Вы находитесь на странице: 1из 40

Applied Catalysis A: General, 105 (1993) l-40 1

Elsevier Science Pubhahers B.V., Amsterdam

APCAT A2612

Review

Dimerization of ethylene to butene-1 catalyzed by


Ti(OR’)4-A1R3

Abdul Wahab Al-Sa’doun


Research Department, SABIC Industrial Complex for Research and Development, P. 0. Box
42503, Riyadh 11551 (Saudi Arabia)
(Received 20 January 1993)

Abatract

The catalytic dimerization of ethylene stands as the moat selective and economical route to polymeri-
zation grade butene-1. This approach is currently exemplified by the IFFGABIC process for the produc-
tion of butene-l wherein ethylene is dimerized over the modified Ziegler-Natta homogeneous catalysts
Ti(OR’),-AIR, (R’ =C&,alkyl or aryl, R=C,-CBalkyl). The mass of thepublishedpatentsand the
technical literature dealing with the ethylene dimerization over various transition metal based com-
plexes is overwhelming. Therefore, the literature taken into account in this review was selected in order
to give more weight to fundamental studies and less to patent claims. This paper is dedicated to dis-
cussing the chemical perspectives of the dimerization reaction over the titanium-based catalytic system
with its main objective being to get a more decisive picture on the performance of these catalysts. This
review paper identifies the components of this homogeneous system and discussea, in fair detail, the
factors controlling its selectivity to butene-1. Other different themes covered include: structure of the
active species, kinetics, mechanistic considerations and the principal reaction parameters. Additionally,
this eurvey is supplemented by a summarized account for the corresponding supported titanate catalysts
and their usage as a dimerlzation component in the dual functional catalytic systems.

Key words: aluminium/titanium molar ratio; butene-I; ethene dimerization; titanate complex;
triethylaluminium; Ziegler-Natta catalyzta

CONTENTS

1 Introduction ...............................................................................................................................
2
2 Industrial routes to butene-l .................................................................................................... 3
3 The products spectrum of the ethylene dimerization reaction .............................................. 4
4 Components of the dimerization catalyst ................................................................................ 5

Correspondence to: Dr. A.W. Al-Sa’doun, Research Department, SABIC Industrial Complex for
Researchand Development, P.O. Box 42503, Riyadh 11551, Saudi Arabia. Tel. ( +966-l )4985555,
fax. ( +966-1)4980398.

0926-860X/93/$06.00 0 1993 Elsevier Science Publishers B.V. All rights reserved.


2 A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40

4.1 Transition metal compounds ............................................................................................... 6


4.2 Co-catalysts ........................................................................................................................... 6
5 Factors governing the selectivity of the catalyst ..................................................................... 7
5.1 Chain-transfer frequency ..................................................................................................... 7
5.2 Ethylene conversion ............................................................................................................. 8
5.3 Catalyst pre-treatment ......................................................................................................... 8
6 Structure of the catalytic species .............................................................................................. 10
7 The reaction parameters ........................................................................................................... 12
7.1 Al/Ti molar ratio .................................................................................................................. 12
7.2 Reaction temperature ........................................................................................................... 14
7.3 Nature of modifier ................................................................................................................ 15
7.4 Ethylene pressure ................................................................................................................. 17
7.5 Nature of coordinated ligands ............................................................................................. 18
7.6 Nature of the triaIkyiaIuminum .......................................................................................... 19
7.7 Concentration of the titanate complex ............................................................................... 21
7.8 Nature of solvents ................................................................................................................. 21
7.9 Reaction time ........................................................................................................................ 22
8 The kinetics of the dimerization reaction ............................................................................... 23
9 Mechanistic considerations ...................................................................................................... 27
9.1 The bimetallic intermediate mechanism ............................................................................ 27
9.2 The cyclic intermediate mechanism ................................................................................... 29
10 The heterogeneous titanate systems ........................................................................................ 30
10.1 The supported titanate catalysts ....................................................................................... 30
10.2 The titanate catalyst in dual functional systems ............................................................. 32
11 Outlook ....................................................................................................................................... 35
12 Summary .................................................................................................................................... 36
Acknowledgements ......................................................................................................................... 36
References ....................................................................................................................................... 36

1. INTRODUCTION

Butene-1 is a basic petrochemical building block of captive requirements.


Not only can it be converted to polybutene-1 and butylene oxides, but its larg-
est utilization is as a co-monomer with ethylene for the production of higher
strength and higher stress crack resistance polyethylene resins (LLDPE and
HDPE). This share accounts for approximately 80% of the butene-1 market
(Fig. 1) and is projected to grow by 7-9% annually over the next few years with
the worldwide demand exceeding 1.2 billion pounds by 1995 [ 11.
This market trend has been the driving force behind the multiple endeavors
to develop an industrial process that is capable of manufacturing, at relatively
low pressure, a polymerization grade butene-1.

2. INDUSTRIAL ROUTES TO BUTENE-I

Butene-1, a versatile chemical intermediate to a wide variety of industrial


products, is the first member of the even-numbered linear 1-alkenes which
have diversified applications [ 21. The major industrial routes for butene-1 are:
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 3

n Valeraldahyddr

n 0th~~ (mwwptans, rsobutyl


phwok and lolventa

n Polybutmel

jj$, Butylsne ox,de

El Polyathyi en0 comonomer


I

Fig. 1. The industrial uses of butene-1.

1. Refinery operations of crude oil: About 15% of the generatedeffluents


from the flmd catalytic cracker in the refineriesis butene-1.
2. Steam cracking of C, hydrocarbon streams: Approximately 27% of the
butadiene-freeC, fraction from steam crackingrepresentbutene-1.
3. Ethyleneoligomerizationprocesses:Butene-l is producedas a by-product
in the ethylene-basedoligomerizationprocesses [ 3 1. Dependingon the type of
the catalytic systems utilized, the three commercial approaches for linear l-
alkenesare:
(i) ZkgZcr technology. In this approach, a catalytic (Gulf-Chevron Process)
or stoichiometric (Ethyl Process) amount of triethylaluminumis utilized to
chain-growethyleneand simultaneouslydisplacethe resultantlinear1-alkenes.
(ii) Nickel bused technology. This approach is practiced in the SHOP process
(Shell Higher Olefin Process) which produces a mixtureof 1-alkenesby step-
wise chain growth of ethylenein the presenceof a posphinatednickel chelates
homogeneouscatalyst [4]. Shell does not sell butene-l in the open merchant
market, as it is used internallyfor the metathesisof the CzO+into the indus-
trially desiredlower linear 1-alkenecuts.
(iii) Zirconium based technology. This approach is practiced in the Idemitsu
Process wherein a zirconium chloride in association with organoaluminum
compounds catalyzesthe ethyleneoligomerizationto linear 1-alkenes [ 51.
4. Ethylene dimerizationprocess: The march in the catalytic dimerization
of ethylene into butene-l was pioneered in 1952 by the systematic studies of
Ziegler which were originally aimed at producing higher-chainpolymers via
the growthreaction of the organoaluminumcompounds (multipleinsertionof
ethylene into the Al-C bonds). One particularbatch gave the opposite result,
namely the quantitativeformation of butene-l from ethylene. The cause be-
hind this unexpected result was attributed to the adventitiouspresence of a
traceamountof nickelsalt detectedin the autoclavewhichwas madeof chrome-
nickel steel and, in contrast to the normal practice, had been cleaned with
4 A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40

TABLE 1

Global butene-1 production capacity, 1993 (thousand metric tons)

Source of butene-l North America Europe Rest of the world Total

Refinery and 140 120 40 300


steam cracking
Ethylene 100-115 50-69 20-25 170-210
oligomerization”
Ethylene 105 105
dimerization

a A range for butene-l production capacity has been estimated due to the product distribution
flexibility of the various 1-alkene processes.

nitric acid [ 61. This field was later on the subject of interestfor researchersin
industry and academia,which ultimatelypaved the way to the recent devel-
opment of the IFP process for the selectivedimerizationof ethyleneto butene-
1 over a homogeneoustitaniumbased catalytic system [ 71. Since 1987,exten-
sive process modification, contributed by SARIC and IFP, enabled the smooth
runningof the first and the world largestplant at Petrokymea,Saudi Arabia,
which has a nameplatecapacity of 56699 metric tons/year. As a resultof that
collaboration,the two partiesjointly own this technologywhich will from now
on be referredto as IFP-SARIC technologyfor butene-1production.Today, there
are seven licensedbutene-1 plants with three of them alreadygone on stream
in Thailand, China and India [ 81.
The breakdown of the 1993 global butene-l production capacities (gener-
ated from the aforementionedindustrialroutes) is presentedin Table 1.

3. THE PRODUCT SPECTRUM OF THE ETHYLENE DIMERIZATION REACTION

The exposure of two moles of ethylene to a modified Ziegler-Natta (Z/N)


catalystresultsin the quantitativeformation of butene-1togetherwith a small
amountof highermolecularweightby-products (the quantityof whichis highly
dependenton the catalystemployed and the reactionconditions) accordingto
the following equation:

Catal.
2CHz==CHz
- CH@H;$H-CHI + by-products (3.1)
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 5

TABLE 2

The hydrocarbon product’~~composition of ethylene dimerixation over Ti (OBu),-AlEt, [ 91


Reaction conditions: Al/Ti=3, T=55”C, PC, =22 atm

Total hydrocarbon (wt.-%) Relative composition (wt.-%)

C1 91.7 Butene-1 99.8


Butane 0.2
CB 8.0 3-Methylpentene-1 25.5
1-Hexene 6.4
2-Ethylbutene- 1 63.9
2-Hexene 2.6
3-Methylpentene-2 0.1
3-Methylpentadiene 0.1
1,4-Hexadiene 0.9
2,4-Hexadiene 0.5
C8 0.3

The formation of these by-products, which complicates the efficacy of an


industrialprocess, can be minimizedby the judicious choice of the catalystand
the reaction conditions. The major hydrocarbon products of the ethylene di-
merixationreaction are listed in Table 2 [91. In additionto the listedproducts,
a trace amount of polyethyleneis often generated.
Several literaturereports claimed that the selectivity of the titanate cata-
lysts to butene-l may exceed 98 wt.-% throughthe proper control of the reac-
tion parameters [ 10,111.However,one should not rule out the likelihoodthat
this selectivityis in fact accounting for the presence of a mixtureof butene-l
(in a selectivity not exceeding 92 wt.-% ) and isohexenes (predominantly 2-
ethylbutene-1and 3-methylpentene-l ).

