

Figure 26: Proposed reaction mechanism for the aldol condensation of 4-nitrobenzaldehyde with acetone in the presence of the promoting silanol groups [55]

In 2015, Lauwaert et al. [56] synthesized different acid-base catalysts by functionalizing an

SBA-15 support with (3-iodopropyl)trimethoxysilane. The iodo-group is subsequently replaced

with the desired functional groups, i.e. a secondary amine and an acid site separated by one

carbon atom (see Figure 27). In this work, the effect of the acid strength of the promoting site is

investigated while carefully controlling the arrangement of the sites because the iodosilane

cannot form hydrogen bonds, it does not have any tendency to be grafted in a clustered manner.

The low average site density indicates that the amines will most likely interact with the acid

groups on the same linker and potentially also with surrounding silanols, but not with the acid

groups on a neighbouring linker.

Page 52: Green aldol condensation: synthesis, testing and design of

Introduction 31

Figure 27: Schematic representation of the synthesis of SBA-X with X referring to the functional group, i.e. 'I' for the iodo-group, 'A' for the secondary amine, 'AL' for the alcohol group,

'CA' for the carboxylic acid and 'PA' for the phosphoric acid [56] To investigate the intramolecular acid strength effects on the cooperativity, the activity of the

newly synthesized materials were assessed through aldol condensation experiments with 4-

nitrobenzaldehyde and acetone at 45 °C. Incorporating an intramolecular alcohol on the amine

site increases the TOF, indicating that in this case, the active and promoting sites are in a more

favourable configuration to catalyze the aldol condensation (compared to a structure in which

only intermolecular surface silanols promote the activity of an amine site). This observation can

be attributed to the joint movement of the amine and promoting alcohol site when they are

located on the same linker. Due to this joint movement, the amine and alcohol can bend together

toward a neighbouring silanol on the silica surface, resulting in two promoting sites in the

vicinity of the amine active site, i.e., the intramolecular alcohol function and the intermolecular

surface silanol. Subsequently, the two reactants can simultaneously be activated by the

formation of a hydrogen bond instead of the consecutive activation when there is only one

promoting site in the vicinity of the amine and lead to an increase in catalytic activity.

However, changing the alcohol to a stronger acid site such as a carboxylic acid or a phosphoric

acid decreases the TOF, confirming that stronger acid sites have a negative impact on the activity

of the amines. These results suggest that the optimal promoting sites for the aldol condensation

are H-bond donors, such as alcohols and silanols, and not strong acid sites.

Page 53: Green aldol condensation: synthesis, testing and design of

Introduction 32

Lauwaert and co-workers concluded that materials combining the amine with an alcohol exhibit

the highest TOFs, indicating that bringing an alcohol in close proximity to the amine is beneficial

for the cooperativity between the two sites, on the condition that the inclusion of such

intramolecular alcohol functions does not significantly increase steric hindrance [56].

1.4.2 Effect of the base functionality type

In 2010, Kubota et al. [57] used the same SBA-15 material as silica support for the

immobilization of primary, secondary and tertiary amines. In a post-synthetic modification, the

mesoporous silica SBA-15 was dissolved in toluene and after adding the corresponding

aminoalkyl groups, the mixture was stirred under reflux for 2 hours. The catalytic activity of

these materials was tested in the aldol condensation of 4-nitrobenzaldehyde with acetone at

30 °C. Both the primary (TOF = 4.2 h-1) and secondary (TOF = 26.0 h-1) amine-immobilized

SBA-15 achieved nearly 100 % conversion after 6 hours. The very high activity displayed by the

secondary amine-immobilized SBA-15 material is probably due to the facile formation of

enamine, which is a key reaction intermediate. However, the reaction catalyzed by a tertiary

amine-immobilized SBA-15 resulted in almost no reaction.

This group also investigated the effect of SBA-15 as an additive on the amine-catalyzed aldol

reaction. When only the homogeneous, organic bases were used, the reactions barely reached a

few percentages of conversion after 6 hours. Surprisingly, the activity increased significantly by

simply adding SBA-15 to the secondary amine, forming a physical mixture of the organic base

and the mesoporous silica. For the primary amine, the activity did not change much when

SBA-15 was added; while the use of tertiary amines, either with or without SBA-15, led to almost

no product [57].

Table 3: Effect of SBA-15 as an additive on the amine-catalyzed aldol reaction [57]

Catalyst Additive Time [h] Total conversion [%]

Primary amine None 6 1

SBA-15 6 5

Secondary amine None 6 1

SBA-15 1 26

SBA-15 6 >99

SiO2 1 18

Tertiary amine None 6 0

SBA-15 6 1

None SBA-15 6 0

Page 54: Green aldol condensation: synthesis, testing and design of

Introduction 33

On the other hand, in the absence of any amine functionality SBA-15 was totally inert, which

indicates that high activity is only realized when the reaction is carried out in the presence of

both the amine and silicate additive. Thus, besides the presence or absence of SBA-15, the

activity of this catalytic system is dependent on the type of amine used.

The reaction was also performed with a mixture of the secondary amine and amorphous SiO2.

These results indicate that a periodical structure has advantages over an amorphous structure.

Due to a preferable arrangement and direction of the silanol group on the surface of SBA-15, the

rate-enhancement ability per one surface silanol in an ordered mesoporous silica is higher than

that in amorphous silica. To verify that the silanol group of the SBA-15 material is involved in

the mechanism, end-capping of the silanol groups by trimethylsilylation was carried out. Adding

this mesoporous material to a reaction mixture with a secondary amine caused a significant

decrease in the reaction rate. After 6 hours the conversion was around 30 % (compared to more

than 99 % conversion with the non-trimethylsilylated SBA-15). This indicates the importance of

an acid-base cooperative system [57].

By comparing the results of the amine-SBA-15 mixed system and amine-immobilized SBA-15

after 30 minutes of reaction, the effect of immobilization could be investigated. For the primary

amine, the conversion increased from 3 % in the mixed system to 24 % in the immobilized

system, while for the secondary amine, the conversion increased significantly from 15 % in the

mixed system to 92 % in the immobilized system. This indicates that immobilization results in

much better enhancement of the reactivity than simple physical mixing. Also the reusability of

the amine-immobilized mesoporous silica was tested. Therefore, the mixture was filtered and

the catalyst was washed thoroughly with benzene and recovered. X-Ray Diffraction confirmed

that the framework structures of the catalysts didn’t change after the reaction. The recycled

amine-immobilized materials retained their very high conversions and a constant yield of the

aldol products [57].

In the case of a primary amine-functionalized mesoporous silica nanoparticle (AP-MSN),

Kandel et al. [47] observed an inhibition of the reaction kinetics at high concentrations of 4-

nitrobenzaldehyde. They suggested that the catalytic sites of the material could be blocked by

the formation of a stable Schiff base (see Figure 28), which not only eliminates active sites but

also blocks diffusion in the pores.

Page 55: Green aldol condensation: synthesis, testing and design of

Introduction 34

Figure 28: Formation of a Schiff base between 4-nitrobenzaldehyde and the aminopropyl group of AP-MSN [47]

By comparing the infrared spectra of the functionalized material before and after the reaction,

they confirmed the formation of an imine intermediate, which was stable even after washing and

drying the material. This suggested a chemical transformation of the aminopropyl group rather

than only physisorption of 4-nitrobenzaldehyde to the surface of the particles. Comparison of

the nitrogen content of the material before and after the reaction by elemental analysis revealed

that approximately 70 % of amine groups formed the imine (the amine loading varied from 1.0

mmol/g before reaction to 1.7 mmol/g after the formation of the stable intermediate).

In 2014, Lauwaert et al. [55] proposed a reaction mechanism for the primary-amine-catalysed

aldol condensation between 4-nitrobenzaldehyde and acetone. Figure 29 represents a further

elaborated version of this reaction mechanism which was used to determine the effect of the

amine structure and base strength [58].

Figure 29: Proposed reaction mechanism for the aldol condensation of 4-nitrobenzaldehyde with acetone [58]

Page 56: Green aldol condensation: synthesis, testing and design of

Introduction 35

When no promoting silanol groups are available, the first step of the reaction mechanism is the

formation of an enamine (1) through a nucleophilic addition of acetone to the amine active site.

This happens in two steps: first a carbinolamine is formed which, subsequently, dehydrates with

formation of the enamine. In the case of a primary amine the carbinolamine can also dehydrate

with formation of an imine (3) which inhibits the further reaction (see Figure 30). For the actual

aldol condensation to take place, the enamine reacts with 4-nitrobenzaldehyde yielding an

iminium ion (2). The primary reaction product (aldol) is formed by a water-assisted desorption

of this iminium ion. Finally, the aldol product releases water and forms the secondary reaction

product (ketone). In the presence of promoting silanol groups, the nucleophilic reactions will be

facilitated due to the formation of hydrogen-bridge interactions between carbonyl moieties and

the silanol groups, represented by the species indicated by (1a), (2a), (3a) and (4a) [58].

Figure 30: Nucleophilic addition of acetone to a primary amine [58] In the present work an acetone excess has been used. As a result, the direct interaction between

4-nitrobenzaldehyde and the primary amine is suppressed. Therefore, the formation of the

inhibiting imine (3) from 4-nitrobenzaldehyde can be considered limited. In order to investigate

the effect of the catalyst properties on the aldol condensation, all catalysts were subject to

identical operating conditions. As can been seen in Table 4, pronounced differences in turnover

frequency are observed as a function of the base that was grafted on the silica.

Page 57: Green aldol condensation: synthesis, testing and design of

Introduction 36

Table 4: Turnover frequencies obtained with both the unpromoted base as the corresponding cooperative acid-base catalysts [56, 58]

Amine active site Type of amine TOFunpromoted (s-1) TOFpromoted (s-1)

Aminopropane Primary 2.0 *10-4 7.80 *10-4

Methylaminopropane Secondary 1.0 *10-3 3.30 *10-3

Ethylaminopropane Secondary 3.61*10-5 1.41*10-4

Cyclohexylaminopropane Secondary 9.6 *10-6 3.1 *10-5

Phenylaminopropane Secondary - 7.0 *10-6

Diethylaminopropane Tertiary 1.2 *10-5 3.9 *10-5

Primary amines exhibit a reasonable turnover frequency. While both cyclohexyl and phenyl

substituents lead to lower values of the turnover frequency; methylaminopropane resulted in

the highest turnover frequencies, clearly indicating the important effect of the substituent on

secondary amines. In case of a secondary amine the formation of any imine is prevented due to

the lack of a second hydrogen on the nitrogen, resulting in a higher reaction rate for this type of

amine. By changing the substituent of a secondary amine from a methyl to an ethyl group, the

TOF is lowered with about one order of magnitude. This suggest that steric hindrance matters

even for relatively small substituents. With tertiary amines extremely low TOF values were

observed. Because for tertiary amines, the formation of the carbinolamine intermediate is no

longer possible due to the absence of hydrogen atoms on the nitrogen atom, explaining the low

activity of the tertiary amine [56, 58]

The data reported in Table 4 clearly show how the activity exhibited by amine groups is

enhanced by the presence of silanols, which again confirms the cooperativity between the acid

and base sites. In their previous work Lauwaert and co-workers [55] showed that this

cooperative effect is also depended on the arrangement of the active site on the catalyst surface.

As already explained, primary amines tend to be positioned in a clustered manner on the silica

surface due to hydrogen bonding in the synthesis mixture. On the other hand, the precursor

leading to a secondary amine doesn't have this tendency and shows no significant deviations

from ideality. Leading to a secondary amine being grafted in a random manner on the silica

surface. As we can see in Figure 31, this results in a higher percentage of promoted secondary

amines in the silanol-to-amine ratio of 0 to 1.7 in comparison to the primary amines.

Page 58: Green aldol condensation: synthesis, testing and design of

Introduction 37

Figure 31: Percentage of amines promoted as a function of the silanol-to-amine ratio; Diamonds: catalysts functionalized with primary amine;

Squares: catalysts functionalized with secondary amine [58] At high conversions, a selectivity towards the secondary ketone product of about 16 % is

obtained with the aminopropane active site while a selectivity of about 10 % is obtained with

the methylaminopropane active site. Hence, the nature of the active site seems to influence the

obtained product distribution.

Although many primary, secondary and tertiary amines have similar base strengths, these

amines exhibit significant differences in catalytic activities for the aldol condensation.

Cyclohexylaminopropane has the highest base strength and it is unable to form the inhibiting

imine, but the active site showed a low activity. This can be explained by steric hindrance

induced by the bulky cyclohexyl group close to the amine, increasing the activation energy of the

reaction. They concluded that a secondary amine with a small substituent avoiding steric

hindrance seems to be the optimal amine type to catalyze the aldol condensation between 4-

nitrobenzaldehyde and acetone [58].

Page 59: Green aldol condensation: synthesis, testing and design of

Introduction 38

1.4.3 Effect of the distance between the base and acid functionality

Although efficient dual activation promotes the reaction, the acid and base catalytic components

can also interact in a non-productive manner based on their relative strength and proximity,

quenching the activating behaviour of the individual sites; and rendering the catalyst inactive if

the acid and base sites strongly self-associate [50, 59].

Yu et al. [49] reported the synthesis of bifunctional catalysts with controlled spatial arrangement

of the functional groups by controlling the steric hindrance. The authors tried to immobilize

different amine groups of lysine on mesoporous silica. The acid-base distance can be adjusted

within a range of one to five carbon atoms, to prevent the mutually destruction of the acidic site

and basic site. This resulted in two materials: proximal-C-A-SBA-15 and maximum-C-A-SBA-15,

which are shown in Figure 32.

Figure 32: Proximal-C-A-SBA-15 and maximum-C-A-SBA-15 [49] The synthesized materials were tested in the aldol condensation of 4-nitrobenzaldehyde with

acetone at 50 °C and their catalytic properties are shown in Table 5.

Table 5: Catalytic properties [49]

Catalyst a A [%] B [%] Total conversion b [%] TON c

Proximal-C-A-SBA-15 71 19 90 7.4

Maximum-C-A-SBA-15 57 17 74 6.0

a 0.5 mmol 4-nitrobenzaldehyde in 10 mL acetone stirred at 50°C for 20h. b Conversion calculated by 1H

NMR spectroscopy analysis. c Number of moles of aldehyde converted per 1 mol of -COOH.

While containing almost the same amount of lysine, proximal-C-A-SBA-15 showed a higher TON

value than maximum-C-A-SBA-15, indicating that the rate per active site was enhanced by the

proximal acid-base distance. The activity of the catalysts decreased as the acid-base distance

increased, suggesting that cooperative surface catalysis relies on the two functional groups

being close enough to each other on the surface of the mesoporous material to interact with

each other and with the reacting molecules. A cooperative mechanism was suggested by Yu and

Page 60: Green aldol condensation: synthesis, testing and design of

Introduction 39

coworkers. In Figure 33, the dual activation of electrophiles and nucleophiles at the acidic sites

and the basic sites is elaborated. The aldehyde gets activated by the acidic sites, and acetone is

deprotonated to generate imine formation by the basic sites, which then attacks the electron-

deficient aldehyde to generate the aldol product [49]. Compared to the mechanism proposed by

Zeidan and co-workers (Figure 23) the latter seems to be too simplistic, given the new

information. By including possible surface intermediates and cooperative interactions, this

mechanism is more comprehensive, more complete. Note that in Figure 33 the molecule

assigned with the asterisk is missing three carbon atoms to be correct.

Figure 33: Proposed reaction mechanism for the aldol condensation reaction on the solid support [49]

Page 61: Green aldol condensation: synthesis, testing and design of

Introduction 40

While the importance of tuning the acid strength has been demonstrated, the impact of the

length and flexibility of amines in their cooperative interactions with the silanols had not yet

been studied. Brunelli et al. [53] investigated the effect of the spatial separation of the amines

and silanols by systematically varying the alkyl chain length of the organosilane grafted onto a

well-defined SBA-15 mesoporous silica support. Using a set of materials functionalized with

alkyl linkers varying from methyl to pentyl (see Figure 34), they revealed that the cooperativity

between the amine and silanols could be tuned by controlling the linker length [53, 60].

Figure 34: Aminosilanes grafted onto SBA-15 with controlled linker lengths [53] The materials were tested for their catalytic efficiency in the aldol condensation of 4-nitro-

benzaldehyde with acetone at 50 °C. The linker length had a significant impact on the observed

reactions rates, as can be seen in Figure 35. The observed rate increased with the number of

carbons up to the propyl chain. Increasing the linker length to butyl or pentyl, resulted in

identical catalytic rates, suggesting that amine-silanol quenching reactions do not occur and that

a propyl linker is sufficiently long to allow for effective cooperative reactions [53].

Figure 35: Conversion of 4-nitrobenzaldehyde in the aldol condensation with acetone at 50 °C for the catalysts with different alkyl linkers [53]

Page 62: Green aldol condensation: synthesis, testing and design of

Introduction 41

The material with the shortest chain (C1) showed an interesting behaviour. The conversion

rapidly increased to 10 % within 1 hour, after which the conversion increased slightly over time.

This suggests a relatively rapid reaction of the aldehyde with the amine sites on the silica

surface. But the single methylene linker does not permit sufficient flexibility of the amine site

to allow the reaction to proceed through a silanol-activated pathway. Instead, the reaction

proceeds much slower, as it would for a base-only catalyzed reaction at a rate comparable with

homogeneous 3-propylamine [53, 60].

