Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

G Model

CATTOD-9236; No. of Pages 7 ARTICLE IN PRESS


Catalysis Today xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Catalysis Today
journal homepage: www.elsevier.com/locate/cattod

Two-step conversion of LLCN olefins to strong anti-knocking alcohol


mixtures catalysed by Rh, Ru/TPPTS complexes in aqueous media
Nikolaos C. Kokkinos a,b,∗ , Nikolaos Nikolaou a , Nikolas Psaroudakis c ,
Konstantinos Mertis c , Sophia Mitkidou a,b , Athanassios Ch. Mitropoulos a,b
a
Eastern Macedonia and Thrace Institute of Technology, Faculty of Engineering, Department of Petroleum & Mechanical Engineering, Ag. Loukas, 654 04
Kavala, Greece
b
Hephaestus Advanced Research Laboratory, Division of Petroleum Forensic Fingerprinting, Faculty of Engineering, Eastern Macedonia and Thrace Institute
of Technology, Ag. Loukas, 654 04 Kavala, Greece
c
University of Athens, Faculty of Chemistry, Department of Inorganic Chemistry, Panepistimiopolis Zografou, 157 71 Athens, Greece

a r t i c l e i n f o a b s t r a c t

Article history: A two-step conversion of light–light cracked naphtha (LLCN) olefins to strong anti-knocking alcohol mix-
Received 16 May 2014 tures is proposed as a potential solution to the serious negative aspects from the use of gasoline ether
Received in revised form 28 June 2014 oxygenates (MTBE, ETBE, TAME) from the refineries. Aqueous biphasic Rh/TPPTS-catalysed hydroformy-
Accepted 30 July 2014
lation reaction of olefins present in a Greek refinery naphtha cut comprises the first part of the two-step
Available online xxx
proposed process. The second part of the proposed LLCN upgrade process is the in situ hydrogenation
of the produced aldehydes to the corresponding alcohols catalysed by Ru/TPPTS complex in aqueous
Keywords:
media. Both catalytic systems of Rh/TPPTS and Ru/TPPTS have been generated in situ by direct addition
Naphtha
Refinery process
of the corresponding catalyst precursors to TPPTS in aqueous media; and they were revealed as effective
Aqueous biphasic hydroformylation catalytic systems for biphasic hydroformylation and biphasic hydrogenation of complicated mixtures,
Aqueous biphasic hydrogenation respectively. The ultimate fuel will contain more oxygen; it will have better combustion properties and
Ruthenium higher octane numbers.
Rhodium/TPPTS © 2014 Elsevier B.V. All rights reserved.

1. Introduction The success of RCH/RP process lies at the use of RhH(CO)(TPPTS)3


water-soluble catalyst, which is eventually recovered with an easy,
In 1973, Manassen [1] introduced the idea of biphasic catalysis efficient and environmentally friendly way. It was shown that
by suggesting the use of two immiscible liquid phases, one hold- the same catalytic system applied in RCH/RP process could also
ing the catalyst and the other the substrate. At the same time, Joó hydroformylate mixtures of olefins. Rh/TPPTS complexes have been
worked on homogeneous aqueous solutions at Debrecen, Hungary, reported as effective catalytic systems for the hydroformylation
contributing significantly on the development of aqueous biphasic of C4-alkene mixtures to C5-aldehydes [10,11]. Furthermore, Bari-
catalysis [2]. Then, the initial idea of biphasic catalysis was appre- celli and coworkers achieved hydroformylation of C6-alkenes in
ciably stimulated by Kuntz’s applied chemistry at Rhône-Poulenc aqueous biphasic media [12] and after few years, they mentioned
[3,4]. Following Kuntz’s experimental work and Joó’s continuing the catalytic activity of RhH(CO)(TPPTS)3 complex in the biphasic
research work [5,6], Cornils’ group at Ruhrchemie AG managed to hydroformylation reaction of C5–C7 olefins from a naphtha cut to
develop a commercial process for the oxo-synthesis of propylene to the corresponding aldehydes achieving up to 86% conversion of
butyraldehyde in a two phase medium [7–9]. In 1984, the so-called the olefins during 200 h of reaction time [13]. We have recently
Ruhrchemie/Rhône-Poulenc (RCH/RP) process was the first suc- examined aqueous biphasic Rh/TPPTS-catalysed hydroformylation
cessful industrial scale-up of aqueous biphasic catalysis applied in reaction of olefins present in the refinery light–light cracked naph-
a hydroformylation reaction, with a production capacity of 100,000 tha (LLCN) from the Aspropyrgos refinery (Hellenic Petroleum S.A.)
tonnes per annum in the plant of Oberhausen, Germany. in Greece [14,15]. The highest conversion of 95.4% of the olefins
present in the real LLCN was observed at 70 ◦ C, 100 bar, at a short
reaction time (6 h). This was the first part of the two-step con-
version process of LLCN olefins to strong anti-knocking alcohol
∗ Corresponding author at: Department of Petroleum & Mechanical Engineering,
mixtures (Scheme 1).
Faculty of Engineering, Eastern Macedonia and Thrace Institute of Technology, Ag.
Loukas, 654 04 Kavala, Greece. Tel.: +30 2510 462231; fax: +30 2510 462231. The second part of the LLCN upgrade process is the in situ
E-mail address: nikokkinos@mwpc.gr (N.C. Kokkinos). hydrogenation of the produced aldehydes of the hydroformylated

http://dx.doi.org/10.1016/j.cattod.2014.07.058
0920-5861/© 2014 Elsevier B.V. All rights reserved.

Please cite this article in press as: N.C. Kokkinos, et al., Two-step conversion of LLCN olefins to strong anti-knocking alcohol mixtures
catalysed by Rh, Ru/TPPTS complexes in aqueous media, Catal. Today (2014), http://dx.doi.org/10.1016/j.cattod.2014.07.058
G Model
CATTOD-9236; No. of Pages 7 ARTICLE IN PRESS
2 N.C. Kokkinos et al. / Catalysis Today xxx (2014) xxx–xxx

Scheme 1. Two-step conversion of LLCN olefins to strong anti-knocking alcohol mixtures catalysed by Rh, Ru/TPPTS complexes in aqueous media.