4. COMPONENTS OF THE DIMERIZATION CATALYST

The homogeneousZ/N catalytic system, comprised of titanium(IV) alkox-


ides in association with trialkylaluminum,was originally envisioned for the
production of high-molecular-weightpolyethylene [ 121, styrene [ 131, and
conjugateddiene [ 14-161. These catalysts have repeatedlybeen the subject of
modification aiming at the selectiveproduction of oligomeric [ 17-211 and di-
merit products [ 22,231. This versatilityof the Z/N catalyst indicatesthat the
ethylene dimerixationand/or oligomerizationcould be regardedas (degener-
ate) polymerizationreactions [ 241. It could also be interpretedas an indica-
tion for the coexistence of a competing march between the dimerixation,oli-
gomerixation,and polymerizationreactions.However,which of thesereactions
prevailsis dependent on the nature of the catalyst’s metal, its oxidation state
and the reaction conditions employed [ 251.
6 A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40

TABLE 3

Performance of selected Group IVB transition-metal complexes as catalysts for the ethylene di-
merization [ 281
Reaction pressure=21 atm, Al/Ti=3.2, Al/THF=2.0, t=60 min, toluene, M=Ti or Zr; co-
catalyst = AlEts

Catalyst Temperature Cat. cont. Conversion to C, Selectivity to butene-l


(“C) (mmol/l) (g C&r M. h) (wt.-%)

Ti(OBu)l 50 6.9 562 000 97.3


Ti(OEt)l 50 6.9 162 000 96.6
TiCl, 22 5.2 6251 6.6
Cp,TiCl, 8 5.2 14 2.2
Zr(OPr), 50 6.9 5579 95.3

4.1. Transition metal compounds

Group VIII transition metals are widely known to favor the B-hydrogen
transfer (abstraction) and thereforealkene dimerixationand oligomerization
catalysts are observed preferentiallywithin these metal complexes. In con-
trast,with Group IV to VI transitionmetalspropagationprevailsand predom-
inantly polymerizationcatalysts are found in this class. Moreover, with the
same metal, chain propagation decreases with increasingoxidation state. It
must be stressedthat this classificationis made merelyon empiricalbasis and
severalexceptionsare known. Further,the metal charactercan be modXed by
adding ligands. Donor ligands tend to increase, acceptor ligands to decrease
the chain propagation [ 241. On the basis of this account, complexes of tita-
nium and zirconium have been found to be the most effective dimerization
catalysts [26-281. As illustrated in Table 3, the catalysts Ti (OR’ )&lI&
(R’ = C1-Cs alkyl or aryl, R= C1-C6alkyl) have been ascertainedto be “the”
catalysts of choice for the selective dimerization of ethylene into butene-1,
primarilybecause: (i) they have a very low isomerizationactivity (very small
amounts of butene-2 are formed) and; (ii) their specific activity is relatively
high [ 23,271. In connection with the latter feature, it has been reported that
these catalysts demonstrate a remarkableactivity for dimerixingethylene to
butene-1, with a conversion exceeding 184 999 mol of ethylene/m01catalyst
per hour [291.

4.2. Co-catalysts

The co-catalyst utilized in the ethylene dimerixationis commonly a Lewis


acid (typicallya trialkylaluminumcompound) [ 231. The multiplerolesplayed
by the aluminumalkyl in the dimerixationreaction are: (i) the releaseof free
coordination sites in the titanate complex; (ii) the withdrawalof the electron
A. W. Al-Sa’doun / Appl. Cat& A 105 (1993) l-40 7

density surroundingthe titanium metal center (the reduction of the valency


of the titaniumcenter) and; (iii) the generationof one or more Ti-C bonds by
exchangingits alkyl groups with the alkoxide groups of the titanate complex
[ 231. It is interestingto note that the aluminumsite and the ligand attached
to it remainpractically inert duringthe reaction [ 29,301.
It is importantto emphasizethat no other alkylating/reducingagent (even
those with metalsbelongingto the GroupIRA such as GaE&or I-), matches
the performanceof trialkylahnninumas a co-catalyst for the titanate precur-
sor. This has been attributedto the tendency of these organo-aluminumcom-
pounds to adopt a bridging association, owing to the bonding characteristics
of the aluminumatoms dictated by their electron deficiency [ 311. This theo-
reticalconsiderationis in agreementwiththe findingsof a researchstudywhich
revealedthat the employmentof other reducingagents (such as Mgh or BuLi)
for the alkylation of the titanate complex did not compensate for the role of
ALEt,in generatingan active dimerizationspecies, suggestingthat the latter
must be part of the active site [ 321.
Moreover,there have been some reportswhich suggestedthat the aluminum
alkyls influence the course of the dimerization reaction, for instance, an in-
crease in the acidity of these compounds causesa decreaseof the chain growth
and vice versa [ 23,281.

5. FACTORS GOVERNING THE SELECTIVITY OF THE CATALYST

In addition to the reactionparameters,the followingfactors affect the selec-


tivity of the Ti (OR)4-AlRg systemsto butene-1:

5.1. Chain-transfer frequency

For a particularethylene dimerizationreaction, irrespectiveof the utilized


catalyst, it is highly desirableto have the rate of the chain-transferpredomi-
nant with respect to that of the chain-propagation.This impliesthat the level
of the by-products formed in the dimerization reaction is related to the fre-
quency of the hydrogen transfer with respect to the chain-propagationreac-
tions [ 33 1. This frequencyis in turn dependenton the conditions underwhich
the hydrogenatoms are shifted within the chain-growthproduct [ 23,25,27]. It
has been reported that this frequency is associated with the electron affinity
of the catalyst’smetal center,hence, it can be controlledby thejudicious choice
of the coordinated ligands [25]. In addition, it has been suggestedthat the
chain-transferdomination can be achieved by a balanced increase of the re-
action temperatureso as to activate the C-H bond, a necessity for the occur-
rence of the hydrogen-transferreaction [23]. However, in view of the high
activation energy of the hydrogen displacement,that reaction should not be
consideredas a simple electronic transfer.
8 A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40

5.2. Ethylene conversion

A full control of the chain-growth could only be achieved when the generated
products display a sufficiently low coordination affinity to the catalyst’s metal
center [ 231. This has not been always attainable, especially when these prod-
ucts exhibit little or no branching, such as for instance the formed butene-1
which possesses sufficient reactivity to co-dimerize with ethylene at the active
centers [23]. One way of preserving the catalyst’s selectivity to dimers is by
using high concentration of ethylene so that its conversion occurring on the
active site will be limited to dimerization. Another advantageous aspect of
working with higher ethylene concentration is related to the fact that there is
less chance of isomerizing butene-1 to butene-2, which often becomes more
facile when the former coordinates to the titanium metal center. This could be
partly attributed to the steric hindrance arising from the concomitant coordi-
nation of two alkene molecules at the active metal center [ 23,251. This expla-
nation has been validated by a research study which reported a marked inhi-
bition in the isomerization of the produced butene-l into butene-2 when excess
ethylene was maintained in the reaction zone [ 341. This trend was ascribed to
the reactivity of ethylene which exceeds that of butene-1 by 2-3 orders of mag-
nitude, imparting that the former will react with the active titanium species
once they are introduced to the reaction medium.

5.3. Catalyst pre-treatment

It is known that the conditions under which the catalyst is prepared have a
considerable impact on the concentration and the type of the intermediate
complexes taking part in the formation of the active dimerization species [ 351.
This account is based on the outcome of a research study which noted a marked
increase in the butene-1 yield when the catalyst’s components were mixed in
the presence of ethylene [ 361. That verdict has been consolidated by the ob-
servation of a sharp reduction in the proportion of polyethylene coupled with
an increase in the productivity of the Ti ( O-iPr)l-AIEta catalyst upon mixing
its components in the presence of ethylene [ 37,381.
Not only ethylene exerts the aforementioned influence on the catalyst’s se-
lectivity; pre-treatment of the Ti(OBu),-Al (iBuS) catalyst with a balanced
amount of hydrogen enabled the generation of butene-l in a selectivity of 92
wt.-% (at T=60”C, Pm =2 atm, Pc, = 2 atm, Al/Ti = 5.5,800 ml dichloroe-
thane ) [ 391. These observations agree with a relevant report which noted an
acceleration of the dimerization rate when either hydrogen or a hydrogen do-
nor was present inside the reactor upon mixing the catalyst components [ 231.
This profile was ascribed to the fact that hydrogen may act as a transfer agent
through hydrogenolysis of the Ti-C bond, thus enabling the liberation of the
dimer. Along this line, pre-treatment of the Ti (OBu),AlE& catalyst with hy-
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 9

drogen resultedin a shorter induction period coupled with an accelerateddi-


merizationrate [34]. However, it has been noted that the catalyst pre-treat-
ment with hydrogen was associated with a detrimental increase in the
proportion of ethane, butane and isomeric butene-2 mixture.
Within this context, Belov et al. [40] reportedthat the pretreatment of the
Ti ( OBu),-AlEt catalystwith an ethylene-hydrogenmixture,ensuredhigher
yield and selectivity to dimer and trimer with a significant reduction in the
proportion of the generatedpolymer. The authors have also noticed that the
product distribution (dim& vs. trimer ) was found to be dependenton the cat-
alyst pretreatment conditions, with the highestbutene-l yield obtained when
the reactionwas conductedat 27” C, 15 atm and an ethylene-hydrogenmixture
in a ratio of 70 vol.-% to 30 vol.-%, respectively.The products spectrumof that
reactionwas comprisedof 76 wt.-% butene-1,23.9 wt.-% hexenestogetherwith
traces of polymer (0.1 wt.-%) [40]. The results generatedby this study are
presentedin Table 4.
In a different approach, Belov et al. [41] reportedthat the treatmentof the
titanate catalystwith oxygen allows a lo-20% increasein the dimeryield with
a concomitant enhancementof the selectivityto butene-1. The optimal cata-
lytic performancewas achievedwhenthe introducedoxygenconcentrationwas
in the range of 0.2-2.0 equivalentswith respect to trialkylaluminum.By way
of analogyto that approach,treatmentof a heptanesolutionof Ti (OBu), with
air containing500 ppm (vol. ) water for 1 min improvedthe catalyst’s selectiv-
ity to butene-l from80wt.-% to 85wt.-% (at T=70”C,Pc, =2 atm, Al/Ti=5)
[421.

TABLE 4

The impact of pre-treating the catalyst Ti ( OBu),-AlEte with ethylene-hydrogen gas mixtures on
the product selectivity [ 401

Catalyst pretreatment conditions Butene-1 Product selectivity (wt.-%)


yield
Hz T P Time (g/g Ti-hr) Butene-1 Hexenes Polymer
(vol.-%) (“C) (atm) (min)

No treatment 46.0 92.4 4.9 2.7


63 27 2.5 7 277 77.6 22.0 0.4
63 25 10 10 84.0 75.0 24.9 0.1
20 25 10 10 118 89.0 10.8 0.2
20 100 2.5 1 222 89.9 9.9 0.2
95 0.0 10 12 105 69.2 30.6 0.2
5 80 0.2 5 112.5 79.3 20.4 0.3
30 27 15 7 347 76.0 23.9 0.1
10 A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40

6. STRUCTURE OF THE CATALYTIC SPECIES

Despite the different opinions about the exact nature of the active dimeri-
zation species, there is unanimous agreement that the reaction of the titanate.
complex with the trialkylaluminum compound is a complicated one. A variety
of complex species are generated, depending on the ratio of the two catalyst’s
ingredients in solution ( Al/Ti). The different physico-chemical methods em-
ployed to elucidate the structures of those species have proposed various spe-
cies ranging from trivalent to divalent with Al-Ti bridging structures, depend-
ing on the Al/Ti ratio, the reaction time, and the temperature [12,13,15,43-
521. In some studies, the extent of reduction was thought to be small (at low
Al/Ti molar ratio), affording primarily Tin’ species in solution. In other cases
(at high Al/Ti ratio), however, Tin species were considered as a likely product.
Dzhabiev et al. [ 461 detected an electron spin resonance (ESR) signal with
a g-value of 1.957 and a well-resolved hyperfiie of 11 components at Al/Ti
molar ratio of 2. These components were ascribed to the equivalent interaction
of the unpaired spin of titanium with two 27Alnuclei [ 481. Based on these data,
complex ( 1) has been proposed in which the titanium atom was regarded as a
trivalent state (Fig. 2). Hirai et al. [ 441 employed ESR spectrometry to deter-
mine the products over a wide range of Al/Ti ratios. Complex species (2 ) and
(3 ) were proposed to be formed at Al/Ti ratios < 2 and > 3 respectively (Fig.
2 ) . The fivefold coordination of the Ti3+ ion in these complexes may account
for their high instability [ 471.
Christenson et al. [ 511 employed chemical ionization mass spectrometry to

Et

/ \,f \/ ‘Et
Et
BU BU

Complex ( 1) { AVTi = 2.0 )

Bu

Bu

Complex (2) {AUTi<2} Complex (3) {Al/Ti > 3)

Fig. 2. Proposed active species of Ti(OR’ ),-AIR3 generated at different Al/Ti molar ratios as
detected by electron spin resonance spectroscopy.
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 11

detect the structureof the species generatedat different Al/Ti molar ratios.
The authorsproposed the three complex species presentedin Fig. 3.
Differentstructuresfor the intermediatespecieshavebeen proposed, subject
to the Al/Ti ratio. For instance, at rather low Al/Ti molar ratio ( < 2)) only
the pre-active complex (4) was generated [51]. When the Al/Ti molar ratio
is > 2.0, complex (6) with a formal oxidation state of 3 + is generated At even
higher Al/Ti ratio ( > lo), the trimeric complex (6) with a formal oxidation
state of 2 + is formed. At Al/Ti = 10.7, reductionof Tin’ of the titanateto Tim
state occurred,with approximately60% of the titaniumexistingas a Tim spe-
cies [ 121.The Tim species are known to promote the polymerizationreaction
via a process of coordination and repetitiveinsertion of ethylene into the Ti-
C bond [52]. Interestinglyenough, complexes (6) and (6) contains a Ti-C
bond which is believed to be an essentialfeatureof the active catalytic species
for the chain-growthreactions [ 531.
It is known that the dimerizationreaction does not proceed when the AlITi
molar ratio is less than two. Hence, it has been proposed that the primary
catalyticallyactive complex in the reaction of the titanate with the triethyl-
aluminum is complex (5). Within this domain, a recent study utilii the
carbon monoxide poisoning method to estimatethe maximumnumber of ac-
tive dimerizationsites, generatedfrom mixing the two ingredientsin solution
(on the assumptionthat one moleculeof carbon monoxide is adsorbedon each
active site at 30’ C, 1.1 atm, Al/Ti = 5.4)) revealedthat the maximumnumber
of those sites was approximately35% of the total titaniumcontent [ 541.
Beyond an Al/Ti molar ratio of 10.6 and even 13, there may be an agglom-
eration of the catalytic speciesleadingto a gradualdecline in the reaction rate
on one hand and a markedincreasein the level of polymer formedon the other.