With an intermediate linker length (C2), the linker enabled a portion of the amines to act

cooperatively with the silanol groups. The increase in catalytic activity with the longer linkers

suggests that a certain degree of flexibility of the alkyl linker is required before the amine and

silanols can efficiently cooperate [60].

These results are in contrast with the work of Yu et al. [49] for the heterogeneous carboxylic

acid and amine system, where the additional flexibility associated with longer chains between

the functional groups resulted in decreased activity. This difference in behaviour may suggest

that the silanols on the surface of the silica are sufficiently weak (pKa around 7.1), such that they

do not protonate the amines, unlike the stronger carboxylic acid (pKa around 5). Therefore,

silanols may behave more like a hydrogen-bonding partner than an acid in these catalysts [60].

Another important aspect of the amine-silanol cooperativity is the curvature of the surface. With

an average carbon-carbon bond length of 0.154 nm, the aminosilanes will approximately reach

0.5 nm into the pore. This means that for the SBA-15, with a pore diameter of 6.5 nm, the surface

appears relatively flat on the length scale that the aminosilane can extend. So by changing the

pore diameter, a significant change in the cooperativity between the amine and silanol groups

should be expected. Brunelli et al. [60] synthesized a mesoporous material with a much smaller

pore diameter (2.3 nm) and noticed that the catalytic cooperativity of the intermediate linker

(C2) increased when placed inside the small pore diameter material and was equivalent to that

of the C3 linker in the larger pore diameter catalyst. Changing the curvature of the pores brought

the aminosilane into closer proximity with the surface silanols, changing the flexibility

requirements of the linker to achieve high catalytic activity [60].

Page 63: Green aldol condensation: synthesis, testing and design of

Introduction 42

1.4.4 Effect of the reaction medium

It is well-known that the selection of solvents can have an important effect on the rates of

reactions. Such an effect is explained by the contribution of solvation energy to the total free

energy of the systems and by stabilization of the transition states with a subsequent reduction in

the free energy of activation. The solvent effects can change preferences for various possible

pathways over the potential energy surface of the reaction system [48].

Zeidan et al. [45] suggested that the acid and base groups in the mesoporous material would be

in equilibrium between the free acid and base and the ion pair that results from neutralization;

and that the solvent in which the reaction is performed, would have an important effect on this

equilibrium. To examine the effect of solvent polarity on conversion, the reactions were carried

out in a variety of co-solvents of which the results are summarized in Table 6. In polar, protic

solvents, such as water and methanol, the equilibrium lies towards the ion pair because proton

exchange is rapid and the protic solvent will stabilize the ion pair the most. Leading to only a

small conversion around 30 % after 20 hours. The conversion more than doubles to about 70 %

upon moving to polar, aprotic solvents, such as diethyl ether and chloroform. Non-polar, aprotic

solvents, such as hexane and benzene, cause slower exchange of the protons, thus moving the

equilibrium in favour of the free acid and base; giving further improvements in conversion to

about 90 % in the case of hexane.

Table 6: Solvent effect on the catalysis [45]

Cosolvent (1:1) A [%] B [%] Total conversion [%]

Water 27 7 34

Methanol 14 9 23

Diethyl ether 52 13 65

Chloroform 60 9 69

Hexane 75 13 88

Benzene 62 12 74

Mesoporous silica nanoparticles (MSNs) functionalized with primary amines are poor catalysts

for the aldol reaction in hexane because of the formation of an imine intermediate. This low

activity of AP-MSN was overcome by using a secondary amine-functionalized material (MAP-

MSN) where the methyl group on the aminopropyl prevented the formation of imine and gave a

3-fold increase in the apparent rate constant in hexane [48].

Page 64: Green aldol condensation: synthesis, testing and design of

Introduction 43

However, switching the reaction medium between hexane and water led to a reversal in

activities between the primary and secondary amine-functionalized material. These differences

in behaviour are not the result from the solvents directing each catalytic reaction through

different pathways. In order to provide a more in-depth explanation, Kandel et al. [48]

investigated the different effects of the solvent properties on the catalytic activity.

Corresponding to the findings by Zeidan and co-workers [45], the polarity of the solvent had a

negative effect on the activity of bifunctionalized materials. As polarity increases, proton

transfer between the acidic silanols and the basic amine can take place, reducing the availability

of the deprotonated amine required to perform enamine catalysis. Because the reaction

catalyzed by AP-MSN in water doesn't follow this trend, the enhanced activity of the catalyst in

this solvent cannot be due to polarity.

The solvent can also have an effect on the different equilibria involved in the reaction. Water can

be considered as a reagent in the last step (part 3b of Figure 26), where it combines with the

enamine intermediate to form the aldol product. Thus, in the overall conversion (as shown in

Figure 36), water is a product only in the inhibition route (red) but is not part of the net

reactants or products of the aldol route (black). Using water as a the reaction solvent, shifts the

overall equilibrium toward the formation of the aldol product and minimizes the inhibition

pathway. Kandel and co-workers concluded that the catalytic activity of both functionalized

materials in water should be a balance between the inhibitory effects of polarity and the

promoting effect of the solvent on equilibria [48].

Figure 36: Overall reaction pathways for 4-nitrobenzaldehyde and acetone in the presence of AP-MSN: inhibition (red) and aldol reaction (black) [48]

Nonetheless, it has to be mentioned that in the above conclusion it was assumed that the aldol is

the main product of the reaction. If the most abundant product were the enone or enal, water

would be formed as a by-product of the reaction and as a consequence would play no role at all

in shifting the equilibrium. However, analysis of the product distribution revealed that the aldol

was the major product for the reactions with selectivities around 80 % to 90 %.

Page 65: Green aldol condensation: synthesis, testing and design of

Introduction 44

1.4.5 Effect of the hydrophobicity of the support

In 2014, Zhe An et al. [61] synthesized mesoporous silicas with alternating hydrophobic and

hydrophilic blocks in the pore wall by tailoring the molar ratio of 1,4-bis(triethoxysilyl)benzene

(BTEB) and tetraethylsilicate (TEOS). Both materials also contained 10 mol% of 3-

mercaptopropyl-trimethoxysilane (MPTMS). The materials were abbreviated as HHB-B8T2 and

HHB-B5T5. For BxTy, x/y represents the molar ratio of BTEB and TEOS. To compare the

catalytic activity of these materials, they also prepared mesoporous silicas with a hydrophobic

surface (containing only BTEB and MPTMS) and mesoporous silicas with a hydrophilic surface

(containing only TEOS and MPTMS). The mercaptopropyl group was later used to functionalize

these materials with L-proline (using chloromethylstyrene as the linker) via a thiol-ene click

reaction (see Chapter Materials & methods).

The catalytic asymmetric aldol reaction of 4-nitrobenzaldehyde with cyclohexanone at 25 °C was

performed. When cyclohexanone was used as solvent, the mesoporous silica with hydrophobic

surface gave a 58 % yield in 24 hours. With HHB-B8T2-Pro and HHB-B5T5-Pro the reaction

resulted in a yield of respectively 42 % and 27 %. For the material with hydrophilic surface, no

reaction was observed (< 3 % in 48 hours). The catalytic reaction was then performed in water.

For the hydrophobic surface, the yield increased to 96 % in 24 hours. HHB-B8T2-Pro and HHB-

B5T5-Pro respectively gave 83 % and 48 % yield. Also the hydrophilic surface afforded a higher

yield of 18 % in 24 hours.

Figure 37: Presentation of a mesoporous silica with alternating hydrophobic and hydrophilic blocks in the pore wall and the different interfaces inside the channels [61]

Therefore, the catalytic activity relies both on the surface properties of the catalyst solids as well

as on the reaction medium. The superior activity of hydrophobic or hydrophobic and hydrophilic

Page 66: Green aldol condensation: synthesis, testing and design of

Introduction 45

alternating catalyst to hydrophilic catalyst can be attributed to the synergistic effects of the

hydrophobic blocks in their channels. In neat condition, these channels advanced the access of

organic reactants to the catalytic sites, hereby accelerating the reaction. In aqueous medium, see

Figure 37, the mesoporous channels were able to attract the organic reactants from the water

and constrain them, increasing the reactant concentration around the catalytic sites [61].

1.5 Scope of the thesis

The aldol condensation of furfural with acetone will be performed in cooperation with the thesis

of Elien Laforce. In comparison with 4-nitrobenzaldehyde, furfural doesn’t contain an electron-

withdrawing group, resulting in a lower reaction rate. Higher reaction temperatures are

required to obtain significant conversions. In industry, the product stream of furfural contains

water. Therefore the hydrophobicity of the catalyst support will be tuned during the synthesis.

Recent research indicated that periodic mesoporous organosilicas (PMOs) have a greater

activity in the aldol condensation than functionalized silica materials. In this thesis, a series of

PMO materials will be synthesized using vinyltriethoxysilane (VTES), tetraethyl orthosilicate

(TEOS), 1,2-bis(triethoxysilyl)ethane (BTEE) and 1,4-bis(triethoxysilyl)benzene (BTEB) as

precursors. Different ratios of these precursors are added during the synthesis to vary the

hydrophobicity of the support.

Afterwards, the vinyl group of VTES will be used to functionalize the PMO materials with

cysteine via a thiol-ene click reaction. After sufficient washing procedures to remove any

remainings of physisorpted cysteine, the functionalized PMO materials are characterized via

elemental analysis to determine the amine loading. The mesoporous structure is verified via

nitrogen physisorption and X-Ray diffraction.

The aldol condensation of 4-nitrobenzaldehyde with acetone will be performed to compare the

catalytic activity of these bifunctional materials. The reaction will also be performed in the

presence of water to investigate the influence of the hydrophobicity of the support on the

catalyst's activity. The used catalyst will be recovered to test the reusability of these

heterogeneous catalysts.

Page 67: Green aldol condensation: synthesis, testing and design of

Introduction 46

Page 68: Green aldol condensation: synthesis, testing and design of

Materials and methods 47

Chapter 2

Materials and methods

This chapter describes the techniques and procedures which were employed for the

characterization of the synthesized Periodic Mesoporous Organosilicas and the analysis of the

catalytic experiments. Also the specifications of the used devices will be mentioned.

2.1 Materials

1,2-bis(triethoxysilyl)ethane (abcr, 97 %), 1,4-bis(triethoxy-silyl)benzene (abcr, 95 %),

2-Hydroxy-4-(2-hydroxyethoxy)-2-methylpropiophenone (Irgacure 2959, Sigma Aldrich, 98 %),

4-nitrobenzaldehyde (Acros Organics, 99 %), Acetone (Acros, 99.6 %, ACS reagent), Acetone

(Chem-Lab NV, 99+%), Acetonitrile (Acros, for HPLC), Brij S10 (Sigma Aldrich, Mn ~711),

Disolol (Chem-Lab NV, 97 %, 3 % IPA), Ethanol (Fiers, 96 %, 2 % MEK + 2 % IPA), Furfural

(Sigma Aldrich, 99 %), KCl (Roth, >99.5 %), L-cysteine (Acros Organics, 99+ %), methyl 4-

nitrobenzoate (Sigma Aldrich, >99 %), n-Hexane (Acros Organics, pure), Tetraethyl orthosilicate

(abcr, 98 %), Toluene (Acros Organics, 99.8 %), Vinyltriethoxysilane (abcr, 98 %).

2.2 Elemental Analysis

Elemental analysis, also called CHNS-analysis, is a qualitative and quantitative analysis which

allows the determination of the composition (in percentages) of the elements carbon, hydrogen,

nitrogen and sulphur. This measurement provides absolute values with a very high accuracy and

precision for the main elements in organic and inorganic compounds.

First the sample is weighed into a tin bowl and is placed in a combustion furnace with an excess

of oxygen. This very high concentration of oxygen is needed for the combustion of the sample.

During the combustion, the elements which have to be analyzed are converted into gaseous

components: CO2, H2O, NOx and SO2. With the aid of the carrier gas, helium, the gas mixture is

transported to a gas chromatograph. Here, the gases are separated from each other and the

detection of the individual components can take place [62].

The elemental analysis was realized by means of a Thermo Flash 2000 device. During the

measurement, V2O5 was added as a catalyst.

Page 69: Green aldol condensation: synthesis, testing and design of

Materials and methods 48

2.3 Nitrogen physisorption

Nitrogen physisorption is a technique to determine the specific surface area, the pore size, the

pore volume and pore size distribution. But it also gives insights into the shape of the pores.

An amount of nitrogen gas adsorbs at 77 K, a constant temperature at which the adsorption is

only dependent on the pressure and on the interaction between the adsorbent and the

adsorbate. Based upon the strength of this interaction, all adsorption processes can be divided

into two categories, namely physisorption and chemisorption. If the cross-section of the

adsorbed gas molecule is known and one can determine how much gas is needed to obtain a

monolayer on the surface of the sample, then one can quite easily calculate the surface area [29,

32, 63]. Plotting the amount of adsorbed nitrogen gas versus the relative pressure result in

adsorption isotherms which, according to Brunauer, Deming, Deming and Teller (BDDT), can be

divided into five types, as shown in Figure 38.

Figure 38: Different types of adsorption isotherms [63] Type I isotherms are encountered when adsorption is limited to one molecular layer. This

condition is fulfilled when chemisorption or physisorption at microporous adsorbentia occurs.

This adsorption isotherm is also often referred to as the Langmuir isotherm. Type II isotherms

are most frequently encountered when adsorption occurs on non-porous or macroporous

materials. The inflection point of the isotherm (point B) usually occurs near the completion of

the first adsorbed monolayer. With increasing relative pressure, second and higher layers are

completed until, at saturation, the number of adsorbed layers becomes infinite. Type III and

Type V isotherms are a variation on type II and type IV respectively. The adsorption is facilitated

by interactions with pre-adsorbed molecules, and hence, multilayer adsorption occurs before

monolayer adsorption is complete.

Page 70: Green aldol condensation: synthesis, testing and design of

Materials and methods 49

Type IV isotherms apply to mesoporous materials. The isotherm can be divided into different

regions, as can be seen in Figure 39. First, at low relative pressures, the micropores are filled

with nitrogen. From the location of point A, the micropore volume can be determined. With

increasing relative pressure, nitrogen molecules are starting to adsorb on top of each other.

During the formation of these multilayers, the amount of adsorbed gas is more or less linearly

related to the increase of the relative pressure. This region (point A-B) of the adsorption

isotherm will be used to calculate the specific surface area via the BET theory. At even higher

relative pressures, the adsorbed molecules stack until the nitrogen molecules will condense. The

result of this capillary condensation (point B-C) is a strong increase in the amount of adsorbed

gas. The slope of this increase determines the pore size distribution. The steeper the increase,

the more pores with similar diameters are filled at the same time and thus is an indication for

the uniformity of the material. When the capillary condensation is completed, the pores are

completely filled with liquid adsorbate, so that no further adsorption is possible, and we will see

a plateau in the isotherm. The height of this plateau determines the total pore volume. A typical

hysteresis can be seen: the desorption of the adsorbate occurs at lower relative pressures than

the adsorption. This is the consequence of a difference in condensation and desorption enthalpy.

The form of the hysteresis is very important, as it predicts the shape of the pores [32, 63].

Figure 39: Type IV isotherm of SBA-15 [32] In order to determine the specific surface area of a porous material, the BET theory (Brunauer,

Emmett and Teller) and the following BET-equation [64] are used:

𝑧

(1 − 𝑧)𝑉=

1

𝑐𝑉𝑚𝑜𝑛+

(𝑐 − 1)𝑧

𝑐𝑉𝑚𝑜𝑛

Page 71: Green aldol condensation: synthesis, testing and design of

Materials and methods 50

in which,

z (= p/p°) is the relative pressure;

V is the adsorbed volume at pressure p;

Vmon is the volume corresponding to monolayer adsorption;

C is a BET constant.

When the left hand side of the equation is plotted as a function of the relative pressure (in the

range 0.05 < p/p° < 0.35) a straight line is obtained. The volume corresponding to monolayer

adsorption can be determined from the slope and the intercept of this straight line. If one

assumes that the adsorbate is an ideal gas, the number of moles of adsorbate at monolayer

adsorption can be obtained from:

𝑚𝑚𝑜𝑛 = 𝑉𝑚𝑜𝑛

22400

and with this value, the specific surface area, expressed per gram of adsorbent, can be

calculated:

𝑆𝐵𝐸𝑇 = 𝑚𝑚𝑜𝑛

𝑎 𝑁𝐴 𝐴

in which,

a is the mass of the adsorbent;

NA is the number of Avogadro;

A is the cross-section of one adsorbate molecule (for N2: 0.162*10-18 m²).

The nitrogen adsorption and desorption measurements were performed at 77K on a Tristar

3000 of Micromeritics. Samples were dried at 100°C under vacuum overnight.

2.4 X-Ray Diffraction

X-Ray Diffraction (XRD) is a versatile and non-destructive technique to quickly obtain detailed

structural and phase information of materials. When these X-rays (0.01 - 10 nm) hit the sample

they are scattered in all directions. However, if the material possesses an ordered structure, a

portion of the X-rays can be reflected and undergo constructive interference [62]. This leads to

Bragg's law, which describes the condition on the scattering angle θ for the constructive

interference to be at its strongest:

𝑛𝜆 = 2𝑑 sin 𝛳

Page 72: Green aldol condensation: synthesis, testing and design of

Materials and methods 51

In which,

λ is the wavelength of the X-rays;

d is the interplanar distance between two lattice planes;

θ is the scattering angle;

n is an integer called the order of the reflection.