LLCN to the corresponding alcohols catalysed by Ru/TPPTS detector (FID) and an Agilent 6890N/5973N GC/MS system both
complex in aqueous media (Scheme 1). Ru/TPPTS complexes of them employing a Petrocol DH 150 capillary column (150 m
have been reported as effective catalytic systems for the length, 0.25 mm diameter, 1.0 ␮m film). Column temperature
hydrogenation of aldehydes. Fache et al. [16] examined Ru/TPPTS- was kept initially constant at 40 ◦ C for 5 min. Then it was raised
catalysed hydrogenation reaction of propionaldehyde in aqueous at 200 ◦ C at a rate of 2 ◦ C min−1 and remained there for 45 min.
media. Then, Grosselin et al. [17] first reported the catalytic Helium (99.999%) was used as a carrier gas with a flow rate 20 cm/s
system RuCl3 ·xH2 O/TPPTS for the hydrogenation of 3-methyl-2- at 175 ◦ C and 65 psig. In order to clean the column from various
butenal in aqueous two-phase system (water–toluene). Moreover, remaining redundant compounds the oven temperature was kept
the hydrogenations of 3-phenyl-2-propenal [18–21], 2-butenal at 230 ◦ C for 180 min (the temperature of both the injector and
[18,19] and fructose [22] catalysed by Ru/TPPTS complexes detector was 280 ◦ C) prior to the analysis.
have been reported successfully. Also, Tilloy et al. [23] inves-
tigated the Ru/TPPTS-catalysed hydrogenation of various alde- 2.2. Analytical method
hydes in the presence of cosolvents or cyclodextrins. Our
research group studied for the first time the aqueous biphasic The methods of gas chromatography with FID and mass spec-
hydrogenation of a hydroformylated naphtha model (mixture troscopy were used for qualitative analysis of the LLCN, of the
of 2-methylvaleraldehyde, 2-ethylbutyraldehyde, hexanal, hep- hydroformylated and of the hydrogenated products. In the quan-
taldehyde, 2-phenylpropionaldehyde and hydrocinnamaldehyde) titative analysis of the substrates and of the products, it was
catalysed by Ru/TPPTS complex, which was generated in situ important to correct the areas of the reported peaks of the GC, due to
by direct addition of RuCl3 ·xH2 O catalyst precursor to TPPTS the existence of organic compounds in various homologous series.
(TPPTS = P[m-C6 H4 SO3 Na]3 , trisodium salt of m-trisulfonated Owing to the complexity of the substrate and the lack of literature
triphenyl phosphine) [24,25]. The upgraded LLCN could potentially as far as it concerns the characterisation of complicated organic
replace the harmful gasoline ether oxygenates (MTBE, ETBE, TAME) mixtures, the response factors (fi ) have been calculated by apply-
from the refinery blended gasoline pool [14,15,24,25]. ing Ongkiehong’s equations [26,27] and taking into consideration
In the current research work, a comprehensive overview of the the theoretical approaches of the response in a FID [28–30] for both
two-step conversion of LLCN olefins to strong anti-knocking alcohol the reactants and the products of the reactions (Scheme 2).
mixtures catalysed by Rh, Ru/TPPTS complexes in aqueous media
take place. In addition, the salt effect in aqueous biphasic hydro-
2.3. Two-step conversion of LLCN olefins to alcohol mixtures
genation of a hydroformylated naphtha model is examined along
with the influence of the reaction time in biphasic hydrogenation
The hydroformylation and hydrogenation reactions as well as
of aldehydes present in a real refinery hydroformylated naphtha
the analysis of the products were carried out at the Petroleum &
cut.
Natural Gas Chemistry & Technology and the Chemical Process Sim-
ulations Laboratories of the Department of Petroleum & Mechanical
2. Experimental details
Engineering, of Eastern Macedonia and Thrace Institute of Technol-
ogy. All manipulations and reactions were performed under argon
2.1. Materials and instrumentation
or nitrogen atmosphere using standard Schlenk techniques [31].
In the beginning of the experimental procedure for both
LLCN was granted by Aspropyrgos refinery (Hellenic Petroleum
catalytic reactions (hydroformylation, hydrogenation), deoxy-
S.A.). RhCl3 ·3H2 O and TPPTS were purchased from Alfa Aesar
genation of distilled demineralised water was carried out via an
and used as catalyst precursor and ligand for hydroformylation
ultrasonic cleaning bath under high vacuum for 3 h prior to use.
reactions, respectively. 2-Ethylbutyraldehyde, heptaldehyde, 2-
Zero time of the reaction was taken when the temperature inside
phenylpropionaldehyde and hydrocinnamaldehyde were obtained
the autoclave reached the preferred value. At the end of every
from Aldrich. Hexanal was obtained from Merck and 2-
run the autoclave vessel was rapidly cooled in ice and salt, the
methylvaleraldehyde was purchased from Alfa Aesar. RuCl3 ·xH2 O
stirring was stopped and the reactor was vented before opening
was purchased from Aldrich and used as catalyst precursor
for hydrogenation reactions. Toluene extra pure and NaCl were
obtained from Merck. Syn-gas (CO/H2 = 1/1), H2 , N2 , Ar, Zero Air,
and He 5.0 N were purchased from Aeroscopio Hellas and Axarlis.
CO/H2 (1/1) and hydrogen 5.0 N were used in the hydroformy-
lation and hydrogenation reactions, respectively, without further
purification. Na2 SO4 was used for drying and was obtained from
Panreac.
All the reactions were performed in 100 mL Autoclave Engineers
batch reactor which was equipped with a low carbon stainless
steel vessel (316 LSS), an electromagnetic agitation system, a
cooling–heating system, a central control unit (URC) and a sam-
Scheme 2. The response factors (fi ) equations for olefins, aldehydes and alcohols,
pling valve. The substrate and the products were analysed using a respectively based upon Ongkiehong’s results. Mi : molecular weight of a compound
Perkin Elmer 8700 gas chromatograph (GC) with a flame ionisation i, ni : the amount of carbon atoms of a compound i.