R’ R’
R
R’o\ /“I / R
\m/o\m/R
R,,/m\oO”\ R R'O' ‘0’ 'R
R’ R'

Complex ( 4 ) { Al/l? < ‘1.0 } Complex(S) {Ani > 2.0)

Comples (6) {Al/T%> 10)


Fig. 3. Complex species of Ti( OR’ ),-AIR, formed at various Al/Ti molar ratios as detected by
ma88 spectrometry.
12 A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40

\ /
Ti

I
R’
\ =Ti/o\AI/R
/” \R/ \
R

L R=l&;R'=&1

Fig. 4. The inactive hexametric aluminium alkoxide complex.

This profile could be ascribed to the presence of many different inactive Ti-Al
complexes [ 521. Further, the agglomeration products are assumed to arise from
either: the dimerization, trimerization, or tetramerization of the active dimer-
ization species (6 ) . An inactive complex species such as the hexameric Ti-Al
complex (7) presented in Fig. 4, is thought to be generated from the dimeri-
zation of the active complex (6) taking place at that Al/Ti ratio [ 551.
To conclude, it is clear that the activity of the dimerization catalyst is strongly
dependent on the Al/Ti molar ratio [ 561. The relative concentrations of com-
plexes (4)) (6), and (6) in equilibrium at various Al/Ti molar ratios are re-
sponsible for the variation in the catalytic activity. It can also be suggested
that depending on the Al/Ti molar ratio, equilibrium is established in the cat-
alytic solutions between the various complex formulations having different
compositions, structures, hence, different catalytic performances.

7. THE REACTION PARAMETERS

The reaction parameters, which have a manifest influence on the course of


the ethylene dimerization, are:

7.1. TheAlITi molar ratio

The Al/Ti molar ratio is believed to be the most critical factor which deter-
mines the course of the ethylene dimerization over the Ti( OR’ ),-AIR, cata-
lyst. [ 52,57-591. This is in agreement with a long-standing assumption ad-
vanced by Natta [ 601 which suggested that the reaction is selective to butene-
1 when the Al/Ti molar ratio is < 10, whilst a mixture of dimer, oligomer and
polymer are produced at higher ratios, with the balance tipped toward the for-
mation of high-molecular-weight polyethylene when Al/Ti > 20. This has later
been confirmed by several studies which concluded that ethylene dimerizes
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 13

AI ITi MOLAR RATIO

Fig. 5. Effect of Al/Ti molar ratio on the Ti (OR),-AI& performance at 50 ’ C, 20 atm.

selectivelyto butene-l at lower Al/Ti ratios, whereashigherAl/Ti ratios are


associated with a marked loss in the activity of the dimerixationcenters, due
to the presence of free A&&, [ 56,61,62]. In contrast, low Al/Ti ratios are re-
portedly associatedwith a lower ethylene conversion rate whereashigh Al/Ti
ratios are associatedwith a higherethyleneconversion rate [ 521.
Beach and Kissin [38] noticed a pronounced dependence of the catalyst’s
productivityon the Al/Ti molar ratio. This was in contrast to the initial activ-
ity which was found to be independenton that ratio. For instance,at lowerAl/
Ti ratios (Al/Ti= 5-6), the initial activity of the catalyst was 1.15 1 mmol-’
min-‘, compared to a rangeof 0.7-1.3 1mmol-’ min-’ at higherratios. How-
ever, since different Al/Ti ratios are associatedwith different rates of catalyst
deactivation,the total catalytic productivity has been increasedwhen the re-
action was conducted at lower Al/Ti molar ratio. For example,at an Al/Ti of
5-6 (conducting the reaction for 1 h at an ethylene pressureof 6.2 atm) the
amount of the formed butene-l was about 2.5-3.5 mol mmol-‘, dropping to
0.7-1.0 mol mmol-’ at an Al/Ti of 23. This trend has been ascribed to the
considerablysmallerdeactivation rates of the dimerizationactive centers oc-
curringat that low Al/Ti molar ratio [ 381.
The combined impact of the Al/Ti molar ratio and the reactiontemperature
on the catalyst’s performance has been the subject of a study conducted by
Belov et al. [ 631. The authorsreportedthat both the reactiontemperatureand
14 A. W. AI-Sa'doun / Appl. Catal. A 105 (1993) 1-40

the A1/Ti molar ratio determines the alkylation rate of the titanate compo-
nents. An optimal catalyst activity has been detected at A1/Ti range of 3.0-5.0
(subject to the reaction temperature). It was also noticed that no marked in-
crease in the proportion of the generated polymers was detected as the A1/Ti
molar ratio surpassed this range [63 ]. In view of that, it has been suggested
that there exists an optimum range of A1/Ti ratios, where the catalyst activity
remains somewhat reasonable and the level of polyethylene produced is kept
minimal. Studies of the Ti(OBu)4-AIEt3 catalyst defines the boundaries of
this range at 4 < A1/Ti < 6, upon operating at moderate temperature and pres-
sure [63,64]. This correlation is presented in Fig. 5 which illustrates the effect
of different A1/Ti ratios on the catalyst's activity and the polymer proportion.

Z 2. Reaction temperature

Results of the systematic studies conducted by Henrici Oliv~ and Olivd


[21,65] on the ethylene oligomerization catalyzed by Z/N catalysts, have re-
vealed that an increase in the reaction temperature is normally associated with
the formation of higher-molecular-weight products. This has been considered
to be the result of the higher magnitude for the activation energy of the chain
propagation in comparison to that of the chain-transfer reaction (E~<ED)
[21 ]. Under these conditions, the monomer molecules tend to co-dimerize with
the produced dimer giving rise to the formation of a significant amount of
linear and branched higher oligomers. A similar trend has been observed in the
dimerization reactions wherein higher levels of hexenes have been associated
with the temperature increase [52,63 ]. Furthermore, the reaction temperature
is reported to have a controlling influence on the generation of the polar inter-
mediate titanium species. This, in turn, indicates that the temperature influ-
ence on the product molecular weight could be taken as a good evidence for the
ionic character of the active dimerization species [66].
A recent study of ethylene dimerization catalyzed by the Ti (OR')4-All~
systems revealed that the reaction exhibits a strong dependence on tempera-
ture [28]. With an increase in the reaction temperature, the conversion of
ethylene and the selectivity to butene-1 decreases. A similar profile has been
observed by Pillai et al. [52 ] in their systematic study of the titanate catalyst
(Table 5 ).
The decline in the ethylene conversion may be attributed to the decrease in
the monomer solubility at higher reaction temperature. On the other hand, the
poor selectivity to butene-1 could be attributed to the fact that increasing the
reaction temperature (at a constant pressure) provokes an increase in the par-
tial pressure of the formed butene-1 and a concomitant decrease in the ethyl-
ene partial pressure which leads to the subsequent decrease in the concentra-
tion of the latter in the liquid phase. In addition, the higher proportion of
polymer generated at higher reaction temperature could be ascribed to the
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 15

TABEL 5

The modifier’s influence on the selectivity of the catalyst Ti (OR’ ),-A@& [ 521
1 h, C&C 1,=5OO ml, Pc, = 12.5 atm, Al/Ti molar ratio = 8

Temperature Conversion Selectivity (wt.-%)


(“C) (wt.-%)
Butene-1 Hexenes Octenes Polymer

33 97.5 75.3 21.2 2.2 1.3


49 98.1 66.3 31.6 1.0 1.1
62 93.2 64.0 31.8 1.0 3.1

TABLE 6

Selected thermodynamic parameters for ethylene dimerization over Ti(OBu),-AlEh at 26°C


[521

Parameter Heptane G-C,, alkane

AE, (kcal/mol ) 8.73 8.52


AH (kcal/mol) 8.13 8.04
AC (kcal/mol ) 15.24 11.94
AS (cal/K mol) - 24.35 - 13.0

higher deactivationrate of the active dimerizationspecies [ 52,62,67] leading


to the quantitativegenerationof the Tim and Tin species,which are known to
promote the ethylenepolymerization [ 23,651.
Despite the adverse impact of the temperatureascension on the catalyst’s
selectivityto butene-1, Belov et al. [63] noted that the optimal range of the
Al/Ti ratio of 6-10 (at a temperaturerangeof 10-30” C ) could be shifteddown
to 3-4 when the reactionwas conducted at highertemperature( > 50” C ) . This
indicates that increasing the temperatureby 40-50°C scored a 2.5-3.0 fold
reduction in the aluminum alkyl consumption which is positively significant
from a technical and economical perspective [ 631.
The thermodynamic parameters for the ethylene dimerization have been
measuredby Pillai et al. [ 521 and their estimatedvalues are listed in Table 6.
The low values of the enthalpy (AH) and entropy of activation (AS) can be
interpretedas an evidence for the existence of a rigid transition state. More-
over, these values are also indicative of a relatively strong coordination and
steric disposition of the monomer molecule and the intermediatedimer on the
active metal center [ 52,681.
7.3. Nature of modifier
The nature of the modifiers, also known as auxiliaryligands, has a distin-
guishableimpact on the performance of the dimerixationcatalysts. Strongly
16 A. W. AI-Sa'doun / Appl. Catal. A 105 (1993) 1-40

coordinating ligands (hard ligands) undermine the catalyst activity to a sub-


stantial level,whereas weakly chelating ligands have no apparent influence on
the catalyst performance [69].
The incorporation of modifiers within the Ti (OR')4-AIR3 catalyst is essen-
tial because even small amounts of by-products (2-5 wt.-% of the consumed
ethylene) have an adverse effect on the entire course of the dimerization re-
action. Such modifiers are used in an amount commensurate with the catalyst
components (their preferable ratios with respect to Ti and Al are 0.01/10 and
0.01/1.0 respectively). Beside reducing the extent of the polyethylene propor-
tion, by suppressing the active centers responsible for polymerization, the
modifier stabilizes the Ti zv complex, believed to be the active dimerization
species [70 ].It is also known that the utilization of these modifiers diminishes
the catalyst's usual activity [50,71 ].
The typical modifier is a Lewis base or a polar organic compound of a suffi-
ciently soft (non-ionic) nature combined with a mild Lewis basicity to avoid
neutralizing the needed Lewis acidity of the aluminum alkyls co-catalysts [72 ].
Cyclic ethers, particularly tetrahydrofuran (THF), have demonstrated to have
the best characteristics to fulfillthese prerequisites, and accordingly they have
been used together with the titanate catalysts. Being good coordinating li-
gands, the THF molecules are thought to stabilizethe titanate complex via the
formation of a hexacoordinated species which cannot otherwise be generated
[64]. It has also been suggested that the THF/Ti molar ratio plays an impor-
tant role in determining the catalyst's selectivityfor butene-1, with the pref-
erable ratio being 2:1. Higher ratios ( > 10) lead to a considerably slower re-
action associated with a rather poor selectivityto butene-1 [64,71 ].
The data of Table 7, generated by Pillai et al. [52 ] in their systematic study
of the Ti (OR') 4-AIEt3, shows the impact of different modifiers on the product
spectrum of the dimerization reaction. From among the listedmodifiers, THF
stands as the best compromise as it demonstrated a good selectivityto butene-