Figure 40: Diffraction of X-rays [65] A selected wavelength from a polychromatic beam of X-rays will only be reflected for a specific

angle. By varying the detection system at an angle of 2θ, various wavelengths can be examined.

The measured intensity of the reflected X-rays is plotted as a function of twice the angle of

incidence, 2θ. The obtained diffractograms are characteristic for certain structures of the

material.

Figure 41: Scheme of the geometry of the XRD device [66] For hexagonal structures, the XRD data can be used to determine the distance between two

adjacent pores, also known as the width of the unit cell a0. And, as can be seen in Figure 42,

Page 73: Green aldol condensation: synthesis, testing and design of

Materials and methods 52

in combination with the pore diameter (dp) and the distance between two lattice surfaces (d100),

the wall thickness of the material (e) can be calculated using the following formula.

𝑎0 =2𝑑100

√3

𝑒 = 𝑎0 − 𝑑𝑝

Figure 42: Representation of a hexagonal (P6mm) structure The X-ray diffraction patterns were recorded on a ARL X'TRA Diffractometer of Thermo

Scientific which is equipped with a Cu Kα tube and a Peltier cooled lithium drifted silicon solid

stage detector.

2.5 Diffuse Reflective Infrared Fourier Transform

(DRIFT) Spectroscopy

Infrared spectroscopy is primarily used for structure identification. By radiating light with

appropriate energy into a sample, the molecule will absorb a certain amount of energy and a

transition between two energy levels can be induced. The precise energy difference between

each of these levels is function of the type of chemical bond. Thus, knowledge of these

differences will therefore lead to identification of functional groups. The irradiated infrared rays

can be either absorbed or reflected. The reflection can occur in two different ways: specular

reflectance or diffuse reflectance. With DRIFT, this last type of reflection is used, wherein the

radiation enters into a theoretically infinitely thin layer, and then reflects on a few particles in

the sample. In order to increase the fraction of diffuse reflectance, the sample is finely pulverized

into smaller particles of approximately five micrometers. If the sample absorbs the infrared rays

too much, potassium bromide (KBr) is added to the sample as a non-absorbent matrix [67].

Page 74: Green aldol condensation: synthesis, testing and design of

Materials and methods 53

Diffuse Reflective Infrared Fourier Transform (DRIFT) spectroscopy was performed on a hybrid

IR-RAMAN spectrophotometer, namely Nicolet 6700 (Thermo Scientific) with a MCT-A detector

which gets cooled with liquid nitrogen and uses a Graseby Specac diffuse reflective cell.

Measurements were performed under vacuum at 120 °C.

2.6 Aldol condensation of furfural with acetone

The experiments on furfural were performed by Elien Laforce. Silica gel 60 was grafted with a

methyl substituted secondary amine. It was mentioned in the literature study that the

length of the linker has an influence on the catalytic activity. Therefore, N-methylamino-

propyltrimethoxysilane (MAPTMS) was chosen as precursor because Brunelli et al. [53] proved

that a propyl chain leads to the best activity of silanol assisted amine catalysts. After calcination

at 700 °C, to remove the remaining surfactant, a silica material with approximately 1.1 silanol

groups per nm² is obtained. Next, this material is dissolved in dry toluene (99.8 %, Acros

Organics) and MAPTMS is added. The mixture is heated to 120 °C and stirred during 24 hours

under inert atmosphere. By varying the amount of precursor, catalysts with different silanol-to-

amine ratio are obtained. Following the grafting procedure, the catalyst is washed with

chloroform to remove any remainings of physisorpted organic components. Finally, the

bifunctional catalyst are dried at room temperature, under vacuum. The monofunctional

catalysts are prepared by treating the bifunctional catalysts with hexamethyldisilazane

(abcr, 98 %) after which the free silanol groups are covered with trimethylsilane and cannot

longer play a promoting role in the aldol condensation.

The samples of the aldol condensation of furfural and acetone are analysed with a gas

chromatograph. Toluene is used as internal standard. However, no hexane was added as co-

solvent because this would lead to a phase separation with furfural. In the reference reaction,

hexane is added as co-solvent because non-polar, aprotic solvents cause slower exchange of the

protons. Thus moving the equilibrium in favour of the free acid and base, resulting in higher

conversions [45].

2.7 Aldol condensation of 4-nitrobenzaldehyde with

acetone

The activity of each catalyst was assessed via the aldol condensation of 4-nitrobenzaldehyde and

acetone at 45 °C. An acetone excess is used in all experiments in order to suppress the direct

interaction between 4-nitrobenzaldehyde and the amine site which yields an inhibiting imine.

Hexane was added as co-solvent to improve the reaction rate.

Page 75: Green aldol condensation: synthesis, testing and design of

Materials and methods 54

The experiments were performed in a 25 mL two-neck round-bottom flask equipped with a

condenser and a septum. First, the functionalized catalyst was added to the flask. To compare

the activities, 4 mol% with respect to the 4-nitrobenzaldehyde concentration was added to

every flask. The reaction mixture was prepared separately by mixing the desired amounts of

acetone (50 vol%), n-hexane (co-solvent, 50 vol%), 4-nitrobenzaldehyde (0.03 mmol/mL) and

methyl 4-nitrobenzoate (internal standard, 0.022 mmol/mL). Afterwards 4 to 8 mL of the

reaction mixture (depending on the available quantity of catalyst) was injected into the flask

which contains the catalyst and the flask was immediately placed in an oil bath at 45 °C. The

moment the flask was placed in the oil bath was taken as the start of the reaction. The reaction

was monitored for 4 hours by taking a sample every 30 minutes (about 100 µL) of the reaction

mixture. Typically 0.9 mL of acetone was used to wash the syringe needle and to transfer the

sample to a vial. Afterwards the catalyst was separated from the sample by means of

centrifugation. Finally, the samples were analyzed using a reversed-phase high-performance

liquid chromatograph (RP-HPLC), from Agilent (1100 series). The HPLC was operated at a

column temperature of 30°C using a gradient method with water (0.1% Trifluoroacetic acid) and

acetonitrile (HPLC grade) as solvents. In this gradient method the volumetric percentage of

acetonitrile is varied from 30% to 62% over a period of 7 minutes. The components were

identified using a UV-detector with a variable wavelength. Quantification of the different

components in the reaction mixture was performed by relating the peak surface areas to the

amount of internal standard, methyl 4-nitrobenzoate [68].

2.8 High Performance Liquid Chromatography (HPLC)

High performance liquid chromatography is a technique used to separate, identify and quantify

the components in a mixture. The mobile phase is a liquid and high pressures are needed to

guide the eluent through the column. In normal phase chromatography, the column is packed

with a hydrophilic, polar stationary phase. In reversed phase chromatography, the column is

packed with a hydrophobic, non-polar stationary phase [62]. The latter was used for the

characterization of the reaction mixture. Separation results from adsorption of hydrophobic

molecules onto the hydrophobic solid support in a polar mobile phase. A mixture of solvents,

water (0.1 % trifluoroacetic acid) and acetonitrile, is used as mobile phase with varying

composition in time to increase the efficiency of the separation. Decreasing the mobile phase

polarity by adding more organic solvent, reduces the hydrophobic interaction between the

solute and the solid support, resulting in desorption. Depending on their interaction with the

stationary phase, the compounds flowing through the column each have a specific retention time

(see Table 7) . The fractions coming off the column are detected and quantified with the use of a

Page 76: Green aldol condensation: synthesis, testing and design of

Materials and methods 55

UV detector which measures the difference between the intensity of the incident light, 𝐼0, and the

light transmitted after the detector cell, 𝐼. The intensity of absorption is measured in terms of

absorbance, A, which is assumed to be proportional to the concentration of the analyte, c, and

the optical path length, L. This is combined into the Lambert-Beer law:

𝐴 = log (𝐼0

𝐼) = 𝜀 ∗ 𝐿 ∗ 𝑐

In which,

ε is a constant known as the molar absorptivity at a given wavelength.

The detection sensitivity of a component can be increased by varying the wavelength of the UV

detector. Every component has a characteristic wavelength which will be maximal absorbed.

Quantification of the different components is the reaction mixture was performed by relating the

peak surface areas to the amount of internal standard. Calibration factors are determined by

measuring calibration curves of known mixtures of the different components with the internal

standard. These calibration factors are related to the sensitivity of the detector, and thus are

dependent on the used wavelength. These wavelengths for maximal absorption and the

corresponding calibration factors can be found in Table 7. Samples of the aldol condensation of

4-nitrobenzaldehyde with acetone were analysed on a RP-HPLC, Agilent 1100 Series.

Table 7: Components from the reaction mixture with their retention times, wavelength for maximal absorption and calibration factors

Component Retention time

(min)

Wavelength

(nm)

Calibration

factor

Acetone 1.9 265 759.31

4-hydroxy-4-(p-nitrophenyl)-butan-2-one 4.2 272 1.2936

4-nitrobenzaldehyde 5.9 265 0.9711

4-(p-nitrophenyl)3-butene-2-one 7.1 305 0.5349

4-methylbenzoate 7.8 260 1

2.9 NRTL-HOC code

An existing code based on the Non-Random Two-Liquid (NRTL) and Hayden-O’Connell (HOC)

methods has been modified which allows for an initial estimation of the vapour-liquid

equilibrium. This code accounts for the thermodynamic non-ideality of both the liquid as well as

the gas phase via so-called activity coefficients and fugacity coefficients. The activity of

component 𝑖 in a liquid mixture is calculated from the mole fraction of this component in the

Page 77: Green aldol condensation: synthesis, testing and design of

Materials and methods 56

mixture and an activity coefficient, 𝛾𝑖 . The fugacity of a component in a gas mixture is calculated

from its partial pressure and a fugacity coefficient, 𝜙𝑖 [69].

𝑎𝑖𝐿 = 𝛾𝑖𝑥𝑖 (1)

𝑓𝑖𝐺 = 𝛷𝑖𝑝𝑖 (2)

The Non-Random Two-Liquid (NRTL) method [70] is a widely accepted method to calculate

activity coefficients and other thermophysical data of compounds in liquid multi-component

mixtures without the explicit use of experimental data. This method make use of binary

interaction parameters to describe the interactions between two molecules or two functional

groups in the mixture. Equations of state such as Hayden-O’Connell (HOC) [71] are used to

calculate the fugacity coefficient of a gas phase. Several studies showed that the NRTL method in

combination with the HOC method describes the non-ideality and vapour-liquid equilibria of

mixtures of acetic acid, methanol, methyl acetate and water well [69]. The adapted code also

worked well for mixtures of acetone, furfural, toluene and water.

The NRTL model is a popular method to determine thermodynamic data for liquid mixtures. The

model is based on Scott’s two-liquid model [72] and on an assumption of non-randomness. The

activity coefficient can be determined from equation 3. The NRTL model contains three

parameters, 𝑎𝑖𝑗 , 𝑏𝑖𝑗 and 𝑐𝑖𝑗 , which describe the binary interactions between two components in

the liquid mixture and a parameter 𝛼𝑖𝑗 , which describes the non-randomness [69]. These

parameters are obtained from the commercial simulation software Aspen Plus and are given in

Appendix A.

ln 𝛾𝑖 =∑ 𝜏𝑗𝑖𝐺𝑗𝑖𝑥𝑗𝑗

∑ 𝐺𝑙𝑖𝑥𝑙𝑙+ ∑

𝑥𝑗𝐺𝑖𝑗

∑ 𝐺𝑙𝑗𝑥𝑙𝑙(𝜏𝑖𝑗 −

∑ 𝑥𝑟𝜏𝑟𝑗𝐺𝑟𝑗𝑟

∑ 𝐺𝑙𝑗𝑥𝑙𝑙)

𝑗

(3)

𝐺𝑖𝑗 = exp(−𝛼𝑖𝑗𝜏𝑖𝑗) (4)

𝜏𝑖𝑗 = 𝑎𝑖𝑗 +𝑏𝑖𝑗

𝑇+ 𝑐𝑖𝑗 ln(𝑇) (5)

Equilibrium between a liquid and a vapour mixture is established when the chemical potential of

each component 𝑖 in the liquid phase is equal to the chemical potential of that component in the

vapour phase. If the chemical potential of all components is expressed with respect to the ideal

gas phase, as is shown in equation 6, this reduces to an equality of fugacities (equation 7). The

liquid fugacity of component 𝑖, is its activity multiplied with a standard state fugacity. Using the

definitions of fugacities in equations 2 and 8, the ratio of each component in vapour and liquid

phase can be obtained from equation 9. These ratios combined with the molar balances in

Page 78: Green aldol condensation: synthesis, testing and design of

Materials and methods 57

equation 10 and the trivial conditions in equations 11 and 12, results in a set of equations which

has to be solved to the unknown 𝑥𝑖, 𝑦𝑖 , 𝑛𝐿 and 𝑛𝐺 [69].

𝜇𝑖(𝑇, 𝑝) = 𝜇𝑖,𝑔𝑎𝑠° (𝑇) + 𝑅𝑇 ln (

𝑓𝑖

𝑓°) (6)

𝑓𝑖𝐿 = 𝑓𝑖

𝐺 (7)

𝑓𝑖𝐿 = 𝑎𝑖

𝐿𝑓𝑖° = 𝛾𝑖𝑥𝑖𝛷𝑖

𝑆𝑝𝑖𝑆 exp (

𝑉𝑖𝐿(𝑝 − 𝑝𝑖

𝑆)

𝑅𝑇) (8)

𝐾𝑖 =

𝑦𝑖

𝑥𝑖=

𝛾𝑖𝛷𝑖𝑆𝑝𝑖

𝑆 exp (𝑉𝑖

𝐿(𝑝 − 𝑝𝑖𝑆)

𝑅𝑇 )

𝛷𝑖𝑝

(9)

𝑛𝑖 = 𝑛𝐿𝑥𝑖 + 𝑛𝐺𝑦𝑖 (10)

𝑋 = ∑ 𝑥𝑖

𝑖

− 1 = 0 (11)

𝑌 = ∑ 𝑦𝑖

𝑖

− 1 = 0 (12)

This set of equations has to be solved in an iterative way since all 𝐾𝑖′𝑠 are function of the liquid

and gas compositions. An algorithm to solve this set of equations, is given in Appendix A.

Page 79: Green aldol condensation: synthesis, testing and design of

Materials and methods 58

Page 80: Green aldol condensation: synthesis, testing and design of

Valorization of furfural via the aldol condensation 59

Chapter 3

Valorization of furfural via the

aldol condensation

The aldol condensation might have a bright future in the discipline of green, renewable and

sustainable energy, e.g. in the valorization of biocomponents such as glycerol and furfural.

Hemicellulose is a major component of plant cell walls and can be extracted from a wide range

of renewable raw materials. Furfural is an important intermediate during the conversion of

biomass and is produced from the dehydration of xylose, which is one of the main components

of hemicellulose. Furfural is more available than 4-nitrobenzaldehyde and from an industrial

point of view, it's a more relevant reactant to research. For example, furfural can be used as a

solvent for extractive distillation. It can also be used as reactant in the aldol condensation with

acetone, see Figure 43, to produce C8 and C13 alkanes which can be employed as biofuel [3, 73].

Figure 43: Aqueous-phase conversion of sugars and derivatives into liquid hydrocarbon fuels [73]

Page 81: Green aldol condensation: synthesis, testing and design of

Valorization of furfural via the aldol condensation 60

The reaction scheme is given in Figure 44. Furfural and acetone react together and form the

aldol product 4-(2-furyl)-4-hydroxybutan-2-on which after dehydration leads to a ketone 4-(2-

furyl)-3-buten-2-on. This ketone can undergo further condensation with furfural, forming a C13

ketone 1,4-pentadien-3-on-1,5-di-2-furanyl.

Figure 44: Aldol condensation of furfural with acetone [26]

3.1 Aldol condensation of furfural with acetone

The first experiment was performed at the same reaction conditions as the aldol condensation of

4-nitrobenzaldehyde and acetone (i.e. a temperature of 45 °C). This reaction will be more

elaborately discussed in Chapter 5. In Figure 45, the conversion as a function of time is shown

for both aldol condensations. The reaction rate of the aldol condensation of furfural with acetone

is much lower. The turnover frequency of this reaction is about one order of magnitude lower

than the aldol condensation of 4-nitrobenzaldehyde with acetone. This lower reaction rate can

be explained by the fact that the strongly electron-withdrawing nitro group of 4-

nitrobenzaldehyde makes the carbonyl function more susceptible for a nucleophilic attack to

form the aldol product. Because furfural doesn’t contain such an electron-withdrawing group,

the aldol condensation will proceed relatively slow.

Page 82: Green aldol condensation: synthesis, testing and design of

Valorization of furfural via the aldol condensation 61

Figure 45: Conversion as a function of time for the aldol condensation of furfural with acetone (blue) and 4-nitrobenzaldehyde with acetone (red)

3.2 Optimization of the reaction conditions

In collaboration with Elien Laforce, the optimal reaction conditions for the catalysed aldol

condensation between furfural and acetone was investigated. The molar ratio of the reactants

(furfural/acetone) was varied between 0.004 and 0.124, while a reaction temperature between

60 °C and 100 °C was tested. These experiments were performed in a batch reactor (Parr 4560,

300 mL) in which always 1 mmol/L active sites were added via the catalyst.