Please cite this article in press as: N.C. Kokkinos, et al., Two-step conversion of LLCN olefins to strong anti-knocking alcohol mixtures
catalysed by Rh, Ru/TPPTS complexes in aqueous media, Catal. Today (2014), http://dx.doi.org/10.1016/j.cattod.2014.07.058
G Model
CATTOD-9236; No. of Pages 7 ARTICLE IN PRESS
N.C. Kokkinos et al. / Catalysis Today xxx (2014) xxx–xxx 3

it. A separation funnel was used to divide the withdrawn reaction contained un-substituted linear olefins (25.46%), mono-substituted
mixture into two phases. Immediately then, the organic layer was branched olefins (33.81%) and cyclic olefins (2.26%). There were
analysed by gas chromatography. few more peaks in the GC report (indentified as C1 –C3 alkanes and
In the hydroformylations experimental procedure, the Rh/TPPTS C5 dienes) that were rejected, due to their total negligible amount
catalyst used was prepared in situ by direct addition of RhCl3 to being less than 0.05 mol%.
TPPTS in aqueous solutions. The redox reaction between RhCl3 and
TPPTS in water–toluene biphasic system [32] and the syn-gas gen- 3.2. First step: aqueous biphasic catalytic hydroformylation of
erate in situ the water-soluble catalytic complex RhH(CO)(TPPTS)3 , olefins from a refinery LLCN
which is one of the most efficient catalysts of homogeneous hydro-
formylation of olefins [33]. Thus, in a typical experiment, an A series of experiments was carried out with temperatures in the
aqueous solution of the RhCl3 ·3H2 O catalyst precursor (22.6 mg; range of 30–130 ◦ C, pressures between 25 bar and 100 bar, reaction
0.09 mmol) and TPPTS ligand (248 mg; 0.34 mmol) in water (15 mL) times between 3 h and 36 h, TPPTS/Rh molar ratio in the range of
together with LLCN (7.5 mL; 42.90 mmol olefins) in toluene (7.5 mL) 2–6 equiv. TPPTS per atom of rhodium and olefins/Rh molar ratio
were fed in the stainless steel autoclave reactor. The reactor was between 125 and 1500 per atom of rhodium. At the end of all
then purged with syn-gas at room temperature. Later, syn-gas the hydroformylation reactions, the reaction mixture was strictly
(CO/H2 = 1/1) pressurisation and heating were taken place accord- biphasic. The upper organic layer was colourless and it contained
ing to the desired reaction pressure and temperature. the produced aldehydes, while the catalyst remained in the yellow-
In the hydrogenations experimental procedure, a pot with the ish lower aqueous layer [34].
substrate (2-methylvaleraldehyde, 2-ethylbutyraldehyde, hexanal,
heptaldehyde, 2-phenylpropionaldehyde and hydrocinnamalde- 3.2.1. Influence of reaction pressure, temperature and time
hyde in a 1:1:1:1:1:1 molar ratio) was stirred gently in ambient Table 2 presents the activity of the catalyst in the aqueous
temperature in order to achieve better homogenisation of the alde- biphasic catalytic hydroformylation of olefins from a refinery LLCN
hydes mixture. The Ru-TPPTS catalyst used was prepared in situ as a function of syn-gas pressure, temperature and time. The rise of
by direct addition of RuCl3 to TPPTS in aqueous solutions. Thus, in syn-gas pressure in the solution increases CO/H2 dissolution in the
a typical experiment, an aqueous solution of the RuCl3 ·xH2 O cat- liquid phase and thus, CO/H2 concentration at the water–organic
alyst precursor (11.7 mg; 0.04 mmol) and TPPTS ligand (110 mg; interface favouring so the formation and further stabilisation of
0.18 mmol) in water (15 mL) together with a 1:1:1:1:1:1 mixture RhH(CO)(TPPTS)3 catalytic complex [33,35,36]. Subsequently, a
of the six aldehydes (6 mL; 44.21 mmol) in toluene (9 mL) were fed higher conversion was observed by increasing CO/H2 (1:1) pres-
in the stainless steel autoclave reactor. The reactor was then purged sure. The highest conversion of 95.4% of the olefins was observed
with hydrogen at room temperature. Afterwards, hydrogen pres- at 100 bar and 70 ◦ C during 6 h of reaction.
surisation and heating were carried out according to the desired Furthermore, Table 2 shows that the yield of the hydroformy-
reaction pressure and temperature. lation reaction was increasing by raising the reaction temperature
up to 70 ◦ C. For temperatures higher than 70 ◦ C, it was observed
3. Results and discussion a decrease in the activity of the catalytic system in the presence
of CO/H2 . Under the last conditions with TPPTS/Rh molar ratio of
3.1. Light–light cracked naphtha characterisation 3.8 equiv. of TPPTS per Rh atom the catalytic system appeared to
be unstable above 70 ◦ C and for temperature higher than 130 ◦ C
A light–light cracked naphtha characterisation was performed it seems to be totally deactivated [37]. The low conversion of
by using a Perkin Elmer 8700 GC/FID and an Agilent 6890N/5973N 14.3% under 30 ◦ C revealed the mass transfer difficulties normally
GC/MS (MS library NIST 05). The LLCN (Table 1), which was obtained encountered in most biphasic systems without the use of any sur-
from Aspropyrgos refinery in Greece, was constituted by parafins factants.
(6%), isoparafins (31%), olefins (62%), and naphthens (1%). The LLCN Regarding the parameter of reaction time, it is worthy of remark
that 40% of the olefins were hydroformylated within the first 3 h of
Table 1 the experiments. For reaction times more than 3 h, both conversion
Characterisation of LLCN (PIONA Analysis). and yield continued to increase with small and steady steps. After
3 h the reaction rate became much slower. This can be more likely
LLCN
explained by the fact that the remaining branched olefins are less
Components Molar allocation (%) reactive than the linear ones, which almost all of them have been
Parafins 6.13 already consumed within the initial 3 h of reaction time.
n-Butane 0.23
n-Pentane 5.90
3.2.2. Effect of TPPTS/Rh and olefins/Rh molar ratios
Isoparafins 31.46
Isopentane 31.25
The water-soluble ligand used in our catalytic system was
2,2-Dimethyl-butane 0.11 trisulfonated triphenyl phosphine (TPPTS). All the experiments
2,3-Dimethyl-butane 0.10 regarding the influence of TPPTS/Rh molar ratio were carried out
Olefins 61.53 at a temperature of 70 ◦ C and at a pressure of 75 bar. According to
1-Butene 0.11
Table 3, the best results were obtained at a TPPTS/Rh molar ratio
trans-2-butene 0.90
cis-2-butene 0.65 of 2–3 equiv. of TPPTS per Rh atom. However, it must be noted
3-Methyl-1-butene 0.46 that for TPPTS/Rh molar ratios higher than 4, the conversion of the
1-Pentene 1.52 olefins of the LLCN sharply decreased. This could spring from the
2-Methyl-1-butene 5.90 fundamental issue of the generation of a vacant coordination site
trans-2-pentene 16.72
cis-2-pentene 5.56
in the coordination sphere of the catalytic metal centre, allowing
2-Methyl-2-butene 27.45 the coordination of the olefin and so the initiation of the catalytic
Cyclopentene 2.26 hydroformylation of olefins in the water. On the other hand, an
Naphthens 0.88 excess of TPPTS ligand in aqueous media shifts the equilibrium
Cyclopentane 0.88
towards coordinatively saturated complexes, RhH(CO)(TPPTS)3 or
Aromatics 0
RhH(CO)2 (TPPTS)2 , that are less reactive.