TABEL 7

T h e modifier's influence o n the selectivity of the catalyst T i ( 0 R ' )4-AIEt3 [52]


Modifier/A1Ets molar ratio = 0.1; Temp. -- 35 ° C, PTot~ = 12 atm, A I / T i molar ratio = 8

Modifier Conversion Selectivity (wt.-%)


(%)
Butene-1 Hexenes Octenes Polymer

97.5 75.3 21.2 2.2 1.3


PPh 3 90.2 67.3 28.0 1.7 3.0
Et3N 80.2 67.2 29.2 3.6
THF 56.9 78.3 17.1 4.6
PhsPO 38.8 84.3 15.7 - -
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 17

1 coupled with a fairly good ethylene conversion rate and above all without
producing a detectable level of polymer.
The above outcome has been manifestedby Nevzorov and Belov [72] in a
recent study which revealed that the inclusion of various organophosphorus
compoundswithinthe Ti ( OR’ )&lEt, catalystresultedin a markedenhance-
ment of its selectivityto butene-1 with the deterringrate of the side-reactions
increasing in the following order: Ph,PO> BusPO>Bu,PSe. Moreover, the
authors noted an optimal performance of these catalysts at Al/P < 1.0 [ 721.
Similarly,the inclusion of tributylphosphineoxide (BqPO) into Ti (OBu)*-
All&, at a molar ratio of P/Ti=3.0, resultedin the formation of butene-l in
a selectivityof 99.5 wt.-% (n-heptane, 50°C 20 atm, and 6.5 Al/Ti) with ap-
proximately0.1 wt.-% polymer [ 731.
Along this line, some studieshave establishedthat the additionoE Cp,TiClz
( Cp=$-C,HS), OS,or (iPr),NH to the Ti(OBu),-AlEt, catalyst resultedin
a marked decrease in the by-products proportion [ 74,751. Similarly,addition
of MgEtBrto the Ti ( OBU)~-A~(iPr ), catalystin a Mg/Ti ratioof 2.0 improved
its selectivityfor butene-l from 87% to 96 wt.-% (at SOOC).It was also noticed
that the proportion of the formed iso-hexenes was roughlyhalf its originally
detected level [ 761.

7.4. Ethylene pressure

The ethylene dimerixationover Ti (OR’ )&lRs is carried out in a liquid


phase, the composition of which is determinedby the ethylenepressurewhich,
in turn, determinesthe ethyleneconversionrate. Higherpressureis associated
with an increasein the ethylene conversion rate [ 521. This account has been
evidencedby the observation of Sergienkoet al. [76] of a higheryield of bu-
tene-1 and iso-hexenesat higherpressure (Table 8).
From the data of Table 8, it can be noted that, with an increasein the total

TABLE 8

Influence of the reactor’s total pressure on the yield and selectivity to butene- 1 [ 76 ]
[Ti] =7.3 mmol/l, Al/Ti=3; T=50”C; time= 1.25 h, PH, = 1.1 atm; 0.5 1 hexane

P Yield (g/g Ti h) Selectivity to


(at& Butene-l (wt.%)
Butene-l 3-Methylpentene-1 2-Ethylbutene-1

1.1 61.8 1.92 3.84 93.8


3.3 339.5 15.4 36.6 90.8
4.4 474.0 21.1 48.0 91.1
5.5 516.1 25.2 56.8 88.9
6.6 741.3 27.5 62.2 88.9
18 A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40

pressure in the reactor from 3.3 to 6.6 atm, there have been an almost doubling
in the yield of butene-l and iso-hexenes. However, such increases have a det-
rimental, although minor, effect on the reaction’s selectivity to butene-1. This
observation could be interpreted as an indication that higher pressure ensures
higher activity of the catalyst due to an improved diffusion of the monomer
through the reaction mixture to the active dimerization sites.
The initial ethylene pressure has apparently no influence on the catalyst’s
selectivity to butene-1, but on the other hand an increase in the level of the
generated polymer was detected at higher reaction pressure [ 52,61,77]. In line
with this, it has been reported that conducting the dimerization reaction at an
ethylene pressure > 3.0 atm did not have any marked effect on the initial rate
of dimerization [ 631, but at a pressure range of 0.5-2.5 atm the initial rate was
directly proportional to the ethylene pressure. Further, it was revealed that
increasing the ethylene pressure from 0.5 to 17 atm was paralleled by an in-
crease in the polymer content from 0.01 to 1.4 wt.-% (at 2O”C, Pc, =4 atm,
Al/Ti=2.5) [63].

7.5. Nature of coordinated 1igand.s

The nature of the ligands coordinated to the titanium center plays a key role
in stabilizing the structure of the active species, hence, influencing their per-
formance. This indicates that the coordinated ligands govern the course of
formation of the ionic transition state via which the chain transfer proceed
[691.
Henrici-OlivB and Olivh [78] noted a significant influence exerted by the
ligand’s nature on the strength of the Ti-C bond and ultimately on the course
of the dimerization reaction. Further, the authors suggested that the growth
step starts from a configuration in which a monomer is coordinated to the
metal in either a six-member (I) or a four-member (II) transition state [ 791.
The generation of any of these intermediates, and subsequently the product
distribution (dimer vs. oligomer ), is thought to be influenced by the nature of
the coordinated ligands, which are generally categorized into:
I. Electron donor Zigands. Donor ligands reduce the positive charge on the
titanium center, hence giving rise to the formation of a less intensive polari-
zation on the adjacent bonds. This imparts that the polarization will be con-
fined to the a-carbon atom, affording the polar four-member transition state
(I) which favors chain growth products as presented in Route-A, Fig. 6.
II. Electron acceptor ligands. Acceptor ligands increase the positive charge
on the titanium center, resulting in a more pronounced polarization of the
adjacent bonds including the /?-hydrogen atom, hence enabling its inclusion
into the polar six-member transition state (II) which promotes the predomi-
nance of p-hydrogen transfer, giving rise to dimers as presented in Route-B,
Fig. 6.
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 19

(n)

Y-- c”2
I I
Lam _-T-R

La-” + q=CE-R
La : E4ectron aa!eptor ligallda

La : R4ectnm donor ligands

Fig. 6. The influence of the &and’s nature on the course of the dimerization reaction.

Several systematic studies have singled out the aliphatic alkoxides (OR),
particularly when R= 0-iPr or 0-nBu, to be the best candidate ligand for the
selective production of butene-1 [ 36,52,64,70]. This trend could be attributed
to the basicity of the alkoxides which renders the titanium center more elec-
trophilic, hence assist in the formation of the polar active dimerixation sites.
Beside their electronic influence, the large steric effects of these ligands have
been suggested to be the reason behind their selectivity for butene-1 [391.
As illustrated in Table 9, it is interesting to note that the presence of a che-
lating ligand in the titanate complex, such as acetyl acetonate (acac) in the
case of (acac),TiCl, and its ethoxide counterpart (acac)zTi(OEt)z, is often
associated with a poor performance of the complex as a dimerization catalyst
most probably due to the fact that these ligands makes the catalytic sites in-
accessible to the co-catalyst.

7.6. Nature of the triulkylaluminum

Since the nature of the alkylaluminum compounds (Al&) is largely gov-


erned by the nature of their alkyl group, the rate of the dimerization reaction
20 A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40

TABLE 9

The effect of the ligand’s nature on the butene-1 yield [37]


T= 9O”C, P= 7.1 atm, acac = 2,4-pentanedionate, Tol. = CBH,CHB

Catalyst Al/Ti Butene-l yield


(mol/mol) (g/g Ti h)

Ti(OiPr), 5.6 2134.0


Ti(NMez)l 2.9 82.0
(acac)zTiClz 4.9 264.5
(acac),Ti(OEt), 6.1 156.0
Ti(OTol.), 3.7 1450.5

TABLE 10

Influence of the co-catalyst type on the product selectivity [ 761


CataIyst=Ti(OBu),, T=30”C,P=l atm, hexane, t=2 h

Co-catalyst Conversion Product selectivity (wt.-%)


(wt.-%)
Butene- 1 iso-Hexenes Polymer

AiEt, 46.6 65.0 32.9 2.1


AlPr3 53.5 56.3 41.5 2.2
AiBu, 53.1 59.5 39.2 1.3
AI(Bi&H 81.9 95.5 4.5

was found to be facile when R is a short chain aliphaticfragment [ 23,251.This


is in agreementwith the data obtainedby Sergienkoet al. [ 761 which are listed
in Table 10.
These results indicate that dibutylaluminumhydride in combination with
the titanatecomplex is the best choice from the viewpoint of ethyleneconver-
sion. However,this was underminedby the formation of a substantiallevel of
polymericmaterial.Thus, it appearsthat triethylaluminumrepresentsthe best
compromise as it affords the second highest yield of butene-1, and, most im-
portantly, at a reduced.polyethylenelevel.
This observation has been evidenced by the results of several systematic
studiesof the catalystsTi (OR)&&, [ 27,52,81]. When trimethylaluminum
was used as a co-catalyst, poor catalytic activity was observed which was as-
cribed to its existence as dimer in non-polar solvents. It is well documented
that chlorinated aluminumalkyls decreasethe activity and selectivityof the
catalyst [821. A similarprofile has been observedwith the aluminumsesquich-
lorideand diethylaluminumethoxide [ 831. On the other hand,in a long-stand-
ing research study it has been reported that NJV-dialkylaminoalane,
A. W. Al-Sa’doun / AppX Catal. A 105 (1993) l-40 21

H&N ( CHS)2,in combinationwith butyl titanateexhibitedexcellentselectiv-


ity to butene-1 [84].
It has been noticed that the bulkier the R group (in R&l), the slower is the
dimerization rate. For instance, the dimerizationproceeds faster with AU!&
than with Al (iBu),. This profile has been ascribedto the subtle differencesin
Lewis acidity of these aluminumalkyls as well as the steric effect exerted by
the isobutyl group [ 521.

7.7. Concentration of the titanate complex

It is widely accepted that the initial rate of ethylene conversion over


Ti (OBu),-AlEt3 is independenton the catalystconcentration,indicatingthat
the reaction is not diffusion controlled. At highercatalystconcentration,how-
ever, this no longer holds and the rate appears to be determinedby diffusion
of ethylene from the gas phase into the solution [ 121.
Kinetic studiesof the ethylene dimerizationhave revealedthat the use of a
lower concentration of the catalyst precursoralloweda significantdecreasein
the rate of the secondary reactions leading to the catalyst deactivation
[ 43,52,65]. This is in agreementwith the observationof Belov et al. [851 of a
shortercatalystlife upon increasingthe titanateconcentrationfrom 9010B4to
5010~~mol/l. However, that increase in the titanate concentration led to an
increase in the dimerization rate by a factor of 1.5-5, subject to the stirring
rate of the reaction mixture.
Additionally,the increasein the titanateconcentrationhas an adverseeffect
on the catalyst’s selectivity to butene-1. Increasingthe concentration of the
titanate from 1.5. 10e3 to 9.5. 10e3 mol/l increasesthe proportion of hexenes
from 5.2 to 12.4 wt.-% and the polymer from 0.01 to 0.35 wt.-% [63]. This is
in agreementwith the observationof Pillai et al. [ 521 of a marginalfall in the
selectivityto butene-l when the titanateconcentrationwas doubled This pro-
file could be ascribed to the fact that a higher titanium metal concentration
mightlead to the formation of clustersaroundthe activecenters,therebyhind-
ering the monomer to approach them.

