8-PNBD - Applied Polymer Science

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Ring-Opening Metathesis Polymerization of Bicyclo[2.2.

1]hepta-2,5diene (Norbornadiene) Initiated by New Ruthenium(II) Complex


Esra Evrim Yalnkaya,1 Osman Dayan,2 Mehmet Balcan,1 etin Gu
ler1
1
2

Chemistry Department, Faculty of Science, Ege University, Bornova, 35100 Izmir, Turkey
Laboratory of Inorganic Synthesis and Molecular Catalysis, anakkale Onsekiz Mart University, 17020 anakkale, Turkey

Correspondence to: E. E. Yalnkaya (E-mail: [email protected])

The polymerization of norbornadiene (NBD) initiated by a novel ruthenium (Ru)(II) complex (3) containing 1,10 -pyridine-2,6-diylbis[3-(dimethylamino)prop-2-en-1-one] (1) as ligand has been investigated. Ru complexes exhibit more catalytic activity
in the ring-opening metathesis polymerization (ROMP) of NBD when activated with trimethylsilyldiazomethane (TMSD). The influence of the various experimental parameters such as reaction time and temperature, nature of the solvent and catalyst, ratio of the
NBD/Ru, and TMSD addition has been investigated. The polymers have been obtained in high yields with a relatively low polydisper n and M
 w values in a monomodal distribution. Their structures have been determined by means
sity index for ROMP and a high M
of FTIR and 1H-NMR spectroscopy. Thermal properties have been determined via thermogravimetric analysis and DTG methods.
The NBD polymerization results that initiated by Ru-based catalyst coordinated to amine ligand have been compared to initiated by
C 2012 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 000: 000000, 2012
[RuCl2(p-cymene)]2. V
ABSTRACT:

KEYWORDS: norbornadiene; ring-opening metathesis polymerization; ruthenium-based catalyst; characterization

Received 3 September 2010; accepted 12 April 2012; published online


DOI: 10.1002/app.37867
INTRODUCTION

Norbornadiene (NBD) can serve as monomer for the synthesis


of high-molecular-mass polymers with widely varying structure.
This cyclic compound gives quite different chemical polymer
structures, and the resulting variation of properties can be
achieved by the selection of various catalyst systems. The cyclolinear structures are formed in the presence of ring-opening metathesis polymerization (ROMP) catalysts (e.g., based on Mo,
W, Re, or Ru compounds) by opening of the strained five-membered-ring NBD, and thus, the polymer can be synthesized.14
Although olefin metathesis was discovered as early as 1955,5 it
has only achieved a leading role in synthetic methodology in
the last decade. This is mainly attributed to advances in the
field of catalysis and organometallic chemistry, which have been
heavily influenced by the work of Grubbs and coworkers68 and
Schrock et al.9 in developing well-defined transition metalcarbene complexes.10,11 The most notorious representative of new
generation of olefin metathesis catalysts is the metalalkylidene
CHPh)(PCy3)2 developed by Grubbs and cocomplex RuCl2(
workers.12,13 This complex, along with similar monometallic
and bimetallic species, has found numerous applications in

ROMP over the past few years.14,15 Delaude et al. found that
the dimeric ruthenium (Ru) catalyst [RuCl2(p-cymene)]2 was
capable of ROMP-functionalized NBDs in the presence of trimethylsilyldiazomethane (TMSD) as the cocatalyst.16 This
research group also showed that readily available Ru complexes
of the type RuCl2(arene)(PR3) are versatile and efficient promoters for the ROMP of both strained and low-strain olefins
when activated by a suitable carbene precursor such as
TMSD.17,18 ROMP of NBD by using Ru complexes of the type
[Ru(g5-C9H7)Cl(1,5-cyclooctadiene)] was reported by Alvarez
et al.19 with a high molecular weight and narrow polydispersity
index (PDI) at room temperature (RT).
The growing interest in the development of new polymers
formed from ROMP has stimulated researchers to search new
catalysts for this reaction. Although it is reasonable to assume
that Grubbs-type complexes are powerful catalysts for olefin
metathesis, new catalysts must be formulated with the design of
new materials,20,21 especially using transition metal complexes
with different ligands.21,22 Thus, the properties of the polymers
depend on the ligands coordinated to the metal center not
directly engaged in reaction. The most commonly used ancillary
ligands are the phosphines and amine ligands, where they tune

Additional Supporting Information may be found in the online version of this article.
C 2012 Wiley Periodicals, Inc.
V