In order to obtain higher conversions, higher reaction temperatures (80 °C, 90 °C and 100 °C )

were investigated. When higher temperatures (> 60 °C) are applied, one has to consider the

vapour-liquid equilibria of the reactants. In particular for acetone, which has the lowest boiling

point in the reaction mixture i.e. 56 °C, one has to take into account that a fraction of this

reactant will be in the vapour phase. Due to the higher boiling points of toluene and furfural,

110.6 °C and 161.7 °C respectively, only a very small amount will evaporate. Because the molar

ratio of the reactants has an influence on the reaction rate, it is important to know the exact

composition of the liquid phase after the vapour-liquid equilibrium has been established. Thus,

it is necessary to check how much of each component has to be added to the Parr reactor, in

order to obtain the desired concentrations in the liquid phase.

0

10

20

30

40

50

60

70

80

90

0 20 40 60 80 100 120 140 160 180

Co

nve

rsio

n (

%)

Time (min)

Furfural 4-nitrobenzaldehyde

TOF = 2.70 10-4 s-1

TOF = 2.64 10-3 s-1

Page 83: Green aldol condensation: synthesis, testing and design of

Valorization of furfural via the aldol condensation 62

Therefore, an existing code based on the Non-Random Two-Liquid (NRTL) and Hayden-

O’Connell (HOC) methods has been modified which allows for an initial estimation of the

vapour-liquid equilibrium. This code accounts for the thermodynamic non-ideality of both the

liquid as well as the gas phase via so-called activity coefficients and fugacity coefficients. A

detailed description of the NRTL-HOC code can be found in Materials and methods.

The results from this simulation indicate that this system deviates a few percentages from an

ideal mixture. As was expected, only a fraction of the acetone evaporates at higher temperatures.

For toluene and furfural it is correct to assume that everything remains in the liquid phase.

If a reaction temperature of 90 °C is considered, about 1.55 % of the acetone will evaporate,

which corresponds with a pressure increase up to a total pressure of 3.7 bar. The turnover

frequencies as a function of the molar ratio of the reactants (furfural/acetone) are given in

Figure 46. A reverse S-shape curve is obtained with a maximal reaction rate at a molar ratio

(furfural/acetone) of 0.03. These TOF values illustrate that a decent conversion is obtained for

the aldol condensation of furfural with acetone at 90 °C.

Figure 46: Turnover frequencies for the aldol condensation of furfural with acetone at 90 °C as a function of the molar ratio of the reactants (furfural/acetone)

0.0

0.4

0.8

1.2

1.6

2.0

0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14

TO

F (

x 1

0-3

s-1

)

Ratio furfural/acetone (mol/mol)

Page 84: Green aldol condensation: synthesis, testing and design of

Valorization of furfural via the aldol condensation 63

3.3 Effect of water in the reaction mixture

Also the influence of water will be assessed by experiments. This is an important element

because during the dehydration of one mole xylose, one mole of furfural is formed but also three

moles of water are obtained as by-product. Thus, in the industry the product stream of furfural

will always contain water. Therefore for an industrial application of the aldol condensation with

furfural, it would be beneficial for the feasibility if the reaction can proceed at a significant rate

even in the presence of water. If the heterogeneous catalyst can provide a good activity for the

aldol condensation in an aqueous environment, this would eliminate the need for an expensive

removal of water, to obtain pure furfural. For the aldol condensation of furfural with acetone at

90 °C, the decline of the turnover frequency as a function of the volume percentage of water in

the reaction mixture is given in Figure 47. A small amount of 1 vol% or 2 vol% water is already

sufficient to lower the catalytic activity with 23,43 %, respectively 57,17 %.

Figure 47: TOF ratio as a function of the volume % of water in the reaction mixture In this regard, Periodic Mesoporous Organosilicas can be a good solution. During the synthesis of

these materials, one can chose which organosilicas to add. Thus, if one is able to synthesize a

material with hydrophobic pores, the water could be excluded from entering, leading to a

situation in which only the reactants are able to physisorb in the pores. Due to the limited

amount of synthesized PMO materials, these materials were first tested in the aldol

condensation of 4-nitrobenzaldehyde with acetone, as a better comparison with previous

obtained data would be possible.

0.0

0.2

0.4

0.6

0.8

1.0

0 2 4 6 8 10 12

TO

F r

ati

o

Water (vol %)

Page 85: Green aldol condensation: synthesis, testing and design of

Valorization of furfural via the aldol condensation 64

Page 86: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 65

Chapter 4

Synthesis and characterization

of cooperative acid-base PMOs

materials This chapter describes the synthesis, functionalization and characterization of Periodic

Mesoporous Organosilicas. These PMO materials will be used as a bifunctional catalyst in the

aldol condensation of 4-nitrobenzaldehyde with acetone (Chapter 5).

4.1 Synthesis procedure

In Figure 48, a general overview of the synthesis procedure is shown. The first step in the

synthesis is dissolving a surfactant (Brij-76) in acidified water. The solution is stirred at 50 °C

during at least 4 hours. In a second step, the organosilica precursors are added dropwise to the

clear solution. The synthesis mixture is stirred at 50 °C during 24 hours and subsequently aged

at 90 °C for 24 hours. Next, the precipitate is filtered on a glass filter and washed with water and

acetone. In the final step, the surfactant is removed via extraction with acidified ethanol. The

surfactant cannot be removed by means of calcination at 400 - 500 °C, as this would result in the

destruction of the organic functionalities. Thus, the material is refluxed three times with

acidified ethanol at 80 °C during 24 hours to remove the surfactant as much as possible.

Figure 48: General overview of the synthesis

Page 87: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 66

During the synthesis four different precursors, which are shown in Figure 49, have been used:

vinyltriethoxysilane (VTES), 1,4-bis(triethoxysilyl)ethane (BTEE), 1,4-bis(triethoxysilyl)benzene

(BTEB) and tetraethyl orthosilicate (TEOS).

a) b)

c) d)

Figure 49: Precursors added during synthesis. a) VTES: vinyltriethoxysilane b) BTEE: 1,4-bis(triethoxysilyl)ethane

c) BTEB: 1,4-bis(triethoxysilyl)benzene d) TEOS: tetraethyl orthosilicate In each synthesis the total amount of silicon atoms has been kept at a constant value while

varying the ratio of these precursors to obtain materials with different properties. This allows an

experimental assessment of the effect of the different types of support and its hydrophobicity on

the catalytic activity.

Because the co-condensation of VTES with BTEE or BTEB was not yet described in literature,

a specific procedure that guarantees the formation of ordered mesoporous materials was not

available. Therefore two types of surfactants have been tested in this work. First Pluronic P123

was chosen as surfactant, because it is reported that this structure directing agent lead to pore

sizes bigger than 5 nm. However, the synthesis resulted in gel-like materials, possibly indicating

that the synthesis conditions were not favourable for the interactions between P123 and the

silica precursors. Different synthesis conditions (temperature, pH, additives such as KCl to

improve the hydrothermal stability during the crystallization process) were tested but no

precipitation of a white solid was obtained. Therefore, a more general procedure described by

Burleigh et al. [43] was tested. This recipe uses Brij-76 as surfactant, which is capable of forming

ordered mesoporous materials at broader synthesis conditions. The final recipe that lead to PMO

materials is given in appendix B.

Si

O

OOSiO

O

O Si

O

O

O

Si

O

O

O Si

O

O

O

O

Si

O

OO

Page 88: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 67

4.2 Functionalization

After the synthesis of these PMO materials, the amino acid functionality is introduced into the

pores via a post-functionalization. In a UV reactor, cysteine is grafted onto the vinyl group of

VTES via a thiol-ene "click" reaction.

4.2.1 Thiol-ene “click” chemistry

Thiol-ene click chemistry is a twofold concept, in which the term "click chemistry" was

introduced in 2001 by Kolb et al. [74]. A "click" reaction is a stereospecific reaction that meets a

set of stringent criteria. The reaction is modular and it can thus be applied to various products.

The reaction produces high yields and if by-products are formed, they are harmless and easily

removable. From a practical point of view, these "click" reactions should be easy to carry out

with readily available starting materials and reagents, with no solvent or a solvent that is benign

or easily removed. Ideally the reaction is insensitive to the presence of oxygen and water. Also,

the product should be isolated in a simple manner [75, 76].

The term "thiol-ene" refers to the addition of a thiol group to an olefin. This hydrothiolation of a

carbon-carbon double bond is a very fast reaction and can be carried out at ambient

temperature and pressure. The thiol-ene reaction is easy and versatile to use and usually takes

place under radical conditions, photochemically induced (shown in Figure 50) or via a thermal

treatment [74, 75].

Figure 50: Reaction mechanism for the hydrothiolation of a C=C bond in the presence of a photo-initiator and the appropriate UV radiation. In the propagation circle R stands for cysteine

Page 89: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 68

In general, the reaction contains an initiation, propagation and termination step. Initiation

involves the treatment of a thiol group with a photo-initiator, which after radiation results in the

formation of a thiyl radical, R-S°. During the propagation, first a direct addition of the thiyl

radical to a carbon-carbon double bond takes place, hereby an intermediate carbon radical is

formed. In a second step, a chain transmission is carried out to a second thiol molecule. This

results in the desired thiol-ene product and a new thiyl radical. The reactivity of the thiol-ene

reaction is strongly dependent on the chemical structure and steric effects of the olefin and the

thiol group [75, 76].

4.2.2 Recipe

For the post-functionalization of these periodic mesoporous organosilicas with cysteine,

2-hydroxy-4’-(2-hydroxyethoxy)-2-methylpropiophenone (Irgacure 2959) is used as photo-

initiator. The structure of this complex component is shown in Figure 51.

a) b)

Figure 51: Structure of the components used during the post-functionalization a) Irgacure 2959 b) Cysteine

First 0.50 g photo-initiator is dissolved in 10 mL demineralized water and the long vertical

Schlenk flask is placed into a supersonic bath. Meanwhile 0.53 g cysteine is added to a separate

glass bottle with 10 mL demineralized water. The mixture is heated using a heat gun until a clear

solution is obtained. Next, the cysteine and 0.50 g of the PMO material are brought together in

the Schlenk flask. The valve is used to place the mixture under an inert atmosphere with helium

to avoid that the photo-initiator starts reacting with oxygen. Thereafter, the mixture is placed in

the UV reactor (λ = 254 nm) and stirred during 24 hours for the thiol-ene click reaction to take

place. This can be visualized by a change in the colour of the reaction mixture from white to

yellow-brown (see Figure 52). Finally, the solid is filtered and washed several times with warm

water to remove the remaining unreacted or physisorpted cysteine. More details about the

different washing procedures and the evolution of the physisorpted amine concentrations can

be found in Appendix B.

Page 90: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 69

Figure 52: Functionalization of the synthesized PMO materials The structure of the functionalized PMO materials is shown in Figure 53.

Figure 53: Structure of the functionalized PMO material

Page 91: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 70

4.3 First set of materials

A first set of materials were prepared with VTES and BTEE or VTES and BTEB as precursors

(see Figure 54). In both cases the total amount of silicon atoms in the synthesis mixture was kept

at a constant value. The goal of these experiments was to determine the threshold of the VTES

concentration at which a further increase of VTES would lead to a loss of the mesoscopic

ordering of the material.

The exact concentrations of the precursors added in the synthesis are given in appendix B.

Figure 54: Co-condensation of BTEE with VTES or BTEB with VTES

4.3.1 Molar composition of the PMO materials

The molar composition of the different PMO materials is given in Table 8. The letter in the

abbreviation stands for the type of support: E when ethane was used, B when the benzene group

was added. The number indicates the molar percentages of VTES in the global precursor

mixture. Two of these materials, E10 and B10, have been functionalized with cysteine. These

materials are indicated with the letter C added to the abbreviation.

In the next paragraph, these materials are characterized and a comparison between the

materials is made.

Page 92: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 71

Table 8: Molar composition of the PMO materials

Material VTES (mol%) BTEE (mol%) BTEB (mol%)

E10 10 90 0

E20 20 80 0

E30 30 70 0

B10 10 0 90

B20 20 0 80

4.3.2 Characterization

The synthesized PMO materials were characterized with the use of Diffuse Reflective Infrared

Fourier Transform spectroscopy, nitrogen physisorption, X-ray diffraction and elemental

(CHNS) analysis.

4.3.2.1 DRIFT

In Figure 55, typical DRIFT spectra of the ethane type materials are shown. The small peak at

3728 cm-1 is very common for these organosilica materials as it is caused by the silanol groups.

The broad peak between 3700 cm-1 and 3300 cm-1 is due to the presence of adsorbed water on

the sample and overlaps with the carboxylic acid O-H stretch. The N-H stretch of the primary

amine results in a strong peak around 3400 cm-1 – 3250 cm-1. The peaks between 3000 cm-1 and

2850 cm-1 correspond to the C-H stretch. The tiny peak at 1600 cm-1 can be attributed to the

carbon-carbon double bond of E10. This peak isn't visible after the functionalization because the

vinyl group reacts with the thiol group during the thiol-ene click reaction. For the material

E10-C, the peaks of the carbonyl group (1620 cm-1) and the N-H bending of the primary amine

(1590 cm-1) are more dominant. The peaks that appear in the range 1450 cm-1 to 1250 cm-1 can

be attributed to C-H deformation caused by the surfactant which is not entirely removed from

the material. The strong peak around 1100 cm-1 - 1000 cm-1 corresponds to the siloxane bonds.

Lastly, the peaks between 1000 cm-1 and 600 cm-1 can be attributed to Si-C or Si-O-Si vibrations.

Page 93: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 72

Figure 55: DRIFT spectrum of E10 (blue) and E10-C (red)

4.3.2.2 Nitrogen physisorption and elemental analysis

The mesostructure of the materials is confirmed with the use of nitrogen adsorption and

desorption isotherms. The results of the nitrogen physisorption are shown in Table 9.

Table 9: Properties of the PMO materials

Material SBET (m²/g) Vp (cm³/g) a dp (nm) b

E10 983 0.86 3.19

E10-C 487 0.51 3.12

E20 915 0.61 2.68

E30 765 0.38 2.68

B10 971 0.70 3.10

B10-C 456 0.46 3.04

B20 721 0.50 2.75

a The pore volume is determined with the BJH method. b The average pore size is

determined with the BdB-FHH method [77].

As was expected, an increasing amount of VTES has a detrimental effect on the structure and

porosity of the PMO material. The degree of mesoscopic order of the material decreases with

increasing concentration of VTES in the reaction mixture. For the materials containing ethane as

05001000150020002500300035004000

Ku

be

lka

-Mu

nk

(a

.u.)

wavelength (cm-1)

E10 E10-C

Silanol

Water

Si-OSi-O-Si

N-H

C-H

C=C

C=ON-H

O-H

Page 94: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 73

support, the specific surface area decreases from 983 to 765 m²/g, while the pore volume drops

from 0.86 to 0.38 cm³/g. The same trend can be noticed for the materials with benzene in the

pore walls. These negative effects are also observed in the adsorption isotherms. While E10

still displays a type IV isotherm which corresponds to mesoporous materials, E30 has a type I

isotherm which are encountered when adsorption is limited to one molecular layer.

The amine loading of both functionalized materials is determined via elemental analysis and

results for E10-C and B10-C in 0.2870 mmol/g and 0.0692 mmol/g respectively. The specific

surface area and pore volume further decrease as the cysteine is grafted onto the vinyl groups

inside the pores.

4.3.2.3 XRD

The structure of the mesoporous material can be determined with the use of X-ray diffraction.

Figure 56: XRD diffractogram of E10 (blue) and E10-C (red) The XRD patterns in Figure 56 show an intensive reflection at the low 2θ values which

corresponds to the (100) plane. Usually, these PMO materials also exhibit two smaller peaks at

higher 2θ values, corresponding to the (110) and (200) planes. However, these peaks are not

always present which is the case when the planes are not properly aligned to produce a

diffraction peak. In this case only background is observed.

These peaks indicate that the material has a 2D hexagonal (P6mm) mesostructure. The XRD

diffractograms show that the hexagonal ordered mesoporous structure is maintained after the

functionalization with cysteine. A small shift of the (100) reflection towards higher 2θ values is

0.7 1.7 2.7 3.7 4.7

Inte

nsi

ty (

a.u

.)

2ϴ (degrees)

E10 E10-C

Page 95: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 74

observed which can be attributed to a reduction of the pore diameter of the PMO material due to

the introduction of cysteine in the pores and is confirmed by the results of nitrogen

physisorption in Table 9. As explained in Materials and methods, the XRD data can be used to

calculate the width of the unit cell and the wall thickness, these values are presented in Table 10.

Table 10: Physicochemical properties of the synthesized PMO materials

Material d100 (nm) a Unit cell (nm) dp (nm) Wall thickness (nm)

E10 5.39 6.22 3.19 3.03

E10-C 5.31 6.13 3.12 3.01

E20 5.22 6.03 2.68 3.35

E30 5.08 5.86 2.68 3.18

B10 5.32 6.15 3.10 3.05

B10-C 5.08 5.87 3.04 2.83

B20 5.13 5.92 2.75 3.17

a The d-spacing has been determined with the XRD-software WinXRD.

The d100-spacing decreases with increasing amount of VTES, this can be explained by comparing

the chemical structure of the precursors. While BTEE and BTEB have six ethoxy groups, VTES

has only three. Thus, with increasing amount of VTES in the precursor mixture, this results in a

relative smaller amount of siloxane groups in the pore wall after co-condensation. A second

effect that influences the wall thickness is the pore diameter. With increasing amounts of VTES,

the pore diameter decreases because of a loss of the mesoscopic ordering of the material.