7.8. Nature of solvents

It has been reportedthat the natureof the used reaction medium is critical
for the dimerization reaction [ 671. The desired solvent should be of a non-
destructivenaturetoward the dimerizationcatalyst components [ 831. For in-
stance, due to the presence of a “hard” Lewis acid (AlEt,), “hard” solvents
have a detrimentalinfluenceon the ethylenedimerizationas they tend to neu-
tralize the required Lewis acidity [26]. On the other hand, “soft” solvents
compete with the monomer molecules for coordination onto the metal center.
Along this line, it has been suggestedthat performingthe reaction in solvents
of higher dielectric constant promotes the catalytic.activity and often results
22 A. W. Al-Sa’doun lApp1. Catal. A 105 (1993) l-40

in a more rapid dimerization reaction. However, the draw back associated with
such solvents is the marked increase in the rate of the carbon-carbon double
bond isomerization. As a consequence of that, the ethylene dimerization is
better conducted without any solvent since the isomerized products are highly
undesirable [ 19,861.
In addition to the nature of the utilized solvent, its proportion relative to the
catalyst complex exerts a marked influence on its activity and selectivity
[77,83]. The initial rate of dimerization was found to be independent of the
solvent type, whereas the deactivation rate of the active species varied accord-
ing to the nature of the used solvent [43,87]. In toluene and in the media of
other aromatic solvents, the deactivation proceeds at a somewhat greater rate
than in aliphatic hydrocarbon solvents, such as n-heptane. Along this line, it
has been noticed that the presence of chloride or the usage of chlorinated sol-
vents as the reaction medium turns the dimerization into oligomerization or
polymerization [ 43,781. It has been reported that the deactivation rate of the
active centers in different solvents proceeds in the following order [52]: n-
heptane> C,,-C,, alkane > petroleum ether (40-60°C) > p-xylene > ben-
zene > Ccl,. This sequence is in agreement with the outcome of recent studies
which reported a selectivity to butene-1 in access of 93 wt.-% when the dimer-
ization reaction was carried out in butene-l without using any additional sol-
vent (at Al/Ti = 10) [ 85,881. These findings came in line with an earlier sug-
gestion that the judicious selection of the solvent helps in smoothing the
coordination of the monomer molecules to the metal center and the subsequent
liberation of the dimer from the active center [891.

7.9. Reaction time

It is known that the product spectrum of the ethylene dimerization is depen-


dent on the reaction time. A prolonged reaction time brings about an increase
in the dimer yield at the expense of the products linearity [ 30,191. This is in
close agreement with the observation of Belov et al. [ 871 of a marked increase
in the proportion of isohexenes (2-ethylbutene-1 and 3-methylpentene-1) upon
running the dimerization for a prolonged period of time.
Kinetics studies revealed that the ethylene dimerization proceeds via an in-
duction period, the length of which is largely dependent on the Al/Ti molar
ratio and the reaction temperature [54,65]. By increasing the Al/Ti molar
ratios, the induction period (ti) becomes shorter and the maximum rate of
dimerization (R,) becomes higher. A summarized account for this correla-
tion is presented in Table 11.
On the other hand, the reaction temperature has a pronounced influence on
the length of the induction period. For instance, upon conducting the dimeri-
zation reaction at a lower temperature range ( - 40 to + 2O”C), the induction
period was found to be noticeably long. This situation was reversed at higher
temperature range (50 - 1OO’C) whereupon little or no sign of an induction
period was detected [ 91,921
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 23

TABLE 11

Dependence of the dimerization rate on the reaction time and the AI/Ti ratios [54]

Al/Ti ti Gnu” KU
(molar ratio) (8) (8) k G/g Ti h)

2.7 100 210 466


4.0 80 200 529
5.4 70 190 649
8.1 50 150 740
10.7 50 150 823

“Time to reach the maximum dimerization rate (R,), hexane, 3O”C, 1.1 atm.

In their systematic study of the Ti (OR’ ),-A& Pillai et al. [ 521 noticed a
short induction period (10 min at 26’ C) followed by an increase in the for-
mation of butene-l and in the total ethylene conversion which ultimately
reached a plateau in approximately1 h. This has been consideredas evidence
for the fact that once the active dimerixationspeciesare generatedin sufficient
concentration,the ethyleneuptake remainslinearwith time [52]. Beyond this
period, the authorsnoticed the beginningof a deactivationprocess which sub-
sequentlyresultedin a lower yield of butene-1.

8. THE KINETICS OF THE DIMERIZATION REACTION

The kinetic studiesof the ethylenedime&&ion haveprovided some insight


into the detailed mechanism of the reaction [931. These studies have unani-
mouslyconcluded, assumingthe absence of any side reactions,that the overall
rate constant is first order with each of the concentrationsof the catalystcom-
plex and of the monomer [ 37,52,65,94]. This is in agreementwith a kinetic
scheme in which insertion of ethylene is the rate determiningstep for the di-
merixationreaction [57].
The Z/N catalyzedpolymerizationsare characterizedby a rapidactivityevo-
lution of the catalyst that is usually followed by an immediate decay profile
[ 531. Similarly,the ethylene dimerixationkinetics was found to be of a decay
type, i.e. the rate is maximalat the beginningof the reactionand continuously
decreaseswith time (Fig. 7). The plot on the left, showingthe rapid accumu-
lation of butene-l in the system, is an accelerationtype kinetic curve which
indicates that the dimerixationstarts almost immediatelyafter blending the
catalyst components. This trend is opposed to the plot on the right, which
illustratesthe inversecorrelationbetweenthe dimerizationrate constantsand
the reaction time. This rapid decrease in the dimerixationrate constants is
indicative of the high instability of the active dimerixationspecies under the
specified reaction conditions [ 371.
24 A. W. Al-Sa'doun / Appl. Catal. A 105 (1993) 1-40

-6OO

E
- 500

120 J

-400 ~
E 8

._~

-zoo E

;- 4O!o - I00

20 =10 60 80 100 120


Time, mln.

Fig. 7. Kinetics of ethylenedimerization over Ti (OBu)4-A1Et3,heptane at 70°C, 7.12 atm, A1/


Ti=5.6 (taken from ref. 37).

Interestingly enough, the rate constants for the chain propagation (rp) and
that for the chain transfer reactions (rt~) were both found to be dependent on
the concentrations of the active dimerization sites and the monomer alike,
thus:

[MI (8.1)
[M] (8.2)
where [C*] and [M] are the concentrations of the active titanium sites and
the ethylene monomer, respectively. These results have two interesting fea-
tures. First of all, the fact that both rate constants have the same dependence
on the concentrations of the active sites and that of the ethylene may be inter-
preted as the kinetic confirmation for the suggestion that the chain growth and
the chain transfer are in fact occurring on an alternative and competitive basis.
This could be interpreted as a further evidence for the thought that the dimer
or oligomer transformation of ethylene takes place as a degenerated polymer-
ization according to complex kinetics [24]. Additionally, it implies that the
predominance of any of the different products generated in the Z / N catalyzed
chain-growth reactions is subject to the kinetics conditions, as presented in
Table 12.
Woo and Woo [54] applied eqn. (8.1) to estimate the rate constant of the
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 25

TABLE 12

Chain-growth products end conditions for their formation

Product Kinetics condition

Dimer (C,)
Oligomer (C&C,)
Polymer ( > C,,)

ethylene dimerizationover Ti (OBu),-AlE& The estimatedvalue was found


to be 75.2 1mol-l s-l, measuredat 3O”C, 1.1 atm, and an Al/Ti molar ratio of
5.4.
Analysis of the results of several kinetic studies indicates that the kinetic
profile of the ethylene dimerixation,in both the initial period and the deacti-
vation period, is largely governed by the Al/Ti molar ratio and the reaction
temperature [ 551. The specific impact of these parameterson the kinetics of
the dimerizationcould be summarizedas follows:
I. Theinitial period. In the initial period, the dime&&ion rate is governed
by the concentration of the active sites [C*], which in turn, is controlled by
the Al/Ti molar ratio. Duringthis stage,there is a generalconviction that the
ethylenedimerizationdoes not occur below an Al/Ti molar ratio of 2.0 [ 53,551.
Beyond that ratio ( Al/Ti > 2.0)) the dimerizationproceeds through an initial
and normally short induction period that is subsequentlyfollowed by a rate
maximum.Further increase in the Al/Ti molar ratio resultedin a shorter in-
duction period accompaniedwith an enhanceddimerizationrate. Christenson
et al. [ 511 proposed the following patterns for the formation of the pre-active
sites: !

A
Aid- Al,,, (8.3)

Al, +Ti(OBu), St C, (8.4)


whereAid,Al, are the dimeric and the monomeric forms of AlE& respectively.
The pre-active sites (C,) can be transformed to the active sites (C*) by
further reaction of the former with an excess of the monomeric trialkyl alu-
minum in the following manner;
Ki
C, +Al,- C* (8.5)
whereKiis the rate constant for the formation of the active sites at the initial
period.
Woo and Woo [ 541 reportedthat the Kivaluesat variousAl/Ti molar ratios
in the initial period of the reaction were found to be in the range of 14.2-
15.8.103s-’ .
26 A. W. Al-Sa’doun / Appl. Cabal. A 105 (1993) l-40

On the other hand, the deactivation rate for the active dimerization sites
was found to be governedby the Al/Ti molar ratio, changingfrom first order
at Al/Ti < 5.0 to second order at Al/Ti molar ratio> 5.0. [ 37,541. This could
be interpretedas an evidencefor the existenceof the followingdifferent deac-
tivation schemesat different Al/Ti ratios:

C* 4 Cd.l Al/Ti<5.0 (8.6)

2C* 3 Cdq2 Al/Ti> 5.0 (8.7)


whereC& and C& are the deactivatedspecies,the structureof eitherof which
have not yet been unequivocallyestablished.
It is noteworthy that the deactivation of the active sites C* is caused by
either a substantial increase in the Al/Ti ratio or an equivalent rise in the
reaction temperature.
II. The cieactiuatinperial. In additionto the Al/Ti molar ratio, the reaction
temperaturegovernsthe deactivationrate of the active sites. The highertem-
peraturewill almost certainly lead to a considerableaccelerationin the deac-
tivation rate. This profile has been confirmed by the resultsreportedby Pillai
et al. [52] which furnishedevidence for the decrease in the selectivityto bu-
tene-1 upon increasingthe reaction temperature.A similar decay profile was
notedby Woo and Woo [ 541.This profile is presentedin Table 13,whichclearly
shows a steady increasein the deactivationrate constants (&), at higher re-
action temperature.
A comparable correlationpattern has been observed between the deactiva-
tion rate constants (I&) and the Al/Ti molar ratio [ 381. Under the reaction
conditions, the rate of catalyst deactivationis describedby the first-orderlaw
equation:
C*=C, exp( -kdt) (8.8)
where C, is the initial concentrationof the active dimerizationspecies.