Please cite this article in press as: N.C. Kokkinos, et al., Two-step conversion of LLCN olefins to strong anti-knocking alcohol mixtures
catalysed by Rh, Ru/TPPTS complexes in aqueous media, Catal. Today (2014), http://dx.doi.org/10.1016/j.cattod.2014.07.058
G Model
CATTOD-9236; No. of Pages 7 ARTICLE IN PRESS
4 N.C. Kokkinos et al. / Catalysis Today xxx (2014) xxx–xxx

Table 2
Effect of syn-gas pressure, temperature and reaction time in the hydroformylation of olefins from a refinery LLCN catalysed by Rh/TPPTS complex in water–toluene system.

P (bar) T (◦ C) t (h) ␣-Olefins Other olefins Total olefins Total aldehydes Aldehydes Conversion Yield (mol%) TOF (h−1 ) TON
(mol%) (mol%) (mol%) (mol%) (n/iso) (mol%)

25 70 6 2.20 64.70 66.90 31.30 0.46 33.1 31 61 365


50 70 6 0.68 47.59 48.27 50.46 0.39 51.7 50 98 588
75 70 6 0.92 30.97 31.89 66.17 0.25 68.1 66 129 771
100 70 6 0.00 4.62 4.62 94.91 0.11 95.4 95 184 1106
50 30 6 0.73 85.01 85.74 14.26 0.18 14.3 14 28 166
50 50 6 0.00 55.69 55.69 43.99 0.17 44.3 44 85 513
50 90 6 0.68 53.32 54.00 45.19 0.38 46.0 45 88 527
50 110 6 0.87 58.54 59.41 39.94 0.46 40.6 40 78 465
50 130 6 2.51 77.54 80.05 19.10 0.42 19.9 19 37 223
50 90 3 0.87 59.18 60.04 39.02 0.41 40.0 39 152 455
50 90 12 0.75 49.58 50.33 48.80 0.41 49.7 49 47 569
50 90 24 0.44 45.37 45.81 53.37 0.40 54.2 53 26 622
50 90 36 0.00 28.65 28.65 60.10 0.36 71.3 70 23 820

Conditions: agitation = 600 rpm.


Composition: 22.6 mg (0.09 mmol) RhCl3 ·3H2 O, 248 mg (0.34 mmol) TPPTS, 15 mL deoxygenated distilled demineralised water, 7.5 mL LLCN (42.90 mmol olefins), 7.5 mL
toluene, olefins/Rh = 500.

In addition, the effect of the molar ratio of olefins/Rh in the in the range of 3–6 equiv. TPPTS per atom of ruthenium and
aqueous biphasic hydroformylation of olefins present in naphtha substrate/Ru molar ratio between 500 and 2000 per atom of
was thoroughly examined. For this reason, seven olefins/Rh molar ruthenium. At the end of the hydrogenation reaction, the upper
ratios, between 125 and 3000, were examined in order to find the organic layer was colourless and it contained the produced alco-
highest possible catalytic activity at the specific reaction conditions, hols (2-ethyl-1-butanol, 2-methyl-1-pentanol, hexanol, heptanol,
as shown in Table 3. It was observed that as the ratio of olefins/Rh 2-phenyl-1-propanol and 3-phenyl-1-propanol), while the catalyst
was increased up to 1500 so higher was the TOF. The increase in remained in the yellowish lower aqueous layer.
TOF was indeed profound from 250 (TOF 35 h−1 ) up to 1500 (TOF
436 h−1 ) molar ratio of olefins/Rh. For molar ratio of olefins/Rh 3.2.3.1. Influence of reaction pressure, temperature and time. Table 4
higher than 1500 it was noticed a decrease in the activity of the cat- presents the activity of the catalyst in the hydrogenation of the mix-
alytic system in the presence of CO/H2 . Probably, the active metal ture of aldehydes as a function of hydrogen pressure, temperature
centres of the catalyst reached the zenith of their catalytic activity, and time. According to Table 4, a higher conversion was observed
as it is shown in the TOF column of Table 3. by increasing H2 pressure. The positive correlation of pressure and
conversion could be possible explained by the increase of H2 dis-
3.2.3. Second step: aqueous biphasic catalytic hydrogenation of a solution in the liquid phase, during H2 pressure rise. Moreover,
hydroformylated naphtha model the increase of hydrogen concentration in the solution favours the
In the whole range of experiments, the model of hydro- formation and further stabilisation of RuHCl(TPPTS)3 catalytic com-
formylated naphtha consists of 2-methylvaleraldehyde, plex [18,19,38].
2-ethylbutyraldehyde, hexanal, heptaldehyde, 2- Furthermore, a higher yield was observed by increasing the reac-
phenylpropionaldehyde and hydrocinnamaldehyde in a tion temperature up to 100 ◦ C. For temperatures higher than 100 ◦ C
1:1:1:1:1:1 molar ratio. The above mixture of aldehydes (6 mL) it was observed a decrease in the activity of the catalytic system.
composes the substrate of the hydrogenation reactions. The This observation according to Fache et al. [16] and Heinen et al.
1:1:1:1:1:1 molar ratio was chosen for the sake of simplicity; due [22] is due to deactivation and instability of the catalyst system
to the fact that the molar ratio of naphtha components varies from since under these conditions it is favoured the creation of arenes’
region to region. Toluene (9 mL) was used as the organic solvent complexes of the type [Ru(H)(n6 -C6 H5 CH3 )(TPPTS)2 ]Cl, that they
and together with the substrate composes the organic phase of the reduce (oppose to) the activity of the catalytic system [18,19].
biphasic catalytic system. Therefore, it is revealed that the catalyst is not stable in the presence
A series of experiments was carried out with pressures between of hydrogen at temperatures above 100 ◦ C.
10 bar and 70 bar, temperatures in the range of 22–110 ◦ C, reac- The zero time of the reaction (start-up time) was regarded as the
tion times between 10 min and 300 min, TPPTS/Ru molar ratio time moment where the system reaches the preset temperature