Therefore, materials such as E20 en E30 with smaller pore diameters result in a larger wall

thickness.

It is expected that due to the grafting procedure, only the pore diameter would decrease. This

assumption holds true for the material E10. However, the smaller wall thickness of the material

B10-C, compared to the material before grafting, is something what was not anticipated. This

possibly indicates that the PMO material with benzene groups is more sensitive to the grafting

conditions.

Page 96: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 75

4.4 Reference work

In cooperation with the Centre for Ordered Materials, Organometallics and Catalysis (COMOC),

similar PMO materials have been tested for the aldol condensation of 4-nitrobenzaldehyde with

acetone at 45 °C. The results of this reaction and the turnover frequency of the catalysts are

shown in the following figures.

Figure 57: Conversion as a function of time

Figure 58: Comparison of the turnover frequencies at the same conditions

0

5

10

15

20

25

0 20 40 60 80 100 120 140 160 180 200

Co

nve

rsio

n (

%)

Time (min)

Cysteine (0.08 mmol/g)

Cysteine (0.41 mmol/g)

Cysteine (0.84 mmol/g)

0,0E+00

1,0E-04

2,0E-04

3,0E-04

4,0E-04

5,0E-04

6,0E-04

7,0E-04

8,0E-04

9,0E-04

Silica APTES Silica HMDS APTES PMO cysteine(0.84 mmol/g)

PMO cysteine(0.41 mmol/g)

PMO cysteine(0.08 mmol/g)

TO

F (

s-1)

Page 97: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 76

The obtained turnover frequencies are compared to the TOF of a primary amine which is fully

promoted by the presence of silanol groups (Silica APTES) and the TOF of a primary amine

which is unpromoted after treatment of the silanol groups with 1,1,1,3,3,3-hexamethyldisilazane

(Silica HMDS APTES) [58]. The material with a low loading of cysteine exhibits a TOF which is

lower than a primary amine which is fully promoted by surface silanols but higher than an

unpromoted primary amine. This is consistent with previous research about the acid strength of

the promoting site, which showed that carboxylic acids exert a promoting effect on the activity

of amines, even though less pronounced than the promoting effect of surface silanols [50, 56].

Figure 58 also shows that at high loadings of cysteine (0.41 mmol/g and 0.84 mmol/g) the

activity of the catalyst decreases. This is most likely because at high loadings of cysteine, the

intermolecular interactions result in a protonation of the amines by the carboxylic acid located

on a different linker. As a result, the free electron pair of the amine, necessary for the catalytic

aldol condensation cycle, is no longer available for reaction with the reactants [56, 68].

4.5 Second set of materials

Based upon these results, it was decided to synthesize new PMO materials with an approximate

amine loading in the range of 0.05 - 0.15 mmol/g. A second set of materials were prepared with

VTES, TEOS, BTEE and BTEB as precursors (with varying composition). In all cases the total

amount of silicon was kept at a constant value.

The exact concentrations of the precursors added in the synthesis are given in appendix B.

4.5.1 Determination of the theoretical amine loading

By using some empirical calculations, the molar composition of the precursor mixture was

fine-tuned. Based on their molar masses, but keeping in mind that due to the co-condensation

about 50 % of the ethoxy groups form ethanol, the molar mass of the resulting PMO materials

is calculated. Taking into account the molar percentage of VTES and assuming that the

functionalization will result in an equimolar amount of cysteine, the theoretical amine loading

can be determined. However, some of the vinyl groups of VTES will be part of the pore wall and

are not accessible for the thiol-ene click reaction. The percentage of vinyl groups in the pore wall

is estimated from the first two functionalized materials, E10-C and B10-C, of which the amine

loading was determined via elemental analysis.

As will be clarified in the next paragraphs, these calculations appeared to be quite accurate.

Page 98: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 77

4.5.2 Molar composition of the PMO materials

The molar composition of the different PMO materials is given in Table 11. The letter in the

abbreviation represents the type of support: T stands for TEOS, E when ethane was used and B

when the benzene group was added. The binary mixtures combining TEOS with BTEE or BTEB

are abbreviated as TE and TB respectively. And to avoid further complexity with the

abbreviations, these materials are numbered from 1 to 3.

It should be mentioned that, in a strict manner of speaking, the materials containing TEOS

cannot be labelled as PMO materials as they do not fully comply with the definition of periodic

mesoporous organosilicas in which only bridged organosilica precursors are used. It would be

more correct to use the terminology "modified SBA-15" materials or "organosilicas". In the next

paragraphs, the latter term is used as a more general name to denote the entire group of these

synthesized materials.

Table 11: Molar composition of the organosilicas

Material VTES (mol%) BTEE (mol%) BTEB (mol%) TEOS (mol%)

T1 1.0 0 0 99.0

E1 2.0 98.0 0 0

E2 3.0 97.0 0 0

TE1 3.0 64.7 0 32.3

TE2 3.0 32.3 0 64.7

TB1 7.0 0 62.0 31.0

TB2 5.5 0 47.3 47.3

TB3 4.0 0 32.0 64.0

4.5.3 Characterization

The synthesized organosilicas were characterized with the use of Diffuse Reflective Infrared

Fourier Transform spectroscopy, nitrogen physisorption, X-ray diffraction and elemental

(CHNS) analysis.

Page 99: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 78

4.5.3.1 DRIFT

In Figure 59, typical DRIFT spectra of the benzene type materials are shown. The small peak at

3728 cm-1 is very common for these organosilica materials as it is caused by the silanol groups.

The broad peak between 3700 cm-1 and 3300 cm-1 is due to the presence of adsorbed water on

the sample and overlaps with the carboxylic acid O-H stretch. The C-H stretch of the aromatic

benzene group results in a peak around 3100 cm-1. The strong peak at 2980 cm-1 is due to the

=C-H stretch, originating from the VTES precursor, in the TB1 material. The small peaks between

3000 cm-1 and 2850 cm-1 correspond to the C-H stretch of alkanes. For the material TB1-C, the

peaks of the carbonyl group (1720 cm-1) and the N-H bending of the primary amine (1590 cm-1)

are more visible. Around 1400 cm-1 a peak corresponding to the C-C stretch of aromatics is

observed. The other peaks that appear in the range 1450 cm-1 to 1250 cm-1 can be attributed to

C-H deformation caused by the surfactant which is not entirely removed from the material. The

strong peak around 1100 cm-1 - 1000 cm-1 corresponds to the siloxane bonds. Lastly, the peaks

between 1000 cm-1 and 600 cm-1 can be attributed to Si-C or Si-O-Si vibrations.

Figure 59: DRIFT spectrum of TB1 (blue) and TB1-C (red)

4.5.3.2 Nitrogen physisorption and elemental analysis

The mesostructure of the materials is confirmed with the use of adsorption-desorption

isotherms. The results of the nitrogen physisorption are shown in Table 12.

05001000150020002500300035004000

Ku

be

lka

-Mu

nk

(a

.u.)

wavelength (cm-1)

TB1 TB1-C

Silanol

Water

C-H

=C-H

C-H

Si-OSi-O-Si

C-C

N-H

C=O

Page 100: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 79

Table 12: Properties of the organosilicas before and after functionalization

Material SBET (m²/g) Vp (cm³/g) a dp (nm) b Amine loading

(mmol/g) c Before After Before After Before After

T1 860 688 1.08 0.71 4.08 5.14 0.0530

E1 916 907 0.95 0.94 3.75 3.70 0.0592

E2 928 783 0.94 0.82 3.68 3.68 0.0640

TE1 1030 589 1.07 0.64 3.75 3.49 0.0931

TE2 930 616 1.19 0.70 3.87 3.80 0.1010

TB1 898 664 0.92 0.52 3.40 3.19 0.1357

TB2 875 655 0.93 0.48 3.60 2.96 0.1456

TB3 826 401 0.84 0.23 3.54 3.18 0.0882

a The pore volume is determined with the BJH method. b The average pore size is determined with the

BdB-FHH method [77]. c Amine loading was determined via elemental analysis.

Due to the smaller mol percentage of VTES in the materials, the amount of VTES has no

significant effect on the structure and porosity of the organosilicas. For the binary mixtures,

the specific surface area decreases with increasing amount of TEOS. This is a consequence of the

smaller structure of TEOS compared to BTEE and BTEB. While the pore diameter slightly

increases as the material contains more TEOS, and thus resembling more to a SBA-15 material,

which is known to have pore diameters around 5 nm.

In general, the specific surface area and pore volume decrease as the cysteine is grafted onto the

vinyl groups inside the pores. The functionalization of TB3 resulted in some significant changes

of the properties. TB3-C has a type I isotherm which are encountered when adsorption is limited

to one molecular layer. This manifests itself in a low specific surface area of 401 m²/g and a

small pore volume of only 0.23 cm³/g.

After the functionalization, the materials are washed several times with water to remove the

remaining unreacted or physisorpted cysteine. One has to be careful with materials containing

TEOS, as the increased number of silanol groups makes the material more hydrophilic and thus

more sensitive to a treatment with water. It is known that, due to the presence of organic

functionalities in the pore wall, PMO materials are more stable than pure silica materials. The

PMO materials from paragraph 4.3 were refluxed at 120 °C. However, when the material T1,

consisting of 1 mol% VTES and 99 mol% TEOS, is washed at 120 °C, an detrimental effect on the

pores (see Figure 60) was observed. Confirming that this treatment is too harsh for SBA-15-like

materials.

Page 101: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 80

Figure 60: Comparison between the pore size distribution before and after functionalization and washing procedures. Red: 120 °C, Green: 40 °C

With the treatment at 120 °C, three major peaks are obtained in the pore size distribution,

possibly indicating that several layers of the pore wall have been hydrolysed. Also, as a

consequence of the hydrolysis, a fraction of the material is lost, with a remarkable decrease in

specific surface area and pore volume as result, see Table 13. The X-ray diffractogram of T1 and

T1-C, shown in Figure 62, also confirms that the periodic mesostructure is lost after the

treatment with water at 120 °C.

Table 13: Properties of the T1 material before and after functionalization

Material Number of

washes

Temp

(°C)

SBET

(m²/g)

Vp

(cm³/g) a

dp

(nm) b

Amine loading

(mmol/g) c

T1 860 1.08 4.08

T1-C 2 120 377 0.66 5.10 0.0824

T1-C-S2 5 40 688 0.71 5.14 0.0530

Adapting the washing procedure to a milder temperature of 40 °C and a shorter time (2h - 3h)

results in a material (T1-C-S2) which properties are better preserved, see Table 13. However,

still a small increase in the average pore size is observed and as can be seen in Figure 62, the

hexagonal mesostructure is again lost after repeating the washing procedure five times, leaving

a non-crystalline solid. For the binary mixtures, the temperature is increased to 85 °C to obtain a

better solubility of the cysteine.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 5 10 15 20

Po

re v

olu

me

(cm

³/g

)

Pore diameter (nm)

T1

0

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0 5 10 15 20

Po

re v

olu

me

(cm

³/g

)

Pore diameter (nm)

T1-C T1-C-S2

Page 102: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 81

4.5.3.3 XRD

The structures of these organosilicas can be determined with the use of X-ray diffraction.

Figure 61: XRD diffractogram of E1 (blue) and E1-C (red) Typical XRD patterns of the organosilicas, see Figure 61, show an intensive reflection at the low

2θ values which corresponds to the (100) plane. Also two smaller peaks at higher 2θ values,

corresponding to the (110) and (200) planes are visible. These peaks indicate that the material

has a 2D hexagonal (P6mm) mesostructure. The XRD diffractograms show that the hexagonal

ordered mesoporous structure is maintained after the functionalization with cysteine. A small

shift of the (100) reflection towards higher 2θ values is observed. This change can be attributed

to a reduction of the pore diameter of the PMO material because of the introduction of cysteine

in the pores and is confirmed by the results of nitrogen physisorption in Table 12.

The XRD patterns for the other catalysts look similar to the one of E1 and E1-C. Only the

material TB3-C exhibits one broader peak and the two smaller peaks around a 2θ value of 2.7

are no longer visible. This indicates that the functionalization with cysteine influenced the

mesostructure of this material, as was already observed via nitrogen physisorption by an

significant decrease in specific surface area and pore volume.

The X-ray diffractograms of T1, T1-C and T1-C-S2, in Figure 62, confirm that the periodic

mesostructure is lost after the treatment with water, leaving an amorphous silica material.

0.7 1.7 2.7 3.7 4.7

Inte

nsi

ty (

a.u

.)

2ϴ (degrees)

E1 E1-C

Page 103: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 82

Figure 62: XRD diffractogram of T1 (blue) and T1-C (red) As explained in Materials and methods, the XRD data can be used to calculate the width of the

unit cell and the wall thickness, these values are presented in Table 14.

Table 14: Physicochemical properties of the synthesized organosilicas

Material d100 (nm) a Unit cell (nm) dp (nm) Wall thickness (nm)

Before After Before After Before After Before After

TE1 5.80 5.48 6.70 6.33 3.75 3.49 2.95 2.84

TE2 5.88 5.61 6.79 6.48 3.87 3.80 2.92 2.68

TB1 5.51 4.85 6.36 5.60 3.40 3.19 2.96 2.41

TB2 5.55 4.97 6.41 5.74 3.60 2.96 2.81 2.78

TB3 5.61 5.41 6.48 5.93 3.54 3.18 2.94 2.75

a The d-spacing has been determined with the XRD-software WinXRD.

The d100-spacing decreases with increasing amount of VTES (TB1 > TB2 > TB3), this can be

explained by comparing the chemical structure of the precursors. While BTEE and BTEB have six

ethoxy groups, VTES has only three. Thus, with increasing amount of VTES in the precursor

mixture, this results in a relative smaller amount of siloxane groups in the pore wall after co-

condensation. The d100-spacing also seems to be influenced by the amount of TEOS. This

precursor has an enhancing effect on the ordering of the material. With increasing amount of

0.7 1.7 2.7 3.7 4.7

Inte

nsi

ty (

a.u

.)

2ϴ (degrees)

T1 T1-C T1-C-S2

Page 104: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 83

TEOS, the hexagonal structure of the organosilica is promoted and thus slightly increasing the

distance between two lattice surfaces (d100).

A second effect that influences the wall thickness is the pore diameter. Which was already

determined via nitrogen physisorption. It is expected that due to the grafting procedure, only the

pore diameter would decrease. However, for the TE-materials, a small decrease in wall thickness

of around 5 to 10 % is measured. The biggest change in wall thickness is observed for the

material with the highest amount of BTEB in the precursor mixture, TB1. This decrease also

occurs for the material solely containing benzene bridges (B10-C) and is again an indication

that the benzene groups in the support are more sensitive to the grafting conditions than the

ethane groups.

Page 105: Green aldol condensation: synthesis, testing and design of

Synthesis and characterization of cooperative acid-base PMOs materials 84

Page 106: Green aldol condensation: synthesis, testing and design of

Catalytic experiments 85

Chapter 5

Catalytic experiments

The functionalized materials were used as catalyst in the aldol condensation of acetone with

4-nitrobenzaldehyde. To investigate the influence of the catalyst support's hydrophobicity on

the catalytic activity, the reaction was performed both in the absence and presence of water. To

evaluate the regeneration possibilities, the used catalysts were filtered after the reaction and

washed with various solvents.

5.1 Comparison of ethane and benzene bridged

organosilica supports

The materials with only VTES and BTEE or BTEB (E1-C and B10-C) were first tested as catalyst.

The conversion as a function of the time is given in Figure 63.

Figure 63: Results of the aldol condensation of 4-nitrobenzaldehyde and acetone at 45 °C, Blue: E1-C; Green: B10-C

Although E1-C (0.0592 mmol/g) and B10-C (0.0692 mmol/g) have similar amine loadings, both

catalysts have very different reaction rates. The turnover frequencies are calculated from the

slope of the trendline. This value is divided by the molar percentages of amines in the mixture

and by 60 s/min to obtain the number of molecules that can be converted per active site per unit

0

2

4

6

8

10

12

14

16

0 50 100 150 200 250 300 350

Co

nve

rsio

n (

%)

Time (min)

TOF = 1.77 10-4 s-1

TOF = 1.80 10-5 s-1

Page 107: Green aldol condensation: synthesis, testing and design of

Catalytic experiments 86

of time. The turnover frequency of E1-C is about one order of magnitude higher than B10-C. For

a possible explanation it is necessary to take into consideration that the concentration of

reactants in the pores may be different from the concentration of the reaction mixture. The

properties of the pores play an important role on the physisorption of the reactants into the

pores. It is not unlikely that for the material B10-C the environment of the pores is more in

favour of 4-nitrobenzaldehyde to enter due to some interactions with the benzene groups in the

support. This can be justified with the "like likes like" principle. As a consequence of the relative

higher concentration of 4-nitrobenzaldehyde in the pores, this reactant reacts with the primary

amine and may lead to the formation of a stable Schiff base, which eliminates active sites. This

inhibition of the reaction kinetics was also observed by Kandel et al. [47].

Next, by adding TEOS to the precursor mixture, the influence of an increased number of silanol

groups will be investigated by testing the catalysts containing both hydrophobic and hydrophilic

blocks (TE and TB) in the aldol condensation of 4-nitrobenzaldehyde with acetone at 45 °C.