TABLE 13

Dependence of the deactivation rate of the active dimerization species on the reaction tempera-
tures [54]
Al/Ti=9.2, [Ti] ~1.33 mmol/l

Temperature Ethylene pressure Monomer cont. &


(“C) btm) (mol G/l) Wnol Cz)

30 0.918 0.1067 3.79


40 0.798 0.0838 7.23
50 0.636 0.0607 19.62
60 0.419 0.0366 58.83
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 27

TABLE 14

Correlation between the deactivation rate constant ( Kd) and the Al/Ti molar ratios [ 541
T=33”C, P=l.l atm, C+O.1067 mol/l

Al/Ti & &CM


(mol ratio ) (l/m01 C2) (-)

5.4 1.156 0.1257


8.1 3.165 0.3377
9.2 3.798 0.4052
10.7 3.644 0.4729
13.4 5.583 0.5956
16.1 7.304 0.7794

With an increase in the Al/Ti molar ratio, the kinetic order of the catalyst
deactivation changes from first order (eqn. 8.8) to second order (eqn. 8.9) :
k,t= 1/c*- l/C, (8.9)
This was related to the fact that the catalyst mixture generates several active
species, with proportions depending on the Al/Ti molar ratios, some of which
decompose in monomolecular reactions whilst the rest do so in bimolecular
reactions [391. In this regard, it was found that the Kd increases with increas-
ing the Al/Ti molar ratios, indicating a rapid deactivation process, which is
most probably due to the presence of free AlEt at these high Al/Ti molar
ratios. The direct correlation between Al/Ti ratios and the Kd is presented in
Table 14.

9. MECHANISTIC CONSIDERATIONS

It is believed that the proposed scheme for the catalyzed ethylene dimeri-
zation is of the same type as for the “Z/N catalyzed polymerization” (coordi-
nated anionic mechanism) [ 271. Among several mechanistic postulations for
the ethylene dimerization, the most generally accepted are:

9.1. The bimetallic intermediate mechanism

This classical mechanism was originally proposed by Cossee [95,96] who


suggested an alkoxy bridged Ti-Al model as the active dimerization complex.
This model was later evidenced by quantum-mechanical studies. The self-con-
sisted all valence electrons molecular orbital calculations revealed that the
coordination of ethylene to the vacant site imparts a rearrangement of the
catalyst complex from its original trigonal-bipyramidal structure (8) into a
hexa-coordinated octahedral structure (9) with a titanium-ethyl bond at an
intermediate position between the two octahedral sites as depicted in Scheme
28 A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40

1 [ 97,981.This rearrangement is energetically favorable as the energy differ-


ence between these two structures is relatively small (1.43eV) [Ed].
In the light of these calculations, it has been suggested that the driving force
behind the ethylene dimerization appears to be a consequence of the titanium
a-carbon atom activation which in turns, arises from the stabilization of the
titanium’s d-orbital of the octahedral complex as a consequence of the ethylene
coordination [ 571. ESR studies of the reaction pathways revealed that both
the chain propagation and the p-hydrogen transfer, are favored decisively by

Scheme 1. A schematic representation for the bimetallic mechanism.


A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 29

the titanium d-orbital. This implies that the chain growth can occur without
any radicalbreakingand provide a sensiblejustification for the low activation
energyof the dimerixationprocess [ 58 1. It is noteworthythat the description
of the activation process is equally valid for the chain propagation and the
chain terminationalike.

9.2. The cyclic intermediate mechanism

The initial step in this reaction mechanismis the coordination of two mon-
omer molecules to the active metal center leading to the formation of the x-
bonded bis (ethylene) titanium complex ( 13). This step is succeededby the
formation of a metallocyclopentaneintermediate (14) which is subsequently
converted tothe a-bonded mono(butene-1) complex (16)via a 1,3-hydrogen
shift. Dissociation of the latter intermediateproduces the butene-1 as illus-
trated in Scheme 2 [ 9,100,101]:
The main step in this mechanism is a concerted coupling of two ethylene
moleculeson the titaniumatom affordinga titanium(IV) cyclopentanespecies
(Scheme 3). The concerted coupling process is thought to proceed by a step-

Scheme 2. Schematic representation of the cyclic intermediate mechanism


(ligands
aredeleted
for clarity).
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) 1

Scheme 3. Schematic representation for the concerted coupling of ethylene onto the Tiw metal
center [ 991.

wise addition of monomers to the metal center followed by the formation of C-


C bonds. It is important to emphasize that the catalysts which promote this
route should have no isomerizing activity and exhibit high selectivity to the
dimeric product [ 1021.
The cyclic intermediate reaction scheme is marked by a relatively slow for-
mation of butene-l for reasons related to the fact that a coplanar Ti-C-C-H
arrangement is difficult to achieve [ 1031. But on the other hand, the cyclic
character of the titanocyclopentane intermediate species ( 14) offers a reason-
able explanation for the demonstrated selectivity of the titanium-based cata-
lysts to butene-1. Moreover, the absence of the ionic hydrogen species (H+ or
H- ) in the catalytic intermediates ensures the highest possible selectivity to
butene-l and the apparent absence of an isomerization reaction to butene-2
[ 100,101].

10. THE HETEROGENEOUS TITANATE SYSTEMS

The heterogenized catalysts Ti (OR),-AIR, have been utilized either for the
selective production of butene-1 or as a dimerization component in a dual func-
tional catalyst for the in-situ dimerization and copolymerization of ethylene
with the produced butene- 1.

10.1. The supported titanate catalysts

It has been reported that the activity of the homogeneous titanate catalysts
is short lived for reasons related to the reaction of the active species with each
other [ 32,104]. One way to overcome this problem is by supporting these cat-
alysts onto a solid carrier, so as to enhance their overall productivity. With this
in sight, McDaniel and co-workers [32,105] anchored the titanate complex
Ti(OEt)l onto porous carriers such as silica, alumina, and aluminum phos-
phate. Beside their retained dimerization activity, the authors noticed that the
generated heterogenized catalysts have a constant kinetic profile instead of
the rapid evolution and decay of activity, which is a characteristic feature of
A. W. Al-Sa’doun / Appt. Catal. A 105 (1993) l-40 31

their homogenouscounterparts (Fig. 8). Further,a higherand steady dimeri-


zation rate has been achieved, indicating the generationof stabilized dimeri-
zation active sites.
The proportion of the adsorbedtitaniumhas been found to be dependenton
the supporting approach utilized, with the highest loading (3-4 wt.-% Ti)
achieved upon treating the calcined support with a solution of Ti(OEt), fol-
lowed by alkylatingthe resulted slurry with an appropriatealkylatingagent.
This was followed by the introduction of the hexane-rinsedsolid catalyst into
the reactor which contains a solution of AlEt, [32,105]. Emphasis has been
placed on the necessityof havingAll& inside the reactorbefore startup of the
dimerizationreaction, irrespectiveof the type and/or the quantity of the al-
kylatingagent used in the catalyst preparationstage.
In addition, it has been observed that subjectingthe supportsto highercal-
cining temperaturesresults in a sharp increase in the overall activity of the
supportedcatalysts,which was ascribedto the fact that the calciningtemper-
ature affects the density and spacing of the support’s hydroxyl groups [ 321.
Upon calciningthe support at lower temperatures,the closely spaced hydrox-
yls can react in pairs, whereas high calcining temperatures ( > 600’ C ) leave
only singlehydroxyls,which are widely spaced and can thereforereact in a 1: 1
ratio. This is important, in view of the reports which regardedthe monoad-
sorbed species as the only active dimerizationspecies [ 105,106]. The correla-

0
0 5 10 15 20 25 30 35

Reaction Time (min)

Fig. 8. Comparative kinetics of ethylene dimerization over the homogeneous and the heteroge-
neous titanate catalysts (taken from ref. 32 ) .
32 A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40

tion betweenthe catalystactivityand the calciningtemperatureof the support


in the heterogenizedtitanate systems is presentedin Table 15.
The activity of the supportedcatalystshas demonstrateda profound depen-
dence on the proportion of the triethylaluminumco-catalyst. For instance,
addition of l-2 ml of a 15 wt.-% hexane solution of AlEt,, at 80°C and 37 atm
ethylenepressure,to the AlPOd-supportedtitanatecatalystincreasesits activ-
ity from 12 350 to 12 500 g ethylene/g Ti h, with the highestactivity achieved
upon reducing the amount of AlEt3 inside the reactor. Further, the overall
activity of the supportedcatalysts has been doubled to 24 700 g ethylene/g Ti
h upon introducinghydrogenas a co-reactant with ethylene [ 321.
The authorsreportedthat, beside their displayedoutstandingoverall activ-
ity, the supported titanate catalysts have a selectivity to butene-1 of 98.9
wt.-%; (at 95”C, 38.4 atm Pc,, and 3.0 Al/Ti) [32,10/S].Likewisein the non-
supportedsystems,a small amount of polymer was generated (approximately
l-5wt.-%)at the end of the reaction run. However,much of it was recovered
as tiny beads on the catalyst surfaceinstead of being deposited on the reactor
wall as observed with the unsupported catalyst runs [ 321. This is desirable
from an industrialperspective as these catalysts seem to have good potential
with respect to minimizingthe reactor fouling associatedwith the homogene-
ous catalysts.

10.2. The titan&e catalyst in dual functionalsystem

The conventionalTi (OR’ ) *-AIR3catalystshave been used as a component


in dual functionalcatalysts,which are composed of two compatibleactive cen-

TABLE 15

Comparative data on dimerization activity of the homogeneous and the hetereogeneous titanate
catalysts [ 321

Catalyst” support Ti adsorbed Activity


(wt.-%) (g/g Ti h)

Ti(OEt), 3800
solution
Supported 2OO”C-Si02 4.2 315
Ti(OEt),
6OO”C-Si02 3.7 10 800
2OO”C-AlPO, (Al/P=0.9) 2.7 13 800
6OO”C-AlPO, (Al/P =0.9) 2.7 16 900
500”C-AltO3 3.4 4700
800”C-A&O3 2.4 14 300

“Each sample was first reduced by 3-5% MgRz then tested at 85”C, 34 atm with Al/Ti molar
ratio=2-3.
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 33

ters, one for the dimerizationand the other for the subsequentpolymerization
[ 107-1101. The prerequisitefor the dual functional catalysts could be sum-
marizedas: (i) they should not interferechemicallywith each other, and (ii)
they should operate under the same reaction conditions [38]. However, it is
generallyobserved that there exists a significant mutual influence of the two
buildingblocks of dual functional catalystson each other. The polymerization
component can affect its dimerizationcounterparteither by partiallypoison-
ing it or by activatingit. On the other hand, the dimerizationcomponent can
influence its polymerizationcounterpartby selectivelypoisoning some of its
active centers [ 381.
The dual functional catalyst approach has been used by Beach and Kissin
[37,38] for the simultaneousethylene dimerizationto butene-1 and copoly-
merizationof the in-situ formed butene-1 with ethyleneto form branchedpol-
yethylene. This catalyst allows the preparation of LLDPE from ethylene in
one reactor using only a single monomer, ethylene.The dimerizationcompo-
nent of the dual functional catalyst is the conventional alkyl titanate/alumi-
num alkyl dimerizationcatalyst, whereasthe polymerizationcomponent is a
titaniumhalogenidepolymerizationcatalyst.Table 16 contains data generated
from using the dual functional Ti ( 0-iC3H,)4-Al ( C2H5)3-TiC14/MgC12/PE
catalyst [ 381.
The generateddata, listed in Table 16, indicate that the polymerizationac-
tivity of the dual functional catalyst depends primarilyon: (i) the concentra-
tion of butene-1 in the reaction medium (as determinedfrom PE branching)
and, (ii) the concentration of the dime&&ion component [ 381.
Within this context, Sivaram et al. [ 1111 prepared a titanium-magnesium
catalystby mixinganhydrousmagnesiumchloridewith the butyl titanate.The
resultantcomplex has been found to be an effective dual functional catalyst,

TABLE 16

Influence of the dimerization catalyst on the performance of the dual functional catalyst [ 3 ]
Reaction conditions: T=90” C, total reaction pressure = 7.1 atm, concentration of polymerization
catalyst = 0.08 mmol Ti 1-l

Ti(OiPr), Al/Ti Pre-run time Reaction rate PE branching


(mmol/l) (mol/mol ) (min) (g/g Ti atm h ) (Me/1000 C)