Table 3
Effect of TPPTS/Rh and olefins/Rh molar ratios in the hydroformylation of olefins from a refinery LLCN catalysed by Rh/TPPTS complex in water–toluene system.

TPPTS/Rh C C/Rh ␣-Olefins Other olefins Total olefins Total aldehydes Aldehydes Conversion Yield (mol%) TOF (h−1 ) TON
(mol/mol) (mol/mol) (mol%) (mol%) (mol%) (mol%) (n/iso) (mol%)

2 500 0.00 29.33 29.33 69.79 0.27 70.7 70 136 813


3 500 0.00 30.63 30.63 68.42 0.32 69.4 68 133 797
5 500 0.31 42.59 42.90 55.96 0.38 57.1 56 109 652
6 500 0.97 87.56 88.53 10.77 0.28 11.5 11 21 126
4 125 2.20 64.70 66.90 31.30 0.46 33.1 31 15 92
4 250 0.60 62.33 62.93 35.65 0.30 37.1 36 35 208
4 500 0.00 40.52 40.52 58.99 0.32 59.5 59 115 688
4 1000 0.00 16.32 16.32 81.48 0.22 83.7 81 317 1899
4 1500 0.00 24.62 24.62 74.48 0.25 75.4 74 436 2613
4 2000 1.80 94.59 96.39 2.92 0.46 3.6 3 23 140
4 3000 2.00 95.01 97.01 2.30 0.43 3.0 2 18 110

Conditions: T = 70 ◦ C, PCO/H2 = 75 bar, t = 6 h, agitation = 600 rpm.


Composition: 15 mL deoxygenated distilled demineralised water, 7.5 mL LLCN (42.90 mmol olefins), 7.5 mL toluene.

Please cite this article in press as: N.C. Kokkinos, et al., Two-step conversion of LLCN olefins to strong anti-knocking alcohol mixtures
catalysed by Rh, Ru/TPPTS complexes in aqueous media, Catal. Today (2014), http://dx.doi.org/10.1016/j.cattod.2014.07.058
G Model
CATTOD-9236; No. of Pages 7 ARTICLE IN PRESS
N.C. Kokkinos et al. / Catalysis Today xxx (2014) xxx–xxx 5

Table 4 Table 5
Effect of hydrogen pressure, temperature and reaction time in the hydrogena- Effect of TPPTS/Ru and aldehydes/Ru molar ratios in the hydrogenation of the
tion of the hydroformylated naphtha model catalysed by Ru/TPPTS complex in hydroformylated naphtha model catalysed by Ru/TPPTS complex in water–toluene
water–toluene system. system.

P (bar) T (◦ C) t (min) Conversion Yield TOF (h−1 ) TON TPPTS/Ru CHO/Ru Conversion Yield (mol%) TOF (h−1 ) TON
(mol%) (mol%) (mol/mol) (mol/mol) (mol%)

10 80 60 51.1 50 498 498 3 1000 59.6 57 562 562


20 80 60 70.5 70 696 696 4 1000 70.5 70 696 696
30 80 60 71.4 71 699 699 5 1000 65.1 62 618 618
50 80 60 75.7 75 745 745 6 1000 2.7 2 17 17
70 80 60 80.6 78 774 774 4 500 77.7 75 742 742
20 22 120 1.7 1 4 9 4 1500 22.6 19 186 186
20 40 120 2.0 1 7 14 4 2000 15.2 12 121 121
20 60 120 3.7 3 16 33
Conditions: T = 80 ◦ C, PH2 = 20 bar, t = 1 h, agitation = 650 rpm.
20 80 120 87.8 86 424 848
20 100 120 97.4 98 486 973
20 110 120 97.1 94 467 934
20 120 120 97.8 92 453 905 conversion of the mixture of aldehydes sharply decreased; possibly,
20 140 120 94.5 92 454 909 large amount of TPPTS shifts the equilibrium towards coordina-
50 80 10 18.9 18 1090 182
tively saturated species which are less reactive. In addition, four
50 80 45 60.8 60 790 592
50 80 90 87.4 87 572 858 substrate/Ru molar ratios were tested, between 500 and 2000, in
50 80 120 94.3 93 461 921 order to find the highest possible conversion at the optimal reac-
50 80 300 96.1 93 184 919 tion conditions, as shown in Table 5. It was observed that as the
Conditions: agitation = 650 rpm. ratio of aldehydes/Ru was decreased so higher was the conversion.
Composition: TPPTS/Ru = 4, aldehydes/Ru = 1000, 6 mL aldehydes, 9 mL toluene, The optimal substrate/Ru molar ratio for our series of experiments
15 mL deoxygenated distilled demineralised water. can be regarded to be the lower one, the ratio being 500.