5.1.1 Catalysts containing both hydrophobic and hydrophilic blocks

The conversion as a function of the time is given in appendix B. An overview of the turnover

frequencies is given in Figure 64. The black dashed lines indicate the expected turnover

frequency in case linear interpolation between the catalysts E1-C, B10-C and T1-C-S2 can be

applied. The experimentally determined TOF values are for both the TE- and TB-materials lower

than expected. For the TB-materials (red curve) this difference is less noticeable because of the

low TOF values. However, for the TE-materials (blue curve) a large decline is observed when

more TEOS is added to the precursor mixture. . Surprisingly, the activity increases again when

the modified SBA-material (T1-C-S2) is considered.

The lower turnover frequencies obtained from the catalytic experiments indicate that no linear

relationship applies for the activity of the catalysts containing both hydrophobic and hydrophilic

blocks. Adding three different precursors to the precursor mixture complicates the synthesis

because every precursor has its own hydrolysis and condensation rates. When the difference

between these rates is too large, the tendency towards homocondensation reactions increases.

This will be a problem during the co-condensation because the homogeneous distribution of the

different organic functionalities in the framework cannot be guaranteed, resulting in a clustered

distribution of the organosilicas, rather than a random one. After functionalization, the amine

functionalities might be too close to each other and are not fully promoted by neighbouring

silanol groups. Leading to lower TOF-values than would be expected from linear interpolation.

Page 108: Green aldol condensation: synthesis, testing and design of

Catalytic experiments 87

Figure 64: Turnover frequency as function of the molar percentage of TEOS in the catalysts. Blue: TE-materials; Red: TB-materials.

Dashed lines indicate the theoretical TOF values in case of linear interpolation A higher density of silanol groups doesn't increase the catalytic activity of the TE-materials. This

suggests that the amount of silanol groups introduced by the ethane precursor are already

sufficient to promote the active sites in the pores. For the TB-materials, it can be observed that

with an increasing amount of the benzene precursor (TB1 > TB2 > TB3), the activity of the

catalyst decreases. This can again be related to the formation of an inactive intermediate,

a Schiff base, and the consequent low activity of the benzene functionality in the aldol

condensation of 4-nitrobenzaldehyde with acetone, as was shown in Figure 63. Increasing the

amount of TEOS has a positive effect on the activity because the environment of the pores

gradually improves into a situation in which a smaller effect of the “like likes like” principle is

expected. Leading to a decreased number of inactive intermediates, and thus the aldol

condensation can proceed at a higher rate.

5.2 Effect of water in the reaction mixture

The influence of the hydrophobicity of the catalyst support on the catalytic activity in the

presence of water is investigated. This is an important property because for a possible

application in industry, e.g. in the valorization of biocomponents such as glycerol and furfural,

the reactants will most likely be in an aqueous solution. Thus the activity of these bifunctional

catalysts in the presence of water will have a major impact on the potential feasibility of these

materials. The amount of water was determined based on the results of the aldol condensation

0.0E+00

2.0E-05

4.0E-05

6.0E-05

8.0E-05

1.0E-04

1.2E-04

1.4E-04

1.6E-04

1.8E-04

2.0E-04

0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0 100.0

TO

F (

1/

s)

TEOS (mol %)

Page 109: Green aldol condensation: synthesis, testing and design of

Catalytic experiments 88

of furfural with acetone, performed by Elien Laforce (see Chapter 3). A small amount of 1 vol%

or 2 vol% water is already sufficient to lower the catalytic activity with 23,43 %, respectively

57,17 %. Due to the low solubility of water in the non-polar reaction mixture (hexane and

acetone), simulations were performed with Aspen Plus. Eventually, only one volume percentage

of water was added to avoid phase separation.

5.2.1 Catalysts containing both hydrophobic and hydrophilic blocks

An overview of the turnover frequencies for the aldol condensation of 4-nitrobenzaldehyde

with acetone in the presence of water is given in Figure 65. The black dashed lines indicate the

expected turnover frequency in case linear interpolation between the catalysts E1-C,

B10-C and T1-C-S2 can be applied. The experimentally determined TOF values are for both the

TE- and TB-materials lower than expected. This can again be related to complexity of the

precursor mixture. When the condensation and hydrolysis rates of the different precursors do

not match, the homocondensation reaction is favoured during the co-condensation. This results

in materials of which the amine functionalities are not completely promoted by silanol groups

due to the clustering of all the precursors.

Figure 65: Turnover frequency as function of the molar percentage of TEOS in the catalysts. Blue: TE-materials; Red: TB-materials.

Dashed lines indicate the theoretical TOF values in case of linear interpolation

0.0E+00

2.0E-05

4.0E-05

6.0E-05

8.0E-05

1.0E-04

1.2E-04

1.4E-04

1.6E-04

1.8E-04

2.0E-04

0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0 100.0

TO

F (

1/

s)

TEOS (mol %)

Page 110: Green aldol condensation: synthesis, testing and design of

Catalytic experiments 89

The turnover frequencies of the TE-materials both for the reference reaction and the reaction in

presence of 1 vol% water are compared in Table 15. For the most hydrophobic material, TE1-C,

the catalytic activity remains at the same level. It is an interesting result to see that, even in the

presence of water, TE1-C is able to preserve its activity. This indicates that the hydrophobic pore

walls impede the water molecules to enter the pores. Thus the reaction continues as it would in

the situation without water. Interestingly, while in the reference reaction the turnover

frequency of TE1-C was still lower than the one of T1-C-S2, this catalyst now becomes more

active in the presence of water. Materials with more TEOS in the pore wall are more susceptible

to the water and therefore the effect of water is more pronounced. For TE2-C, the catalytic

activity decreases with about 15 % but compared to T1-C-S2, the loss in activity is more than

40 %. This illustrates that the hydrophobicity of the ethane group, even if the latter is in small

concentrations, does have a large influence on the catalytic activity.

Table 15: Comparison of the turnover frequencies for the TE-materials

Material TEOS (mol%) TOFreference reaction (s-1) TOFreaction with water (s-1) Ratio

E1-C 0.0 1.77 x 10-4 1.86 x 10-4 1.049

TE1-C 32.3 7.16 x 10-5 7.70 x 10-5 1.075

TE2-C 64.7 4.41 x 10-5 3.73 x 10-5 0.846

T1-C-S2 99.0 8.67 x 10-5 5.15 x 10-5 0.594

The turnover frequencies of the TB-materials in the absence or presence of water are shown in

Table 16. In comparison to the TE-materials, the presence of water has a much larger effect on

the catalytic activity. The TOF values for the aldol condensation in the presence of 1 vol% water

drop significantly. All TB-materials exhibit a turnover frequency around 9 x 10-6 s-1. Nonetheless,

the influence of the support's hydrophobicity is clear. The most hydrophobic material, TB1-C, is

able to retain most of its activity, which is however only 50 %. And while TB1-C had the lowest

activity in the reference reaction, it is now the most active catalyst. Again, materials with more

TEOS in the pore wall are more susceptible to the water and for these materials, TB2-C and TB3-

C, the activity drops with about 75 %. The catalyst TB2-C has a lower TOF value than TB3-C.

However, according to this explanation, the reverse order should be detected. This deviation can

possibly be attributed to experimental errors.

Page 111: Green aldol condensation: synthesis, testing and design of

Catalytic experiments 90

Table 16: Comparison of the turnover frequencies for the TB-materials

Material TEOS (mol%) TOFreference reaction (s-1) TOFreaction with water (s-1) Ratio

B10-C 0.0 1.77 x 10-5 1.30 x 10-5 0.721

TB1-C 31.0 2.00 x 10-5 9.71 x 10-6 0.486

TB2-C 47.3 3.22 x 10-5 9.04 x 10-6 0.281

TB3-C 64.0 4.45 x 10-5 9.54 x 10-6 0.214

T1-C-S2 99.0 8.67 x 10-5 5.15 x 10-5 0.594

A possible explanation why the ratio of the TB-materials is much lower than the TE-materials

has to do with the Snyder polarity index [78]. Water is one the most polar solvents, and has a

Snyder polarity index of 9, while benzene has a lower value of 3. The polarity of ethane is not

easy to find, because the natural state of ethane is the gas phase and thus it is not included in

lists of the solvents polarity. However, knowing that alkanes are the most non-polar solvents

and based on the polarity indices of hexane, octane and decane, one can assume that the polarity

of an ethane compound would be in the range of 0.1 to 0.3. Comparing the polarity index of

ethane and benzene, it can be expected that the ethane groups in the silica support have a

stronger repulsive effect on water. Thus, the TE-materials are capable to exclude water to a

greater extent from the pores than the TB-materials.

5.3 Regeneration

The application of heterogeneous catalysts results in less energy intensive separation processes.

Also, the regeneration and recycling of used catalyst enhance the overall efficiency of the

process. Therefore, the reusability of these materials was investigated. Spent catalyst was

filtered from the reaction mixture on a Büchner funnel with a sintered glass disc. First, assuming

that no inactive intermediates are formed and thus no important changes occur to the

heterogeneous catalyst, the material is simply washed with a small amount of deionized water

and afterwards with a larger amount of acetone to improve the drying of the sample. When the

regenerated materials were tested in the aldol condensation of 4-nitrobenzaldehyde with

acetone at 45 °C, the turnover frequencies dropped significantly. In general, a decrease in

catalytic activity of heterogeneous materials can be attributed to several causes. First, leaching

of the active sites may occur. However, there is no reason to expect that the carbon-sulfur bond

would be cleaved during the aldol condensation. Secondly, the formation of inactive

intermediates on the surface of the catalyst can block the accessibility of the active sites. This

latter case is most likely to be the main reason for the decreased activity. The presence of

Page 112: Green aldol condensation: synthesis, testing and design of

Catalytic experiments 91

intermediates was confirmed by characterizing the catalysts with DRIFT spectroscopy. In Figure

66, the DRIFT spectra of the catalyst before (TE1-C) and after (TE1-C-R1) catalytic reaction are

compared. Zooming in on the region of relevance, the peaks in the range 1550 cm-1 to 1475 cm-1

are slightly elevated which correspond to the N-O asymmetric stretch. Also, an extra peak

between 1400 cm-1 and 1300 cm-1 which can be attributed to the N-O symmetric stretch is

observed. These peaks indicate the presence of a nitro group in the catalytic material and it

suggests that 4-nitrobenzaldehyde still remains inside the pores and formed an intermediate.

Because in the case that the reactant is only physisorpted, 4-nitrobenzaldehyde would have been

washed out by the regeneration procedure. However, there is no increased intensity for the peak

around 1675 cm-1 which corresponds to the C=N stretching of the inactive imine.

Figure 66: Drift spectrum of TE1-C (red) and TE1-C-REGEN (green) A second regeneration method was tested in which the spent catalyst was thoroughly rinsed

with ethyl acetate, deionized water, and ethanol for several times and dried in vacuum at 40 °C.

This method was described by An et al. [61] and worked fine for materials with alternating

hydrophobic and hydrophilic blocks in the pore wall. However, when these regenerated

materials were tested in the aldol condensation, again very low TOF values were obtained

indicating that the inactive intermediates were still present on the catalytic surface. In Figure 67,

the conversion as a function of time for the TE1-C material is plotted. The other catalysts showed

similar trends in terms of activity.

12501300135014001450150015501600165017001750

Ku

be

lka

-Mu

nk

(a

.u.)

wavelength (cm-1)

TE1-C TE1-C-REGEN

imine

N-O

N-O

Page 113: Green aldol condensation: synthesis, testing and design of

Catalytic experiments 92

It should be noted that the catalyst's activity only drops during the first catalytic reaction. When

the catalyst TE1-C was used for a second and third run, the same turnover frequencies are

obtained. This is a possible indication that a single blockage of the pores occurs during the first

catalytic reaction. Hereby enclosing a fraction of the 4-nitrobenzaldehyde, thus explaining why

only the N-O symmetric and asymmetric stretch of a nitro group is observed in the DRIFT

spectra, and no extra peak which corresponds to the C=N stretching of the inactive imine is

present. Afterwards, the active sites that aren't blocked, remain accessible for subsequent

reactions, resulting in a steady catalyst's activity.

Figure 67: Conversion as a function of time for the TE1-C material. Blue: reference reaction; Red: regeneration 1; Green: regeneration 2

Further research is required to clarify which intermediates are responsible for the blockage or

deactivation of the catalysts.

0,0

0,5

1,0

1,5

2,0

2,5

3,0

3,5

4,0

4,5

5,0

0 50 100 150 200 250 300

Co

nve

rsio

n (

%)

Time (min)

TOF = 7.16 10-5 s-1

TOF = 1.90 10-5 s-1

TOF = 1.95 10-5 s-1

Page 114: Green aldol condensation: synthesis, testing and design of

Conclusions and Future Work 93

Chapter 6

Conclusions and Future Work

The aldol condensation might have a bright future in the discipline of green, renewable and

sustainable energy, e.g. in the valorization of biocomponents such as glycerol and furfural.

At the same reaction conditions, the turnover frequency of the aldol condensation of furfural

with acetone is about one order of magnitude lower than the aldol condensation of 4-

nitrobenzaldehyde with acetone. This lower reaction rate can be explained by the fact that

furfural doesn’t contain an electron-withdrawing group. The strongly electron-withdrawing

nitro group of 4-nitrobenzaldehyde makes the carbonyl function more susceptible for a

nucleophilic attack to form the aldol product. Higher reaction temperatures are required to

obtain significant conversions. When higher temperatures (> 60 °C) are applied, one has to

consider the vapour-liquid equilibria of the reactants. The results from a simulation with an

existing code based on the Non-Random Two-Liquid (NRTL) and Hayden-O’Connell (HOC)

methods, indicate that this system deviates a few percentages from an ideal mixture. As was

expected, only a fraction of the acetone evaporates at higher temperatures. For toluene and

furfural it is correct to assume that everything remains in the liquid phase. If a reaction

temperature of 90 °C is considered, about 1.55 % of the acetone will evaporate, which

corresponds with a pressure increase up to a total pressure of 3.7 bar.

During the dehydration of one mole xylose, one mole of furfural is formed but also three moles

of water are obtained as by-product. Thus, in the industry the product stream of furfural will

always contain water. It was found that a small amount of 1 vol% or 2 vol% water is already

sufficient to lower the catalytic activity with 23,43 %, respectively 57,17 %. In this regard,

Periodic Mesoporous Organosilicas can be a good solution. Because it is possible to tune the

hydrophobicity of the catalyst support during the synthesis, one can synthesize materials which

are able to exclude water from entering, leading to a situation in which only the reactants are

able to physisorb in the pores.

In this work several organosilicas have been synthesized. During the synthesis four different

precursors have been used: vinyltriethoxysilane (VTES), 1,4-bis(triethoxysilyl)ethane (BTEE),

1,4-bis(triethoxysilyl)benzene (BTEB) and tetraethyl orthosilicate (TEOS). In all cases the total

amount of silicon atoms in the synthesis mixture was kept at a constant value. By varying the

ratio of these precursors, materials with different properties were obtained. TEOS was added to

the precursor mixture to create a set of materials with changing hydrophobicity. After the

Page 115: Green aldol condensation: synthesis, testing and design of

Conclusions and Future Work 94

synthesis of these organosilicas, the amino acid functionality was introduced into the pores via a

post-functionalization. Cysteine was grafted onto the vinyl group of VTES via a thiol-ene click

reaction which took place in a UV reactor. Next, the solid was filtered and washed several times

with warm water to remove the remaining unreacted or physisorpted cysteine.

In cooperation with the Centre for Ordered Materials, Organometallics and Catalysis (COMOC),

similar PMO materials have been tested for the aldol condensation of 4-nitrobenzaldehyde with

acetone at 45 °C. These catalytic experiments indicated that at high loadings of cysteine

(0.41 mmol/g and 0.84 mmol/g) the activity of the catalyst decreases. This is most likely

because at high loadings of cysteine, the intermolecular interactions result in a protonation of

the amines by the carboxylic acid located on a different linker. As a result, the free electron pair

of the amine, necessary for the catalytic aldol condensation cycle, is no longer available for

reaction with the reactants. Based upon these results, it was decided to synthesize new PMO

materials with an approximate amine loading in the range of 0.05 - 0.15 mmol/g.

All these organosilicas have been characterized with the use of Diffuse Reflective Infrared

Fourier Transform spectroscopy, nitrogen physisorption, X-ray diffraction and elemental

(CHNS) analysis. The DRIFT spectra confirmed the successful functionalization of the vinyl

groups with cysteine. Properties such as the specific surface area, pore volume and average pore

size were determined via nitrogen adsorption and desorption isotherms. Almost all of them

exhibit a type IV isotherm, which corresponds to mesoporous materials. The amine loading of

the functionalized organosilicas was determined via elemental (CHNS) analysis and was in the

range of 0.05 to 0.15 mmol/g. From the X-ray diffractograms, it was found that the washing

procedure at 120 °C is too harsh for the "modified SBA-15" materials (T1-C). Adapting the

washing procedure to a milder temperature of 40 °C and a shorter time (2h - 3h) results in a

material which properties are better preserved. Although, the mesoscopic ordering is still lost

and an amorphous solid was obtained.