0.134 11.6 60 1030 3.0


0.168 11.1 30 760 7.5
0.202 9.3 30 660 18.0
0.269 7.0 30 500 15.1
0.336 5.6 15 560 22.5
0.403 4.6 15 500 24.0
0.470 5.3 15 440 35.5
0.670 4.6 18 460 45.0
34 A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40

with performance ranging from selective ethylene dimerization to total ethyl-


ene polymerization depending on the Mg/Ti ratios. At a Mg/Ti molar ratio of
4, the resulting catalyst mixture is a homogenous solution, whereas higher ra-
tios (Mg/Ti = 10) give rise to a solid catalyst mixture. The latter was found to
be primarily effective in promoting the ethylene polymerization. Appropriate
blends of these two Mg-Ti systems enabled the concurrent dimerization and
copolymerization of ethylene. At lower ratios, the dissolution of MgClz is fa-
vored by the complexation of the chloride ligand with the vacant coordination
site on the titanium atom, affording the stoichiometric complex MgClz-
[ TNB ] 4, [where TNB is Ti (n-OBu),] . This catalyst demonstrated a selectiv-
ity for butene-1 of 95 wt.-% (at Al/Ti = 100, Pc, = 1 atm, 25’ C ) without giving
rise to polymeric by-products. Comparative data of the dimerization reaction
over this system vs. the conventional Ti( OBU)~-A~E~~is presented in Table
17 [3,111].
These data illustrate that the Mg-treated titanate systems are characterized
by a relatively high initial dimerization rate encountered by a low overall eth-
ylene conversion. Furthermore, the Mg-Ti system demonstrated an initial de-
crease in the selectivity upon the transition from Al/Ti 5 to 50, but that profile
was followed by a sharp increase in the catalyst’s selectivity to butene-1 at Al/
Ti ratio of 100 [3,111]
Several other relevant embodiments have been reported in the literature
wherein the Ti (OBu),-Al ( iBu3) catalyst was utilized either in solution or im-
pregnated onto an inert amorphous carriers (such as SiOz, A1203, AlPOd)
[ 112,113]. The reported selectivity of the generated catalysts was found to be
in excess of 85 wt.-% [ 1081. The evaluation studies have been conducted in a
gas-phase fluidized bed process in conjunction with a TiCl,-MgCl, (x=3,4)

TABLE 17

Comparative results of ethylene dimerization using Mg-Ti catalyst and Ti (OBu),-AlEt, catalyst
131
Reaction conditions: T = 27 ’ C, Pcz = 1 atm, t = 45 min, solvent = C11-CII abanes

Mg/Ti Al/Ti Conversion Initial rate Selectivity %


ratio ratio (%I (mol Cz min-’ atm-’ )
Butene-1 Hexenes Octenes

5 68 6.7 89 8 3
10 66 5.2 79 13 8
50 45 1.9 44 44 12
100 20 1.7 95 3 2
5 96 3.9 79 17 4
10 88 1.5 79 20 1
50 41 2.0 83 25 2
100 39 0.9 79 12 9
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 35

catalyst for the in-situ dimerizationof ethyleneinto butene-l and copolymer-


ization of ethylenewith the formed dimericproduct. It has also been reported
that the two reactions are independentof each other and proceed side by side
in the presence of the two different catalysts [ 112,113].

11. OUTLOOK

It is interestingto note that the various reports which regardedthe dimeri-


zation of ethylene as the first phase of an oligomerizationor polymerization
reaction,havealwaysfallen short of furnishinga sensiblereasoningbehind the
termination of the chain growth and subsequentlywere not able to pinpoint
the conditions under which this chain growthbecomes unstable.When assess-
ing the situation of today one can clearly observe a number of problem areas
associated with the dimerizationcatalysts which have to be solved in the fu-
ture, such as:
(i) the identification of the conditions under which ethylene and hydrogen
moleculesbecome activatedby the metal center,
(ii) the correlation between the modifier’s structural properties and their
polymer inhibition characteristics,and
(iii) the presence of multiple active species as detected by gas chromato-
graphy-mass spectrometry.
Needless to say that one has to aim for just one active species if the sole pro-
duction of butene-l is desired. Along this line, side-reactions are to be con-
trolled to a very large extent. In order to put the future trends in better per-
spective,it mightbe of helpto considerthe essenceof the achieveddevelopment
in this researchdomain. The major advancesin the ethylenedimerizationdo-
main have been made in the area of the titaniumbased catalytic systems.Fur-
ther advances in understandingthe nature of the catalytic species could lead
to the development of a catalyst capable of selectivelydimerizingethyleneto
butene-1 without forming any by-products. There is still a long way ahead in
the development of such ideal catalysts. Homogeneous catalysts that have
reachedthe stage of commercializationwill probably allow most of the above.
Undoubtedlyit will take extensiveefforts to arriveat the above ideal situation
but in particularthe fact that homogeneous catalysts lend themselvesmuch
more readilyto study, modification and “fine-tuning” than do their heteroge-
neous counterparts,presents an enormous potential for steeringthe catalyst
performance and thereby tailoring the product properties. Nevertheless,the
idea of heterogenizingthe homogeneoustitanate catalysts have added a new
dimension to the alreadybustling field of researchin the alkene dimerization
domain. But, in comparison with the vast number of scientific studies on the
homogeneous titanate systems, only few efforts have been directed toward
studying the heterogeneouscatalysts, probably due to their structuralcom-
plexity and the possible pluggingof the active centersby the formed polymeric
36 A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40

materials.This complexitywill undoubtedlybe the focal point that will attract


in the future more research efforts in this domain. However, the future re-
searchtrends in these systemswill predominantlybe determinedby the mar-
ket demand for b&me-copolymer in comparison with the currentlyfavorite
hexene-copolymer.

12. SUMMARY

The analysisof the developmentsmade in the field of the catalytic ethylene


dimerizationto butene-1 singled out the titanium alkoxides in combination
with triethyl aluminumas the most effective and selective catalysts for that
purpose. Amongst these systems,the Ti ( OR)4-AlR~ (R = n-Bu or iPr ) have
demonstratedthe highest activity and selectivity at moderate pressure and
temperature.
It has been shown that the reduction of the titanate complex has a pro-
nounced impact on the course of the dimerixationreaction. Several studies
haveconcludedthat triethylaluminummustbe part of the catalyst’sactivesite.
This is so because,whilst other metal alkyls can alkylatethe titanatecomplex,
the activity of the system develops only when triethylaluminumis present.
Thus, the amount of triethylaluminumneeded to achieve optimum activity,
and the activity profile itself, depends somewhat on the particular catalyst
preparation procedure. It is interestingto note that the proposed active di-
merixationsites contain at least one Ti-C bond which is regardedto be an
essentialfeaturefor the occurrenceof the chain-growthreactions.
The ethylenedimerixationreaction can be influencedby a multitudeof cat-
alyst compositions and reaction parameters.It has been shown that butene-1
can be produced with excellent selectivity and higher conversion rats, under
mild reaction conditions, through the proper selection of an optimum Al/Ti
molar ratio. This ratio could be varied,within a narrowrangeabove and below
which there are adverseeffects on both the rate of the reaction and the selec-
tivity to butene-1.

ACKNOWLEDGEMENTS

I am indebtedto the Directorateof Researchand Developmentin the Saudi


Basic IndustriesCorporation (SABIC ) for grantingme the permissionto pub-
lish this work. Specialthanks are extendedto the refereeswho made very val-
uable comments and suggestionsabout the contents of this paper.

REFERENCES

1 Butene-1 Chemical Profile, Chemical Marketing Reporter, (11 February 1991) 239.
A, W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 37

2 G. Lappin, in G. Lappin and J. Sauer (Editors), Alpha-Olefins Applications Handbook,


Chapter 1, Marcel Dekker, New York, 1989, p. 3.
3 A.M. Ai Jaraliah, J.A. Anabtawi, M.A.B. Siddiqui, A.M. Aittani and A.W. AI Sa’doun,
CataI. Today, 14 (1992) 1.
4 A. Behr and W. Keim, Arab. J. Sci. Eng., 10 (1985) 377.
5 Y. Shiraki and T. Tamura, European Patent 328 728 (1989).
6 P.W. Jolly and G. Wiike, in P.M. Maiths F.G.A. Stone and R. West (Editors), Organic
Chemistry of Nickel, (Organometaihc Chemistry Vol. 2)) Academic Press, New York, 1975,
pp. l-33.
7 A. Hennico, J. Leonard, A. Foreatiere, and Y. Giaize, Hydrocarb. Process., 69 (1990) 73.
8 Petrochemical Handbook ‘91, Hydrocarb. Process., 70 (1991) 144.
9 D. Commereuc, Y. Chauvin and A. Bre, New J. Chem., 10 (1986) 34.
10 G.P. Belov, V.I. Smirnov, T.I. Solov’eva and F.S. D’yachkovskiy, DepositedDoc., VINITI,
p. 1478,1977; Chem. Abs. 90,86645 (1979).
11 B. Schieppinghoff, J. Herwig and H.V. Scheef, German Patent 3 301162 (1983); Chem.
Abstr. 101,152502 (1984).
12 C.E.H. Bawn and R. Symcox, J. Polym. Sci., 34 (1959) 139.
13 M. Takeda, K. Kimura, Y. Nozawa, N. Koide and M. Hisatome, J. Polym. Sci. Part C, 23
(1968) 741.
14 G. Natta, J. Polym. Sci., 48 (1960) 419.
15 G. Natta, L. Porri, and A. Carbonaro, Makromol. Chem., 77 (1964) 126.
16 G. Natta, L. Porri and S. Vaienti, Makromol. Chem., 67 (1963) 225.
17 H. Bestian and K. Ciauss, Angew. Chem., Int. Ed. Engl., 2 (1963) 704.
18 H. Bestian, K. Ciauss, H. Jensen and E. Prinz, Angew. Chem., Int. Ed. Engl., 2 (1963) 32.
19 A.W. Langer, J. Macromol. Sci., Chem. A, 4 (1970) 775.
20 G. Olive-Hemici and S. Olivd, Chem. In. Tech., 43 (1971) 906.
21 G. Olive-Hen&i and S. OlivB, Adv. Polym. Sci, 15 (1974) 1.
22 H. Martin, Angew. Chem., 68 (1956) 438.
23 S.M. Pi&i, M. Ravindranathan and S. Sivaram, Chem. Rev., 86 (1986) 353.
24 P.A. Gaitier, A.A. Forestiere, Y.H. Glaize and J.P. Wauquier, Chem. Eng. Sci., 43 (1988)
1855.
25 W. Keim, A. Beher and M. Roper in G. Wilkinson, F.G.A. Stone and E.W. Abel (Editors),
Comprehensive Organometalhc Chemistry, Vol. 8, Pergamon Press, New York, 1982, p. 371.
26 K. Ziegler, Brit Pat. 787 438 (1957), Chem. Abstr. (1958) 521907.
27 G. Lefebvre and Y. Chauvin, in R. Ugo (Editor), Aspecte of Homogeneous Catalysis, Vol.
1, Carlo Manferdi, Milan, 1970, p. 107.
28 A. Krzywicki and K. Jhonstone, in K.J. Smith and E.C. Sanford (Editors), Progress in
Catalysis (Studies in Surface Science and Catalysis, Vol. 73), Elsevier, Amesterdam, 1992,
p. 155.
29 L.G. Wideman, H.R. Menapace and N.A. Maly, J. Catal., 43 (1976) 371.
30 J. Skupinska, Chem. Rev., 91 (1991) 613.
31 C. Elschenbroich and A. Saizer, Organometailics: A Concise Introduction, VCH, Weinheim,
1989, p. 411.
32 P.D. Smith, D.D. Klendworth and M.P. McDaniel, J. Catai., 105 (1987) 187.
33 G. Hen&i-Olive and S. Olive, in J.C.W. Chien (Editor), Coordination Polymerization,
Academic Press, New York, 1975, p. 291.
34 P.E. Matkoviskii, F.S. D’yachkovskiy, L.N. Rusiyan, V.D. Litvinova, V.N. Belova, G.P.
Strataeva and K.M.A. Brikenehtein, Neftekhimiya, 22 (1982) 317.
35 S. Yamada and I. Ono, Bull Jpn. Pet. Inst., 12 (1970) 160.
36 M. Farina and M. Ragazzini, Chim. Ind., 40 (1958) 816.
37 D.L. Beach and Y.V. Kissin, J. Polym. Sci., Polym. Chem. Ed., 22 (1984) 3027.
38 A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40