(set point) and as the end of the reaction, the moment that the 3.2.3.3. Salt effect. Table 6 compares the activity of the Ru/TPPTS
heating was stopped. Table 4 shows that the 18.9% of the substrate catalyst in the hydrogenation of the mixture of aldehydes with
was hydrogenated within the first 10 min of the experiments and a and without addition of salt. In the presence of salt, it is worthy of
very steep increase of both conversion and yield rates was observed remark that 66% of the aldehydes were hydrogenated to alcohols
within the first 1 h of the reaction. For reaction times more than within the first 15 min of the reaction (TOF = 2607 h−1 ); in contrast
2 h, both conversion and yield seem to be stabilised. The initially with the 20% yield of the corresponding experiment without salt
high increase and further stabilisation of conversion in connection (TOF = 803 h−1 ). According to Fache et al. [16,39], the presense of
with reaction time is a typical characteristic of most hydrogenation salt is a significant parameter in the aqueous biphasic catalysis.
reactions [16,17,21,22,39]. When the salt was added in the aqueous phase, the convertion of
Consequently, it appears that the optimal values of hydrogen the mixture of aldehydes increased; since the cation favoured the
pressure, temperature and reaction time in respect not only to coordination of the aldehyde to the catalyst. For a given cation (e.g.
high conversion and yield of the reaction, but also to the mildest Na+ ) in the Ru/TPPTS-catalysed hydrogenation of propionaldehyde,
conditions for hydrogenation, were 20 bar, 80 ◦ C and 2 h, corre- the reaction rate increased according to the nature of the anion,
spondingly. with I− achieving the best results [16,39]. In fact, the key parame-
ter of the enhanced activity in water depends on the nature of the
salt rather than the precursor used.
3.2.3.2. Effect of TPPTS/Ru and aldehydes/Ru molar ratios. Previous
research works on the hydrogenation of individual aldehydes using
the RuCl3 /TPPTS catalyst have revealed that the conversion is a 3.2.4. Aqueous biphasic catalytic hydrogenation of a
strong function of the ligand to metal ratio [17–19,38]. TPPTS hydroformylated real LLCN
becomes highly soluble to the water phase due to the sulfonate The influence of the reaction time in biphasic hydrogenation of
groups. Subsequently, it is sensible to examine the effect of molec- aldehydes present in a real refinery hydroformylated naphtha cut
ular ratio of L/Me to the catalytic activity of the active complex. All was also investigated. The substrate of this series of experiments
the experiments regarding the influence of TPPTS/Ru molar ratio was the products of the aqueous biphasic catalytic hydroformyla-
were carried out at the temperature and pressure of 80 ◦ C and tion of olefins from a Greek refinery LLCN (step 1). The start-up time
20 bar, respectively. Based on our experiments that are reported of the reaction was regarded as the time moment where the system
in Table 5, it was clear that the best TPPTS/Ru molar ratio was reaches the temperature set point and as the end of the reaction,
4–5 equiv. of TPPTS per Ru atom. According to Grosselin et al. [17] the moment that the heating was stopped. Table 7 reveals a pos-
and Nuithitikul and Winterbottom [38], the lack of TPPTS surplus itive correlation between the reaction time and the conversion of
leads to a reduction in the activity of the catalytic system. However, the including aldehydes to the corresponding alcohols. Particularly,
it must be noted that for TPPTS/Ru molar ratios higher than 5, the there was a significant increase in the conversion of the reaction

Table 6
Salt effect in the hydrogenation of the hydroformylated naphtha model catalysed by Ru/TPPTS complex in water–toluene system.

Salt Substrate P (bar) T (◦ C) t (min) Conversion (mol%) Yield (mol%) TOF (h−1 ) TON

– Aldehydes C6, C7, C9 50 80 15 25.6 20 803 201


NaCl Aldehydes C6, C7, C9 50 80 15 67.8 66 2607 652
– Aldehydes C6, C7, C9 50 80 30 49.9 46 905 453
NaCl Aldehydes C6, C7, C9 50 80 30 86.3 85 1675 838

Conditions: agitation = 500 rpm.


Composition: NaCl/Ru = 70 (184 mg NaCl), TPPTS/Ru = 4, aldehydes/Ru = 1000, 6 mL aldehydes, 9 mL toluene and 15 mL deoxygenated distilled demineralised water.

Please cite this article in press as: N.C. Kokkinos, et al., Two-step conversion of LLCN olefins to strong anti-knocking alcohol mixtures
catalysed by Rh, Ru/TPPTS complexes in aqueous media, Catal. Today (2014), http://dx.doi.org/10.1016/j.cattod.2014.07.058
G Model
CATTOD-9236; No. of Pages 7 ARTICLE IN PRESS
6 N.C. Kokkinos et al. / Catalysis Today xxx (2014) xxx–xxx

Table 7
Effect of reaction time in the hydrogenation of the hydroformylated LLCN catalysed by Ru/TPPTS complex in water–toluene system.

Time (min) Total aldehydes (mol%) Aldehydes (n/iso) Total alcohols (mol%) Alcohols (n/iso) Conversion (mol%) Yield (mol%) TOF (h−1 ) TON

60 89.86 0.11 8.98 0.18 8.3 7 72 72


120 75.87 0.11 22.41 0.14 22.5 21 104 208
240 28.53 0.09 69.04 0.13 70.9 69 170 682

Conditions: T = 90 ◦ C, PH2 = 75 bar, agitation = 600 rpm.