The functionalized materials were used as catalyst in the aldol condensation of acetone with

4-nitrobenzaldehyde. The materials with only VTES and BTEE or BTEB (E1-C and B10-C) were

first tested as catalyst. Although E1-C (0.0592 mmol/g) and B10-C (0.0692 mmol/g) have

similar amine loadings, both catalysts have very different reaction rates. The turnover frequency

of E1-C is about one order of magnitude higher than B10-C. For a possible explanation it is

necessary to take into consideration that the concentration of reactants in the pores may be

different from the concentration of the reaction mixture. The properties of the pores play an

important role on the physisorption of the reactants into the pores. It is not unlikely that for the

material B10-C the environment of the pores is more in favour of 4-nitrobenzaldehyde to enter

due to some interactions with the benzene groups in the support. This can be justified with the

Page 116: Green aldol condensation: synthesis, testing and design of

Conclusions and Future Work 95

"like likes like" principle. As a consequence of the relative higher concentration of 4-

nitrobenzaldehyde in the pores, this reactant reacts with the primary amine and may lead to the

formation of a stable Schiff base, which eliminates active sites and results in an inhibition of the

reaction kinetics. The catalysts containing both hydrophobic and hydrophilic blocks (TE- and

TB-materials) were tested in the same reaction. The experimentally determined TOF values are

for both the TE- and TB-materials lower than expected. Adding three different precursors to the

precursor mixture complicates the synthesis because every precursor has its own hydrolysis

and condensation rates. When the difference between these rates is too large, the tendency

towards homocondensation reactions increases. This will be a problem during the co-

condensation because the homogeneous distribution of the different organic functionalities in

the framework cannot be guaranteed, resulting in a clustered distribution of the organosilicas,

rather than a random one. After functionalization, the amine functionalities might be too close to

each other and are not fully promoted by neighbouring silanol groups. Leading to lower TOF-

values than would be expected from linear interpolation.

To investigate the influence of the catalyst support's hydrophobicity on the catalytic activity, the

reaction was also performed in the presence of water. This is an important property because for

a possible application in industry, e.g. in the valorization of biocomponents such as glycerol and

furfural, the reactants will most likely be in an aqueous solution. Thus the activity of these

bifunctional catalysts in the presence of water will have a major impact on the potential

feasibility of these materials. The turnover frequencies of the TE- and TB-materials were

compared for both the reference reaction and the reaction in presence of 1 vol% water.

Especially the materials containing ethane bridges in the catalyst support appear to be quite

promising. It was an interesting result to see that, even in the presence of water, TE1-C is able

to preserve its activity. This indicates that the hydrophobic pore walls impede the water

molecules to enter the pores. Thus the reaction continues as it would in the situation without

water. Materials with more TEOS in the pore wall are more susceptible to the water and

therefore the effect of water is more pronounced. For TE2-C, the catalytic activity decreases with

about 15 % but for the material T1-C-S2, the loss in activity is more than 40 %. In comparison to

the TE-materials, the presence of water has a much larger effect on the catalytic activity of the

TB-materials. The TOF values for the aldol condensation in the presence of 1 vol% water drop

significantly. The most hydrophobic material, TB1-C, is able to retain most of its activity, which

is however only 50 %. A possible explanation why the ratio (TOFreference reaction/TOFreaction with water)

of the TB-materials is much lower than the TE-materials has to do with the Snyder polarity

index. Water is one the most polar solvents, and has a Snyder polarity index of 9. When the

polarity index of ethane (in the range of 0.1 to 0.3) and benzene (around 3) are compared, it can

be expected that the ethane groups in the silica support have a stronger repulsive effect on

Page 117: Green aldol condensation: synthesis, testing and design of

Conclusions and Future Work 96

water. Thus, the TE-materials are capable to exclude water to a greater extent from the pores

than the TB-materials.

Also the reusability of these materials was investigated. Spent catalyst was filtered from the

reaction mixture on a Büchner funnel with a sintered glass disc and regeneration procedures

with various solvents was performed. When the regenerated materials were tested in the aldol

condensation of 4-nitrobenzaldehyde with acetone at 45 °C, the turnover frequencies dropped

significantly. The formation of inactive intermediates on the surface of the catalyst is most likely

the main reason for the decreased activity. These intermediates can block the accessibility of the

active sites. However, it should be mentioned that the catalyst's activity only drops during the

first catalytic reaction. When the catalyst TE1-C was used for a second and third run, the same

turnover frequencies are obtained. This is a possible indication that an one-time only blockage of

the pores occurs during the first catalytic reaction. Afterwards, the active sites that aren't

blocked, remain accessible for subsequent reactions, resulting in a steady catalyst's activity. It

should be useful to study the cause of this deactivation by identifying the exact intermediates.

And possibly elucidate the mechanism by which these intermediates are formed so this

information can be used as a guidance for the design of a next set of cooperative catalysts.

Different regeneration procedures with other solvents have to be tested in order to regain the

original activities.

In this work, differences in surface arrangements due to clustering of the precursors may have

interfered with a correct comparison between the catalysts activity. It might be useful for further

investigations on the effect of the catalyst support’s hydrophobicity to carefully control the

arrangement of the active sites with respect to each other. A material with alternating

hydrophobic and hydrophilic blocks could provide such possibility. It was noticed that repeating

the washing procedure too much, had a negative effect on the mesoscopic ordering of the

functionalized materials. Therefore it is suggested to look for another solvent with a good

solubility for cysteine. Thus a milder temperature and a shorter contact time of the washing

procedure can be applied.

Finally, it is recommended to examine the functionalization of these organosilicas with

secondary amines containing an alcohol group close to the amine. Because literature suggests

that intermolecular amine-silanol or intramolecular amine-alcohol cooperativity is better than

intramolecular amine-carboxylic acid cooperativity. Leading to catalysts with higher turnover

frequencies.

Page 118: Green aldol condensation: synthesis, testing and design of

References 97

Chapter 7

References

[1] J.N. Chheda, G.W. Huber, J.A. Dumesic, Liquid-phase catalytic processing of biomass-

derived oxygenated hydrocarbons to fuels and chemicals, Angew Chem Int Edit, 46 (2007)

7164-7183.

[2] International Energy Agency, Technology Roadmap - Biofuels for Transport. 2011 &

International Energy Agency, Technology Roadmap - Bioenergy for Heat and Power. 2012 &

European Technology Platform, Biomass Technology Roadmap. 2014.

[3] J.N. Chheda, J.A. Dumesic, An overview of dehydration, aldol-condensation and

hydrogenation processes for production of liquid alkanes from biomass-derived

carbohydrates, Catal. Today, 123 (2007) 59-70.

[4] F. Carey, R. Sundberg, Advanced Organic Chemistry, Fifth ed., Springer Science, 2007.

[5] M.B. Smith, J. March, March's Advanced Organic Chemistry, Sixth ed., Wiley-

Interscience, 2007.

[6] L.G. Wade, Organic Chemistry, Eighth ed., Pearson, 2012.

[7] L.M. Baigrie, R.A. Cox, H. Slebockatilk, M. Tencer, T.T. Tidwell, Acid-Catalyzed

Enolization and Aldol Condensation of Acetaldehyde, J Am Chem Soc, 107 (1985) 3640-

3645.

[8] P.T. Buonora, K.G. Rosauer, L.J. Dai, Control of the aqueous aldol addition under claisen-

schmidt conditions, Tetrahedron Letters, 36 (1995) 4009-4012.

[9] J.P. Bouillon, C. Portella, J. Bouquant, S. Humbel, Theoretical study of intramolecular

aldol condensation of 1,6-diketones: Trimethylsilyl substituent effect, Journal of Organic

Chemistry, 65 (2000) 5823-5830.

[10] R.C. Larock, Comprehensive Organic Transformations, 2nd ed., Wiley-VCH1999.

[11] C. Chapuis, D. Jacoby, Catalysis in the preparation of fragrances and flavours, Appl.

Catal. A-Gen., 221 (2001) 93-117.

[12] P.M. Müller, D. Lamparsky, Perfumes: Art, Science, and Technology, Elsevier Science

Publishers Ltd1991.

[13] Ullmann's Encyclopedia of Industrial Chemistry, 2012.

[14] J.N. Armor, A history of industrial catalysis, Catal. Today, 163 (2011) 3-9.

[15] A.D. McNaught, A. Wilkinson, IUPAC Compendium of Chemical Terminology, 2nd

ed., British Royal Society of Chemistry, Cambridge, UK, 1997.

[16] I. Fechete, Y. Wang, J.C. Vedrine, The past, present and future of heterogeneous

catalysis, Catal. Today, 189 (2012) 2-27.

[17] P.T. Anastas, Benign by Design Chemistry, in: P.T. Anastas, C.A. Farris (Eds.) Benign

by Design: Alternative Synthetic Design for Pollution Prevention, Amer Chemical Soc,

Washington, 1994, pp. 2-22.

[18] E.S. Beach, Z. Cui, P.T. Anastas, Green Chemistry: A design framework for

sustainability, Energy Environ. Sci., 2 (2009) 1038-1049.

[19] R.A. Sheldon, The E factor: fifteen years on, Green Chem., 9 (2007) 1273-1283.

[20] I.D. Cosimo, Encyclopedia of Catalysis, Wiley Online Library2010.

[21] E. Dumitriu, V. Hulea, I. Fechete, A. Auroux, J.F. Lacaze, C. Guimon, The aldol

condensation of lower aldehydes over MFI zeolites with different acidic properties, Micropor

Mesopor Mat, 43 (2001) 341-359.

Page 119: Green aldol condensation: synthesis, testing and design of

References 98

[22] O. Kikhtyanin, V. Kelbichova, D. Vitvarova, M. Kubu, D. Kubicka, Aldol condensation

of furfural and acetone on zeolites, Catal. Today, 227 (2014) 154-162.

[23] S.H. Pyo, M. Hedstrom, S. Lundmark, N. Rehnberg, R. Hatti-Kaul, Self- and Cross-

Aldol Condensation of Propanal Catalyzed by Anion-Exchange Resins in Aqueous Media,

Org. Process Res. Dev., 15 (2011) 631-637.

[24] H.H. Liu, W.J. Xu, X.H. Liu, Y. Guo, Y.L. Guo, G.Z. Lu, Y.Q. Wang, Aldol

condensation of furfural and acetone on layered double hydroxides, Kinet. Catal., 51 (2010)

75-80.

[25] L. Hora, V. Kelbichova, O. Kikhtyanin, O. Bortnovskiy, D. Kubicka, Aldol condensation

of furfural and acetone over Mg-Al layered double hydroxides and mixed oxides, Catal.

Today, 223 (2014) 138-147.

[26] O. Kikhtyanin, D. Kubicka, J. Cejka, Toward understanding of the role of Lewis acidity

in aldol condensation of acetone and furfural using MOF and zeolite catalysts, Catal. Today,

243 (2015) 158-162.

[27] F. Hoffmann, M. Cornelius, J. Morell, M. Froba, Silica-based mesoporous organic-

inorganic hybrid materials, Angew Chem Int Edit, 45 (2006) 3216-3251.

[28] B. Hatton, K. Landskron, W. Whitnall, D. Perovic, G.A. Ozin, Past, present, and future

of periodic mesoporous organosilicas - The PMOs, Accounts Chem Res, 38 (2005) 305-312.

[29] P. Van Der Voort, Solid State Chemistry, Cursus, (2009).

[30] 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, A New Family

of Mesoporous Molecular-Sieves Prepared with Liquid-Crystal Templates, J Am Chem Soc,

114 (1992) 10834-10843.

[31] M. Kruk, M. Jaroniec, C.H. Ko, R. Ryoo, Characterization of the porous structure of

SBA-15, Chem Mater, 12 (2000) 1961-1968.

[32] P. Van Der Voort, Nanoporous Materials, Cursus, (2009).

[33] P. Van der Voort, C. Vercaemst, D. Schaubroeck, F. Verpoort, Ordered mesoporous

materials at the beginning of the third millennium: new strategies to create hybrid and non-

siliceous variants, Phys Chem Chem Phys, 10 (2008) 347-360.

[34] F. Hoffmann, M. Froba, Vitalising porous inorganic silica networks with organic

functions-PMOs and related hybrid materials, Chem Soc Rev, 40 (2011) 608-620.

[35] N. Mizoshita, T. Tani, S. Inagaki, Syntheses, properties and applications of periodic

mesoporous organosilicas prepared from bridged organosilane precursors, Chem Soc Rev, 40

(2011) 789-800.

[36] D. Esquivel, Sintesis, caracterizatíon y aplicaciones de materiales periodícos

mesoporosos organosilícicos, Tesis Doctoral, (2006).

[37] M.H. Lim, A. Stein, Comparative studies of grafting and direct syntheses of inorganic-

organic hybrid mesoporous materials, Chem Mater, 11 (1999) 3285-3295.

[38] D.R. Radu, C.Y. Lai, J.G. Huang, X. Shu, V.S.Y. Lin, Fine-tuning the degree of organic

functionalization of mesoporous silica nanosphere materials via an interfacially designed co-

condensation method, Chem Commun, (2005) 1264-1266.

[39] D.A. Loy, K.J. Shea, Bridged Polysilsesquioxanes - Highly Porous Hybrid Organic-

Inorganic Materials, Chem Rev, 95 (1995) 1431-1442.

[40] B.J. Melde, B.T. Holland, C.F. Blanford, A. Stein, Mesoporous sieves with unified

hybrid inorganic/organic frameworks, Chem Mater, 11 (1999) 3302-3308.

[41] S. Inagaki, S. Guan, Y. Fukushima, T. Ohsuna, O. Terasaki, Novel mesoporous materials

with a uniform distribution of organic groups and inorganic oxide in their frameworks, J Am

Chem Soc, 121 (1999) 9611-9614.

[42] T. Asefa, M.J. MacLachlan, N. Coombs, G.A. Ozin, Periodic mesoporous organosilicas

with organic groups inside the channel walls, Nature, 402 (1999) 867-871.

Page 120: Green aldol condensation: synthesis, testing and design of

References 99

[43] M.C. Burleigh, S. Jayasundera, C.W. Thomas, M.S. Spector, M.A. Markowitz, B.P.

Gaber, A versatile synthetic approach to periodic mesoporous organosilicas, Colloid Polym

Sci, 282 (2004) 728-733.

[44] P. Van der Voort, D. Esquivel, E. De Canck, F. Goethals, I. Van Driessche, F.J. Romero-

Salguero, Periodic Mesoporous Organosilicas: from simple to complex bridges; a

comprehensive overview of functions, morphologies and applications, Chem Soc Rev, 42

(2013) 3913-3955.

[45] R.K. Zeidan, S.-J. Hwang, M.E. Davis, Multifunctional heterogeneous catalysts: SBA-

15-containing primary amines and sulfonic acids, Angew Chem Int Edit, 45 (2006) 6332-

6335.

[46] R.K. Zeidan, M.E. Davis, The effect of acid-base pairing on catalysis: An efficient acid-

base functionalized catalyst for aldol condensation, Journal of Catalysis, 247 (2007) 379-382.

[47] K. Kandel, S.M. Althaus, C. Peeraphatdit, T. Kobayashi, B.G. Trewyn, M. Pruski,

Slowing, II, Substrate inhibition in the heterogeneous catalyzed aldol condensation: A

mechanistic study of supported organocatalysts, Journal of Catalysis, 291 (2012) 63-68.

[48] K. Kandel, S.M. Althaus, C. Peeraphatdit, T. Kobayashi, B.G. Trewyn, M. Pruski,

Slowing, II, Solvent-Induced Reversal of Activities between Two Closely Related

Heterogeneous Catalysts in the Aldol Reaction, ACS Catal., 3 (2013) 265-271.

[49] X.F. Yu, X.B. Yu, S.J. Wu, B. Liu, H. Liu, J.Q. Guan, Q. Kan, The effect of the distance

between acidic site and basic site immobilized on mesoporous solid on the activity in

catalyzing aldol condensation, J. Solid State Chem., 184 (2011) 289-295.

[50] N.A. Brunelli, K. Venkatasubbaiah, C.W. Jones, Cooperative Catalysis with Acid-Base

Bifunctional Mesoporous Silica: Impact of Grafting and Co-condensation Synthesis Methods

on Material Structure and Catalytic Properties, Chem Mater, 24 (2012) 2433-2442.

[51] E.L. Margelefsky, R.K. Zeidan, M.E. Davis, Cooperative catalysis by silica-supported

organic functional groups, Chem Soc Rev, 37 (2008) 1118-1126.

[52] K.K. Sharma, A. Anan, R.P. Buckley, W. Ouellette, T. Asefa, Toward efficient

nanoporous catalysts: Controlling site-isolation and concentration of grafted catalytic sites on

nanoporous materials with solvents and colorimetric elucidation of their site-isolation, J Am

Chem Soc, 130 (2008) 218-228.

[53] N.A. Brunelli, S.A. Didas, K. Venkatasubbaiah, C.W. Jones, Tuning Cooperativity by

Controlling the Linker Length of Silica-Supported Amines in Catalysis and CO2 Capture, J

Am Chem Soc, 134 (2012) 13950-13953.

[54] Y. Kubota, K. Goto, S. Miyata, Y. Goto, Y. Fukushima, Y. Sugi, Enhanced effect of

mesoporous silica on base-catalyzed aldol reaction, Chem. Lett., 32 (2003) 234-235.

[55] J. Lauwaert, E. De Canck, D. Esquivel, J.W. Thybaut, P. Van Der Voort, G.B. Marin,

Silanol-Assisted Aldol Condensation on Aminated Silica: Understanding the Arrangement of

Functional Groups, Chemcatchem, 6 (2014) 255-264.