38 D.L. Beach and Y.V. Kissin, J. Polym. Sci. Polym. Chem. Ed., 24 (1986) 1069.
39 K. Yamada and Y. Kondo, Japanese Patent 62,61932 (1987); Chem. Abstr. 107, 96315
(1987).
40 V.I. Zhukov, N.P. Shestak, G.P. Belov, M.N. Dyadyuova, L.A. Shilov, L.D. Shevlyakov,
F.S. D’yachkovskiy, A.G. Liakumovich and P.A. Vernov, US Patent 4 101600 (1978).
41 G.P. Belov, T.F. Dzhabiev, V.E. Kuzmin, F.S. D’yachkovskiy, M.P. Gerasina, K.A. Briken-
shtein, Soviet Patent 485 102 (1975); Chem. Abstr. 84,44933 (1976).
42 Japanese Patent 57 169 430 (1982), to Mitaui Toateu Chemicals Inc.; Chem. Abstr. 98,
88801 (1983).
43 G.P. Belov, T.F. Dzhabiev and T.F. D’yachkovskiy, in F. Marta and D. KalIo (Editors),
Proc. Symposium on the Mechanism of Hydrocarbon Reactions, Hungary, 1973, Elsevier,
Amsterdam, 1975, p. 507.
44 H. Hirai, K. Hiraki, I. Noguchi and S. Makishima, J. Polym. Sci: A-l, 8 (1970) 147.
45 H. Hirai, K. Hiraki, I. Noguchi, T. Inoue and S. Makiahima, J. Polym. Sci; Part A-l, 8
(1970) 2393
46 T.S. Dzhabiev, R.D. Sabrova and A.E. Shilov, Kinet. KataI., 5 (1964) 441.
47 M. Utesuki and Y. Fujiwara, Bull. Chem. Sot. Jpn., 49 (1976) 3530.
48 E. Angelescu, C. NicoIau and Z. Simon, J. Am. Chem. Sot., 88 (1966) 3910.
49 G.P. Belov, T.S. Dzhabiev and I. Kolesnikov, J. Mol. Catal., 14 (1982 ) 105.
50 T.S. Dzhabiev, F.S. D’yachkovskiy and N.D. Karpova, Kinet. Katai., 15 (1974) 67.
51 C.P. Christenson, J.A. May and L.E. Freyer, in R.P. Quirk (Editor), Transition Metal
Catalyzed Polymerization: Aikenes and Dienes, Harwood Academic, New York, 1983, p.
763.
52 S.M. PilIai, G.L. Tembe, M. Ravindranathan and S. Sivaram, Ind. Eng. Chem. Res., 27
(1988) 1971.
53 Y.V. Kiesin, Isospecific Polymerization of Olefins with Heterogeneous Ziegler-Natta Cat-
alysts, Springer Verlag, New York, 1985, Chap. 5.
54 T.W. Woo and S.I. Woo, J. Catal., 132 (1991) 68.
55 D.C. BradIy in F.G.A. Stone and W.A. Graham (Editors), Inorganic Polymer, Academic
Press, New York, 1962, Chap. 7.
56 T.F. Dzhabiev, Z.M. Dzhabiev, G.P. Belov and F.S. D’yachkovskiy, Neftekhimiya, 16 (1976)
706.
57 L.A. Rodrigez, H.M. Van Looy and J.A. Gabant, J. Polym. Sci. Part A-l, 4 (1966) 1905.
58 C. Masters, Homogeneous Transition Metal Catalysis - A Gentle Art, Chapman and Hall,
New York, 1981, p. 135.
59 J. Vybihal, Chem. Prum., 31 (1981) 64; Chem. Abstr. 96 (1982), 69463.
60 P. Natta, J. Polym. Sci., 34 (1959) 151.
61 G.P. Belov, T.S. Dzhabiev, F.S. D’yachkovekiy, V.I. Smirnov, N.D. Karpova, K.A. Briken-
stein, M.P. Gerasina, V.E. Kuzmin and P.E. Matkovskii, US Patent 3 879 485 (1975).
62 G.P. Belov and V.I. Smirnov, Neftekhimiya, 17 (1977) 3.
63 G.P. Belov, F.S. D’yachkovskiy and V.I. Smirnov, Neftekhimiya, 18 (1978) 223.
64 N. LeQuan, D. Cruypelinck, D. Commereuc, Y. Chauvin and G. Leger, US Patent 4 532 370
(1985).
65 G. Hen&i-OlivB and S. Olivd, Polym. Lett. Ed., 12 (1974) 39.
66 A.W. Langer, Am. Chem. Sot., Div. Petrol. Chem., Preprints, 17 (1972) B119.
67 G.P. Belov, T.F. Dzhabiev and F.S. D’yachkovskiy, in A.E. Shilov (Editor), Fundamental
Research in Homogenous CataIyeis, Vol. 3, Gordon and Beach, 1986, p. 1015.
68 E. Angelescu, A. Angelescu and L.V. Nicolescu, Rev. Roum. Chim., 30 (1979) 523.
69 N. LeQuan, D. Cruypelinck, D. Commereuc, Y. Chauvin and G. Leger, US Patent 4 615 998
(1986).
70 S.M. PiI&, M. PiIIai, S. Sivaram and M. Ravindranathan, European Patent 221206 (1987 ) .
A. W. Al-Sa’doun / Appl. Catal. A 105 (1993) l-40 39

71 N. LeQuan, D. Cruypelinck, D. Commereuc and Y. Chauvin, European Patent 135 441


(1985).
72 V.E. Nevzorov andG.P. Belov, Neftekhimiya, 29 (1989) 340; Chem Abstr., 112 36489 (1990).
73 S. Yamada, I. Ono, H. Abe, K. Nubuka and T. Kazuo, German Patent 2 026 246 (1971);
Chem. Abstr. 87315 (1971).
74 G.P. Belov,T.F. Dzhabiev,F.S. D’yachkoviskiy,V.I. Smimov, N.D. Karpova, C.A. Briken-
shtein, M.P. Gerasina, V.E. Kuzmin and P.E. Matkovskii, German Patent 2 462 771 (1980),
Chem. Abstr. 94,66412 (1981).
75 I.A. Campbell, UK Patent 1447 812 (1976).
76 G.S. Sergienko, V.I. Zhukov and G.P. Belov, Neftekhimiya, 3 1 ( 1991) 50.
77 P.A. Vemov, N.V. Lemaev, A.G. Liakumovich, G.P. Belov, E.V. GutaaIyuk and V.P. Ki-
chigin, Neftekhimiya, 8 (1977) 25; Chem. Abstr. 87 167480 (1977).
78 G. Hen&i-Olive and S. Olivd, Chem Tech., (1981) 746.
79 G. Henrici-Olive and S. Olive, Angew Chem., Int. Ed. Engl., 10 (1971) 105.
80 G. Hen&i-Olive and S. Olivb, Coordination and Catalysis, Verlag Chemie (VCH), Wein-
heim, 1977, p. 146.
81 G.L. Tembe and M. Ravindranathan, Ind. Eng. Chem. Res., 1991 (30) 2247.
82 T. Yano, S. Ikai and K. Washio, J. Polym. Sci. Chem. Ed., 23 (1985) 3069
83 P.E. Matkovskii, L.N. Russiyan, F.S. D’yachoviskiy, K.A. Brikenshtein and M.P. Gera-
sina, Kinet. Katal., 19 (1978) 263.
84 S. Cesca, W. Marconi and M.L. Santostaei, J. Polym. Sci, Part B, 7 (1969) 547.
85 G.P. Belov, A.A. Brikenshtein, M.P. Gerasina, F.S. D’yachkovskiy, Neftekhimiya, 6 (1989)
14; Chem. Abs. 111,154402 (1989).
86 J. Vybihai, Czech Patent 179 183 (1979); Chem Abstr. 92,77131 (1980).
87 J. Vybihai, Czech Patent 179 726 (1979); Chem. Abstr. 92,129603 (1980).
88 J. Schoengut end K. Jiratova, Czech Patent 189 984 (1982); Chem. Abstr. 99,70179 (1983).
89 K. Kawamoto, T. Imanaka and T. Shiichiro, Nippon Kagaku Zasshi, 91 (1970), 39; Chem.
Abstr. 72 99767 (1972).
90 V.I. Zhukov, G.B. Belov, V.Y. Ivolgin, N.P. Shestak, N.V. Kartasheva and G.Y. Yushina,
Neftekhimiya, 22 (1982) 598.
91 G.P. Belov, T.S. Dzhabiev, F.S. D’yachkovskiy, V.I. Smimov, N.D. Karpova, K.A. Briken-
shtein, M.P. Gem&a, V.E. Kuzmin, P.E. Matkovsky, L.N. Russiyan, A.D. Pomogali and
N.M. Chirkov, US Patent 3 969 429 (1976).
92 G.P. Belov, T.F. Dzhabiev, F.S. D’yachkovskiy, V. Smimov, N. Karpova, K.A. Briken-
shtein, M. Gerasina, V. Kuzmin, P. Matkovsky, L.N. Russiyan, A.D. Pomogaiic and N.M.
Chirkov, US Patent 3 911042 ( 1975).
93 G.P. Belov, B.L. Psikha, G.A. Furman, S.P. Davtyan and A.V. Agasaryan, Neftekhimiya,
30 (1990) 609; Chem. Abstr. 114,61477 (1991).
94 Y.V. Kissin and D.L. Beach, J. Appl. Polym. Sci., 29 (1984) 1171.
95 P. Cossee, J. Catai., 3 (1964) 80.
96 P. Cossee in A.D. Keetley (Editor), The Stereochemistry of Macromolecules, Vol. I, Dek-
ker, New York, 1967, Chap. 3.
97 D.R. Armstrong, P.G. Perkins and J.J.P. Stewart, J. Chem. Sot., Dalton Trans., (1972)
1972.
98 0. Novaro, S. Chow and P. Magnouat, J. Catai., 41 (1976) 91.
99 0. Novaro, S. Chow, and P. Magnouat, J. Catai., 41 (1976) 131.
100 D. Commereuc, Y. Chauvin, J. Gaiiard and J. Leonard, Hydrocarb. Process., 63 (1984) 118.
101 Butene-1 and other LLDPE Comonomers, A Report prepared by Chem. Systems Inc., New
York, 1986, p. 275.
102 A. Carbonaro, A. Gerco and G. Daiiosta, J. Organomet. Chem., 20 (1969) 177.
40 A. W. Al-Sa’doun /Appl. Catal. A 105 (1993) l-40

103 R.H. Crabtree, The Organometallic Chemistry of the Transition Metal, Wiley, New York,
1988, p. 40.
104 M.P. McDaniel, Adv. Catal., 33 (1985) 47.
105 M.P. McDaniel, P.D. Smith and D.D. Klendworth, European Patent 230 983 (1987).
106 R.L. Burwell, J. Catal., 86 (1984) 301.
107 D. Commereuc, F. Hugues, N. Le Quan and A. Tauoli, US Patent 4 721762 (1988).
108 K. Cann, M.W. Chen and F.J. Karol, US Patent 4 861846 (1989).
109 A. Delbouille and H. Toussaint, US Patent 3 526 616 (1970)
110 R. Cooper and K. Whiteley, US Patent 4 133 944 (1979)
111 S. Sivaram, Shashikant and V.K. Gupta, Am. Chem. Sot., Div. Petrol. Chem., Preprints,
24 (1984) 595.
112 K. Cann and F.J. Karol, World Organization Patent 8605 500 (1986); Chem. Abstr. 106,
19175 (1986).
113 G. Davidovici, UK Patent 2 231583 (1990).

Вам также может понравиться