Composition: 7.5 mL hydroformylated LLCN, 7.5 mL toluene, 15 mL deoxygenated distilled demineralised water.

from 2 h (22.5 mol%) to 4 h (70.9 mol%), despite the complexity of [12] P.J. Baricelli, E. Lujano, M. Modroño, A.C. Marrero, Y.M. Garcıı́a, A. Fuentes,
the mixture. R.A. Sánchez-Delgado, Rhodium-catalyzed hydroformylation of C6 alkenes and
alkene mixtures – a comparative study in homogeneous and aqueous-biphasic
media using PPh3 , TPPTS and TPPMS ligands, J. Organomet. Chem. 689 (2004)
3782–3792.
4. Conclusion
[13] M.M. Alonso, V. Guanipa, L.G. Melean, M. Rosales, A. Gonzalez, P.J. Baricelli,
Catalytic activity of the RhH(CO)(TPPTS)3 precursor in the biphasic hydro-
RhCl3 /TPPTS and RuCl3 /TPPTS catalysts were revealed as effec- formylation reaction of C5–C7 alkenes from a real naphtha cut, Appl. Catal.
tive catalytic systems for the first and the second part, respectively, A: Gen. 358 (2009) 211–214.
[14] N.C. Kokkinos, A. Lazaridou, C.E. Papadopoulos, N. Nikolaou, N. Psaroudakis,
of the currently proposed two-step LLCN upgrade process in aque- K. Mertis, An environmentally friendly potential refinery process of light-light
ous media. In the first step (hydroformylation reaction) the highest cracked naphtha olefins conversion to valuable oxygenated products, in: Inter-
conversion of 95.4% of the olefins present in a real LLCN was national Conference on Chemical Engineering (CET/CEN2011), 2011.
[15] N.C. Kokkinos, E. Kazou, A. Lazaridou, C.E. Papadopoulos, N. Psaroudakis, K.
observed at 70 ◦ C, 100 bar, at a short reaction time (6 h). The sec- Mertis, N. Nikolaou, A potential refinery process of light–light naphtha olefins
ond step (hydrogenation reaction) was significantly boosted in the conversion to valuable oxygenated products in aqueous media – Part 1:
presence of salt and 66% of the aldehydes were hydrogenated to biphasic hydroformylation, Fuel 104 (2013) 275–283.
[16] E. Fache, C. Santini, F. Senocq, J.M. Basset, Homogeneous catalysis in water. Part
strong anti-knocking alcohols within the first 15 min of the reaction III. The catalytic hydrogenation of propionaldehyde with (RuCl 2L 2) 2, RuHClL
(TOF = 2607 h−1 ). The ultimate upgraded LLCN could potentially 3, J. Mol. Catal. 72 (1992) 337–350.
replace the harmful gasoline ether oxygenates (MTBE, ETBE, TAME) [17] J.M. Grosselin, C. Mercier, G. Allmang, F. Grass, Selective hydrogenation of ␣,␤-
unsaturated aldehydes in aqueous organic two-phase solvent systems using
from the refinery blended gasoline pools with an in situ and an ruthenium or rhodium complexes of sulfonated phosphines, Organometallics
environmentally friendly catalytic process. Further research, with 10 (1991) 2126–2133.
encouraging preliminary results, is in progress not only on the field [18] M. Hernandez, P. Kalck, Water-soluble ruthenium complexes contain-
ing tris(m-sulfonatophenyl) phosphine (TPPTS). Preparation of a series of
of properties optimisation of the ultimate fuel, but also on one-
[Ru(H)((6-arene)(TPPTS)2 ]Cl complexes, [Ru(H)2 (CO)(TPPTS)3 ] and revisited
pot synthesis of strong anti-knocking alcohol mixtures from olefins procedures for previously described ruthenium-TPPTS compounds, J. Mol.
through Rh/Ru tandem catalysis [40]. Catal. A: Chem. 116 (1997) 117–130.
[19] M. Hernandez, P. Kalck, Study of the hydrogenation of ␣,␤-unsaturated car-
bonyl compounds catalyzed by water-soluble ruthenium–TPPTS complexes, J.
Acknowledgments Mol. Catal. A: Chem. 116 (1997) 131–146.
[20] S. Fujita, Y. Sano, B.M. Bhanage, M. Arai, Supported liquid-phase catalysts con-
taining ruthenium complexes for selective hydrogenation of ␣,␤-unsaturated
This research has been co-financed by the European Union aldehyde: importance of interfaces between liquid film, solvent, and support
(European Social Fund – ESF) and Greek National Funds through for the control of product selectivity, J. Catal. 225 (2004) 95–104.
[21] S.-i. Fujita, S. Akihara, M. Arai, Recyclability of water-soluble
the Operational Program “Education and Lifelong Learning” of the
ruthenium–phosphine complex catalysts in multiphase selective hydro-
National Strategic Reference Framework (NSRF) – Research Fund- genation of cinnamaldehyde using toluene and pressurized carbon dioxide, J.
ing Program: THALES. Investing in knowledge society through Mol. Catal. A: Chem. 249 (2006) 223–229.
[22] A.W. Heinen, G. Papadogianakis, R.A. Sheldon, J.A. Peters, H. van Bekkum, Fac-
the European Social Fund. Special thanks are also due to Prof. G.
tors effecting the hydrogenation of fructose with a water soluble Ru–TPPTS
Papadogianakis for his valuable contribution and consultancy on complex. A comparison between homogeneous and heterogeneous catalysis, J.
the research field of aqueous biphasic catalysis and to Dr. A. Chatzi- Mol. Catal. A: Chem. 142 (1999) 17–26.
gakis (Head of Aspropyrgos Refinery Chemical Laboratory, Hellenic [23] S. Tilloy, H. Bricout, E. Monflier, Cyclodextrins as inverse phase transfer cata-
lysts for the biphasic catalytic hydrogenation of aldehydes: a green and easy
Petroleum S.A.) for his support and granting of LLCN samples. alternative to conventional mass transfer promoters, Green Chem. 4 (2002)
188–193.
[24] N.C. Kokkinos, A. Lazaridou, N. Nikolaou, G. Papadogianakis, N. Psaroudakis,
References A.K. Chatzigakis, C.E. Papadopoulos, Hydrogenation of a hydroformylated naph-
tha model (mixture of specific aldehydes) catalysed by Ru/TPPTS complex in
[1] J. Manassen, in: F. Bassolo, R.L. Burwell (Eds.), Catalysis: Progress in Research, aqueous media, Appl. Catal. A: Gen. 363 (2009) 129–134.
Plenum Press, London, 1973, pp. 177, 183. [25] N.C. Kokkinos, A. Lazaridou, C.E. Papadopoulos, N. Nikolaou, N. Psaroudakis, K.
[2] F. Joó, M.T. Beck, Vízoldható foszfinkomplexek redoxireakciói - új lehetőségek Mertis, Aqueous biphasic hydrogenation of the aldehyde content of a hydro-
a homogén oldatkatalízisben, Magy. Kem. Foly. 79 (1973) 189–191. formylated real naphtha cut to valuable oxygenated products, in: International
[3] E.G. Kuntz, FR Patent 2,314,910 (1975). Conference on Advanced Research in Scientific Areas (ARSA-2012), 2012.
[4] E.G. Kuntz, FR Patent 2,338,253, 2,349,562, 2,366,237, 2,733,516 (1976). [26] L. Ongkiehong, Doctoral Dissertation, Technical College of Eindhoven, Amster-
[5] F. Joó, M.T. Beck, Formation and catalytic properties of water-soluble phosphine dam, 1960.
complexes, React. Kinet. Catal. Lett. 2 (1975) 257–263. [27] L. Ongkiehong, in: R.P.W. Scott (Ed.), Gas Chromatography, 1960, Butterworths,
[6] F. Joó, Z. Tóth, Catalysis by water-soluble phosphine complexes of transition London, 1960, pp. 7–15.
metal ions in aqueous and two-phase media, J. Mol. Catal. 8 (1980) 369–383. [28] R.A. Dewar, The flame ionization detector a theoretical approach, J. Chromatogr.
[7] B. Cornils, Hydroformylation, in: J. Falbe (Ed.), New Syntheses with Carbon 6 (1961) 312–323.
Monoxide, Springer-Verlag, Berlin, 1980. [29] L.S. Ettre, Relative response of the flame ionization detector, J. Chromatogr. 8
[8] H. Bach, W. Gick, E. Wiebus, B. Cornils, Preprints International Congress Catal- (1962) 525–530.
ysis, 1984, pp. 417. [30] A. Nosal, J. Skarżewski, Gas chromatographic analysis of 1,4-naphthoquinones,
[9] B. Cornils, J. Falbe, Catalytic reactions in freshwater, in: 4th International Sym- Chromatographia 20 (1985) 19–22.
posium on Homogeneous Catalysis, 1984, p. 487. [31] D.F. Shriver, M.A. Drezdzon, The Manipulation of Air-Sensitive Compounds,
[10] C.W. Kohlpaintner, R.W. Fischer, B. Cornils, Aqueous biphasic catalysis: John Wiley and Sons, New York, 1986, pp. 326.
Ruhrchemie/Rhône-Poulenc oxo process, Appl. Catal. A: Gen. 221 (2001) [32] E.G. Kuntz, O.M. Vittori, Redox chemistry of Pd2+ , Pt2+ , Rh3+ -TPPTS systems in
219–225. water: pH influence on the preparation of low valent TPPTS complexes, J. Mol.
[11] B. Cornils, W.A. Herrmann, Homogeneous catalysts and their heterogenization Catal. A: Chem. 129 (1998) 159–171.
or immobilization, in: B. Cornils, W.A. Herrmann (Eds.), Applied Homogeneous [33] I.T. Horváth, R.V. Kastrup, A.A. Oswald, E.J. Mozeleski, High-pressure NMR stud-
Catalysis with Organometallic Compounds, Wiley-VCH, Weinheim, 2002, pp. ies of the water soluble rhodium hydroformylation system, Catal. Lett. 2 (1989)
603–633. 85–90.