[56] J. Lauwaert, E.G. Moschetta, P. Van Der Voort, J.W. Thybaut, C.W. Jones, G.B. Marin,

Spatial arrangement and acid strength effects on acid–base cooperatively catalyzed aldol

condensation on aminosilica materials, Journal of Catalysis, 325 (2015) 19-25.

[57] Y. Kubota, H. Yamaguchi, T. Yamada, S. Inagaki, Y. Sugi, T. Tatsumi, Further

Investigations on the Promoting Effect of Mesoporous Silica on Base-Catalyzed Aldol

Reaction, Topics in Catalysis, 53 (2010) 492-499.

[58] J. Lauwaert, E. De Canck, D. Esquivel, P. Van Der Voort, J.W. Thybaut, G.B. Marin,

Effects of amine structure and base strength on acid-base cooperative aldol condensation,

Catal. Today, 246 (2015) 35-45.

[59] F. Gelman, J. Blum, D. Avnir, Acids and bases in one pot while avoiding their mutual

destruction, Angew Chem Int Edit, 40 (2001) 3647-+.

Page 121: Green aldol condensation: synthesis, testing and design of

References 100

[60] N.A. Brunelli, C.W. Jones, Tuning acid-base cooperativity to create next generation

silica-supported organocatalysts, Journal of Catalysis, 308 (2013) 60-72.

[61] Z. An, Y. Guo, L.W. Zhao, Z. Li, J. He, L-Proline-Grafted Mesoporous Silica with

Alternating Hydrophobic and Hydrophilic Blocks to Promote Direct Asymmetric Aldol and

Knoevenagel-Michael Cascade Reactions, ACS Catal., 4 (2014) 2566-2576.

[62] J. Schaubroeck, Instrumentele Analyse 1, Cursus, (2011).

[63] A. Verberckmoes, Fysicochemie 2, Cursus, (2012).

[64] Brunauer, Emmett, Teller, Adsorption of Gases in Multimolecular Layers, J Am Chem

Soc, 60 (1938) 309 - 319.

[65] PANanlytical, Bragg's law.

[66] Thermo Scientific, Powder X-ray Diffraction System.

[67] A. Verberckmoes, Spectroscopische Technieken, Cursus, (2011).

[68] J. Ouwehand, J. Lauwaert, Internal discussion, (2015).

[69] J. Lauwaert, J.W. Thybaut, G.B. Marin, Bridging the gap between kinetics in gas phase

and non-ideal liquids exhibiting pronounced polarity effects, (2015).

[70] H. Renon, J.M. Prausnitz, Local compositions in thermodynamic excess functions for

liquid mixtures, Aiche J., 14 (1968) 135-144.

[71] J.G. Hayden, J.P. Oconnell, Generalized method for predicting 2nd virial-coefficients,

Industrial & Engineering Chemistry Process Design and Development, 14 (1975) 209-216.

[72] R.L. Scott, Corresponding states treatment of nonelectrolyte solutions, J. Chem. Phys.,

25 (1956) 193-205.

[73] J.C. Serrano-Ruiz, J.A. Dumesic, Catalytic routes for the conversion of biomass into

liquid hydrocarbon transportation fuels, Energy Environ. Sci., 4 (2011) 83-99.

[74] H.C. Kolb, M.G. Finn, K.B. Sharpless, Click chemistry: Diverse chemical function from

a few good reactions, Angew Chem Int Edit, 40 (2001) 2004-+.

[75] A.B. Lowe, Thiol-ene "click" reactions and recent applications in polymer and materials

synthesis, Polym Chem-Uk, 1 (2010) 17-36.

[76] C.E. Hoyle, C.N. Bowman, Thiol-Ene Click Chemistry, Angew Chem Int Edit, 49 (2010)

1540-1573.

[77] W.W. Lukens, P. Schmidt-Winkel, D.Y. Zhao, J.L. Feng, G.D. Stucky, Evaluating pore

sizes in mesoporous materials: A simplified standard adsorption method and a simplified

Broekhoff-de Boer method, Langmuir, 15 (1999) 5403-5409.

[78] L.R. Snyder, J.J. Kirkland, J.W. Dolan, Introduction to Modern Liquid Chromatography,

Third ed., Wiley, 2009.

Page 122: Green aldol condensation: synthesis, testing and design of

Appendix A 101

Appendix A A1. Parameters

Table 17: Parameters of the NRTL obtained from Aspen Plus

Acetone Furfural Toluene Water

Ace

ton

e

𝑎𝑖𝑗 0.0 -10.7013 0.8404 -3.5004

𝑏𝑖𝑗 0.0 3549.04 -260.308 1347.96

𝑐𝑖𝑗 0.0 0.0 0.0 0.0

𝛼𝑖𝑗 0.0 0.3 0.3 0.3

Fu

rfu

ral

𝑎𝑖𝑗 6.5452 0.0 0.0 -2.7563

𝑏𝑖𝑗 -1015.06 0.0 27.28 1911.42

𝑐𝑖𝑗 0.0 0.0 0.0 0.0

𝛼𝑖𝑗 0.3 0.0 0.3 0.3

To

luen

e

𝑎𝑖𝑗 -0.4612 0.0 0.0 -247.879

𝑏𝑖𝑗 351.203 326.553 0.0 14579.8

𝑐𝑖𝑗 0.0 0.0 0.0 35.582

𝛼𝑖𝑗 0.3 0.3 0.0 0.2

Wat

er

𝑎𝑖𝑗 8.5012 4.2362 627.053 0.0

𝑏𝑖𝑗 -2280.09 -262.241 -27269.4 0.0

𝑐𝑖𝑗 0.0 0.0 -92.7182 0.0

𝛼𝑖𝑗 0.3 0.3 0.2 0.0

A2. Algorithm

An algorithm is given to solve this set of equations in an easy way. In steps 1 to 5 a good initial

value for the vapour fraction and the molar ratio of each component in vapour and liquid phase

is obtained. Afterwards, these initial guesses are used in steps 6 to 11 to iteratively determine

the solution of the set of equations.

Page 123: Green aldol condensation: synthesis, testing and design of

Appendix A 102

Step 1: Set the vapour fraction equal to 0.5

𝜑 =𝑛𝐺

𝑛𝐿 (13)

Step 2: Calculate the molar composition of both phases using equations 14 and 15 and 𝐾𝑖 using

equation 9.

𝑥𝑖 =𝑧𝑖

1 − 𝜑 + 𝐾𝑖𝜑 (14)

𝑦𝑖 =

𝐾𝑖𝑧𝑖

1 − 𝜑 + 𝐾𝑖𝜑

(15)

Step 3: Calculate 𝑋, 𝑌 and 𝜕𝑌

𝜕𝜑 using equations 11, 12 and 16, respectively.

𝜕𝑌

𝜕𝜑=

∑ 𝑧𝑖𝐾𝑖(1 − 𝐾𝑖)𝑖

(1 − 𝜑 + 𝐾𝑖)2 (16)

Step 4: Adapt 𝜑 according to equation 17.

If 𝜕𝑌

𝜕𝜑< 0 𝜑𝑛 =

𝜑𝑛−1

1 − 𝑌

If 𝜕𝑌

𝜕𝜑> 0 𝜑𝑛 = 𝜑𝑛−1(1 − 𝑌)

(17)

Step 5: Go back to step 2 and repeat the iteration until the absolute value of both 𝑋 and 𝑌 is

smaller than 10-4.

Step 6: Calculate the molar composition of both phases using equations 14 and 15.

Step 7: Determine the non-ideality of the mixture based on the NRTL model and the HOC

equation of state and calculate the ratio of each component in vapour and liquid phase using

equation 9.

Step 8: Compare the new 𝐾𝑖𝑛 values with the previous 𝐾𝑖

𝑛−1 values using equation 18.

𝐸 =

√∑ (𝐾𝑖

𝑛 − 𝐾𝑖𝑛−1

𝐾𝑖𝑛 )

2

𝑁𝑖

𝑁

(18)

Step 9: If 𝐸 > 10−6 calculate a new value for 𝐾𝑖𝑛+1 using equation 19 and go back to step 6. If

𝐸 < 10−6 go to step 10.

𝐾𝑖𝑛+1 = √𝐾𝑖

𝑛𝐾𝑖𝑛−1 (19)

Page 124: Green aldol condensation: synthesis, testing and design of

Appendix A 103

Step 10: Calculate 𝑋 and 𝑌 using equations 11, 12 and a new value for the vapour fraction, using

equation 20.

𝜑𝑛+1 = 𝜑𝑛 −

∑𝑧𝑖(1 − 𝐾𝑖

𝑛)

1 + 𝜑𝑛(𝐾𝑖𝑛 − 1)

𝑁𝑖

∑𝑧𝑖(1 − 𝐾𝑖

𝑛)2

(1 + 𝜑𝑛(𝐾𝑖𝑛 − 1))

2𝑁𝑖

(20)

Step 11: If |𝑋 − 𝑌| > 10−5 go back to step 6.

Page 125: Green aldol condensation: synthesis, testing and design of

Appendix B 104

Appendix B B1. Synthesis recipe

3.0 g of Brij-76 (surfactant) is dissolved in 138 mL deionized water and 10 mL HCl (12 M) into a

flask of 250 mL. The solution is stirred at 50 °C during at least 4 hours. The precursors (total

silicon amount: 0.02538 mol) are added dropwise to the clear solution. The synthesis mixture is

stirred at 50 °C during 24 hours and subsequently aged at 90 °C for 24 hours. Next, the

precipitate is filtered on a glass filter and washed with water and acetone. In the final step, the

surfactant is removed via extraction with acidified ethanol. Therefore, 1.0 g of the synthesized

PMO material is dissolved in 50 mL ethanol and 1 mL HCl (12 M) in a flask of 100 mL with reflux.

The mixture is stirred at 80 °C during 24 hours. This surfactant extraction is repeated three

times to remove the surfactant as much as possible.

B2. Precursor mixtures

The exact concentrations of the precursors added in the synthesis of the organosilicas are given

in Table 18.

Table 18: Added amount of precursors

Material VTES (mL) BTEE (mL) BTEB (mL) TEOS (mL)

T1 0.055

5.725

T2 0.109 5.667

TE1 0.164 3.131 1.870

TE2 0.164 1.566 3.740

TB1 0.382 3.285 1.793

TB2 0.300 2.504 2.732

TB3 0.218 1.696 3.701

B3. Washing procedures

After the functionalization, the solid is filtered and washed several times with warm water to

remove the remaining unreacted or physisorpted. It is known that, due to the presence of

organic functionalities in the pore wall, PMO materials are more stable than pure silica

materials. The materials from paragraph 4.3 were refluxed at 120 °C. However, when TEOS is

Page 126: Green aldol condensation: synthesis, testing and design of

Appendix B 105

added to the precursor mixture, one has to be careful because the increased number of silanol

groups makes the material more hydrophilic and thus more sensitive to a treatment with water.

For the “modified SBA-15” materials, the washing procedure was adapted to a milder

temperature of 40 °C. For the materials containing both hydrophobic and hydrophilic blocks, the

temperature is increased to 85 °C to obtain a better solubility of the cysteine. The evolution of

the physisorpted amine concentrations can be found in Table 19.

Table 19: Evolution of the physisorpted amine concentrations

Date washed Temperature (°C) Date analysed Sample Loading (mmol/g)

18/03 40 27/03 T1C-S2-Wash1 1.00499

18/03 40 27/03 T1C-S2-Wash2 0.14736

07/04 40 15/04 T1C-S2-Wash3 0.10745

17/04 40 21/04 T1C-S2-Wash4 0.09010

17/04 40 21/04 T1C-S2-Wash5 0.05297

01/04 120 03/04 TE33-C-Wash1 1.59499

01/04 120 03/04 TE33-C-Wash2 0.19861

15/04 40 21/04 TE33-C-Wash3 0.15042

16/04 85 TE33-C-Wash4

16/04 85 21/04 TE33-C-Wash5 8.46877

17/04 85 21/04 TE33-C-Wash6 0.09310

26/03 120 27/03 TE50-C-Wash1 2.54423

26/03 120 27/03 TE50-C-Wash2 1.10323

02/04 120 15/04 TE50-C-Wash3 0.51067

02/04 120 15/04 TE50-C-Wash4 0.22689

16/04 85 TE50-C-Wash5

16/04 85 21/04 TE50-C-Wash6 0.10580

17/04 85 21/04 TE50-C-Wash7 0.10887

02/04 120 03/04 TE66-C-Wash1 2.98094

02/04 120 15/04 TE66-C-Wash2 2.58385

02/04 120 15/04 TE66-C-Wash3 2.53173

16/04 85 TE66-C-Wash4

16/04 85 21/04 TE66-C-Wash5 0.13922

17/04 85 21/04 TE66-C-Wash6 0.12015

18/04 85 21/04 TE66-C-Wash7 0.10102

03/04 40 15/04 TB33-C-Wash1 3.30192

06/04 40 15/04 TB33-C-Wash2 3.05219

07/04 40 15/04 TB33-C-Wash3 2.86585

16/04 85 TB33-C-Wash4

16/04 85 21/04 TB33-C-Wash5 0.23146

17/04 85 21/04 TB33-C-Wash6 0.19883

18/04 85 21/04 TB33-C-Wash7 0.15714

Page 127: Green aldol condensation: synthesis, testing and design of

Appendix B 106

23/04 85 28/04 TB33-C-Wash8 0.13051

25/04 85 28/04 TB33-C-Wash9 0.14093

03/04 40 15/04 TB50-C-Wash1 3.58992

06/04 40 15/04 TB50-C-Wash2 3.43343

07/04 40 15/04 TB50-C-Wash3 2.99893

17/04 85 TB50-C-Wash4

18/04 85 21/04 TB50-C-Wash5 0.60513

18/04 85 21/04 TB50-C-Wash6 0.17006

23/04 85 28/04 TB50-C-Wash7 0.14657

25/04 85 28/04 TB50-C-Wash8 0.14471

07/04 40 15/04 TB66-C-Wash1 3.14343

07/04 40 15/04 TB66-C-Wash2 2.68894

17/04 85 TB66-C-Wash3

18/04 85 21/04 TB66-C-Wash4 0.58999

18/04 85 21/04 TB66-C-Wash5 0.10181

23/04 85 28/04 TB66-C-Wash6 0.08753

25/04 85 28/04 TB66-C-Wash7 0.08888

B4. Catalytic experiments

The catalytic activity of the materials with both hydrophobic and hydrophilic blocks (TE- and

TB-materials) were tested in the aldol condensation of 4-nitrobenzaldehyde with acetone at

45 °C. The conversion as a function of the time is shown in Figure 68 and Figure 69.

Figure 68: Results of the aldol condensation of 4-nitrobenzaldehyde with acetone at 45 °C, Blue: TE1-C; Green: TE2-C

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

0 50 100 150 200 250 300

Co

nve

rsio

n (

%)

Time (min)

TOF = 7.16 10-5 s-1

TOF = 4.41 10-5 s-1

Page 128: Green aldol condensation: synthesis, testing and design of

Appendix B 107

Figure 69: Results of the aldol condensation of 4-nitrobenzaldehyde with acetone at 45 °C, Blue: TB1-C; Red: TB2-C; Green: TB3-C

B5. Overview of the Labjournal

Subject Page

Introduction 1

Synthesis of 100% BTEE PMO 3

Synthesis of E10 and E40 (recipe described by Maria Chong et al.) 5

Synthesis of E10 and E40 (repeat) 7

Synthesis of E40 (recipe described by Esquivel et al.) 9

Synthesis of E10 (recipe described by Burleigh et al.) 11

Synthesis of E20 and E30 (recipe described by Burleigh et al.) 13

Synthesis of B10 and B20 (recipe described by Burleigh et al.) 15

Functionalization of E10 and B10 with cysteine 17

Kinetic experiment with E10-C and B10-C 18

Synthesis of T1, T2, E1 and E2 19

Synthesis of T1, T2, E1 and E2 (repeat) 21

Characterisation with nitrogen physisorption 22

0.0

0.5

1.0

1.5

2.0

2.5

3.0

0 50 100 150 200 250 300

Co

nve

rsio

n (

%)

Time (min)

TOF = 2.00 10-5 s-1

TOF = 3.22 10-5 s-1

TOF = 4.45 10-5 s-1

Page 129: Green aldol condensation: synthesis, testing and design of

Appendix B 108

Functionalization of T1 with cysteine 23

Functionalization of E1 with cysteine 24

Synthesis of T1, T2, TE1 and TE2 25

Characterisation with nitrogen physisorption 27

Functionalization of E2 with cysteine 28

Functionalization of T1 with cysteine (repeat) 28

Functionalization of TE2 with cysteine 29

Kinetic experiment with E1-C, E2-C and B10-C 30

Synthesis of TE3, TB1, TB2 and TB3 31

Functionalization of TE1 and TE3 with cysteine 33

Functionalization of TB1, TB2 and TB3 with cysteine 34-35

Characterization with nitrogen physisorption 35

Washing procedures 36

Kinetic experiment with TE-materials 37

Kinetic experiment with TE-materials + H2O 39

Kinetic experiment with TE-materials – Regeneration 1 41

Kinetic experiment with TE-materials + H2O – Regeneration 1 43

Kinetic experiment with TE-materials – Regeneration 2 45

Kinetic experiment with TB-materials 47

Kinetic experiment with TB-materials + H2O 49

Kinetic experiment with TB-materials – Regeneration 1 51

Kinetic experiment with TE-materials (repeat) 53

Page 130: Green aldol condensation: synthesis, testing and design of

..