Please cite this article in press as: N.C. Kokkinos, et al., Two-step conversion of LLCN olefins to strong anti-knocking alcohol mixtures
catalysed by Rh, Ru/TPPTS complexes in aqueous media, Catal. Today (2014), http://dx.doi.org/10.1016/j.cattod.2014.07.058
G Model
CATTOD-9236; No. of Pages 7 ARTICLE IN PRESS
N.C. Kokkinos et al. / Catalysis Today xxx (2014) xxx–xxx 7

[34] B. Cornils, W.A. Herrmann, I.T. Horváth, W. Leitner, S. Mecking, H. Olivier- [38] K. Nuithitikul, M. Winterbottom, Crucial parameters in the selective biphasic
Bourbigou, D. Vogt, Multiphase Homogeneous Catalysis, Wiley-VCH Verlag hydrogenation of cinnamaldehyde by biphasic Ru-TPPTS and RhCl(TPPTS)3 cat-
GmbH, Weinheim, 2005. alysts, Catal. Today 128 (2007) 74–79.
[35] I.T. Horváth, J.M. Millar, NMR under high gas pressure, Chem. Rev. 91 (1991) [39] E. Fache, F. Senocq, C. Santini, J.-M. Basset, Homogeneous catalysis in water:
1339–1351. a remarkable salt effect in the hydrogenation of propionaldehyde with
[36] U.J. Jáuregui-Haza, E.J. Pardillo-Fontdevila, A.M. Wilhelm, H. Delmas, Solubility RuCl2 (tppts)3 and related complexes [tppts =(m-NaSO3 C6 H4 )3 P], J. Chem. Soc.
of hydrogen and carbon monoxide in water and some organic solvents, Latin Chem. Commun. (1990) 1776–1778.
Am. Appl. Res. 34 (2004) 71–74. [40] K. Takahashi, M. Yamashita, T. Ichihara, K. Nakano, K. Nozaki, High-yielding
[37] X. Liu, F. Kong, X. Zheng, Z. Jin, Polyether triaryl phosphine oxides for hydro- tandem hydroformylation/hydrogenation of a terminal olefin to produce a lin-
formylation of oleyl alcohol in micellar catalysis, Catal. Commun. 4 (2003) ear alcohol using a Rh/Ru dual catalyst system, Angew. Chem. Int. Ed. 49 (2010)
129–133. 4488–4490.

Please cite this article in press as: N.C. Kokkinos, et al., Two-step conversion of LLCN olefins to strong anti-knocking alcohol mixtures
catalysed by Rh, Ru/TPPTS complexes in aqueous media, Catal. Today (2014), http://dx.doi.org/10.1016/j.cattod.2014.07.058

You might also like