WWW.MATERIALSVIEWS.COM

WILEYONLINELIBRARY.COM/APP

J. APPL. POLYM. SCI. 2012, DOI: 10.1002/APP.37867

ARTICLE

the substitution lability of metal complexes via a systematic variation of their steric and electronic properties.23 Although many
studies of ROMP of NBDs with Mo- and W-containing catalysts2426 and also Grubbs catalysts have been carried out to
date,68 there is still demand for stable and easily handled complexes prepared from cheap starting materials, which make
them preferred catalyst precursors. There has been rarely report
of the Ru(II)amine complexes as catalyst for ROMP of NBD.
Moreover, there are scarcely any studies of NBD synthesis with
the newly synthesized Ru-based catalyst coordinated to amine
ligand by ROMP with reporting 1H-NMR and FTIR spectroscopy, thermal analysis, molecular weight, and investigation of
experimental parameters. This article aims to contribute to development of Ru(II)amine complexes that are able to catalyze
ROMP of the highly strained NBD with high yield. To further
sustain these observations, we have launched a detailed study of
the polymerization of NBD in the presence of new Ru-based
catalytic system with TMSD and compared this new catalyst
with well-known [RuCl2(p-cymene)]2 catalyst to determine the
influence of the ligand on this polymerization. Also, we report
the polymerization of NBD under a variety of experimental
conditions in this article.
EXPERIMENTAL

Materials
All reactions and manipulations were performed under an argon
(Ar) atmosphere by using conventional Schlenk tube techniques.
Ar gas was dried by passing through P2O5 (97%, Aldrich). The
[RuCl2(p-cymene)]2 (2) was prepared according to the method
given in the literature27; the structure and the purity of the
dimer were checked with FTIR and 1H-NMR spectroscopy.
NBD was supplied by Aldrich and distilled from CaH2 and
dried over P2O5 under Ar gas before use. Commercial-grade
solvents were also dried under Ar atmosphere and distilled
before use. Solutions of the catalyst were freshly prepared for
each reaction. All of the chemicals used were reagent grade.
TMSD (Aldrich) was supplied as 2M solution in hexanes. This
solution was further diluted by addition of a suitable solvent.
Characterization Techniques
H-NMR (400 MHz) spectra were recorded on a Varian Unity
400 Spectrometer. FTIR spectra were obtained with a Perkin
Elmer Pyris 1 FTIR Spectrometer. The number- and weight-av n and M
 w ) and PDI (M
 w =M
 n ) of
erage molecular weights (M
the polymers were determined by gel permeation chromatography (GPC). GPC analysis of the polymers was performed on a
Hewlett-Packard HPLC system with a differential refractometer
detector. Tetrahydrofuran (THF) served as the eluent at a flow
rate of 1.0 mL/min. The molecular weights and polydispersities
were reported versus monodisperse polystyrene standards. Thermal degradation was studied with a thermogravimetric analyzer
(Perkin Elmer Pyris 1 TGA/DTA) by the heating of samples
from ambient temperature to 1000 C at 10 C/min under a 10
bar dry N2.
1

Synthesis of 1,10 -Pyridine-2,6-diylbis[3-(dimethylamino)prop2-en-1-one] (1)


This compound was synthesized according to a modified literature procedure.28 2,6-Diacetylpyridine (0.5 g, 3 mmol) was dis-

J. APPL. POLYM. SCI. 2012, DOI: 10.1002/APP.37867

solved in 10 mL of toluene in the Schlenk tube. Then, 3 mL of


N, N-dimethylformamide dimethylacetal was added and stirred
over night at 8090 C in Ar atmosphere. The solvent was filtered by cannula wire, and the residue was dried in vacuo at RT.
Orange color solid, 80% yield: 1H-NMR (CDCl3) d (ppm) 8.12
(t, 1H, J 6.4, pyridine-Hp), 7.917.98 (m, 4H; pyridine-Hm,
HC
N(CH3)2), 6.59 (d, 2H, J 12.4, O
CACH), 3.212.98
(ss, 12H, N(CH3)2). 13C-NMR (CDCl3) d (ppm) 183.5, 154.8,
148.9, 137.5, 123.62, 89.7, 37.3. FTIR (cm1) 3099, 3053, 2911,
1645, 1632, 1584, 1494, 1395, 1261, 1129, 1057, 993, 897, 783.
Synthesis of (3)
[RuCl2(p-cymene)]2 (2) (0.1 g, 0.16 mmol) was dissolved in dry
dichloromethane (10 mL) and subsequently amine ligand (1)
(0.088 g, 0.32 mmol) was added slowly. The reaction mixture
was stirred at RT for a night in Ar atmosphere (Scheme 2). The
solution was concentrated by removing half of the solvent in
vacuo. The final product was precipitated in diethylether, filtered, and dried in vacuo at RT. Brown solid, 90% yield (0.17
g): 1H-NMR (CDCl3) d (ppm) 8.75 (t, 1H, J 8.0, pyridineHp), 8.20 (d, 2H, J 7.8, pyridine-Hm), 7.60 (d, 2H, J 6.2,
CACH), 5.255.45
HC
N(CH3)2), 6.30 (d, 2H, J 6.8, O
(dd, 4H, J1 5.6; J2 6.0, arom.), 3.053.15 (ss, 12H,
N(CH3)2), 2.75 (m, 1H, CH(CH3)2), 2.20 (s, 3H, CH3), 1.15
1.20 (d, 6H, J 7.2, CH(CH3)2). 13C-NMR (CDCl3) d (ppm)
(p-cymene) 103.9, 103.5, 86.7, 83.6, 30.6, 21.1, 18.5; (ligand)
186.7, 160.1, 154.6, 138.1, 127.9, 91.5, 41.3. FTIR (cm1) 3048,
2962, 2865, 1640, 1623, 1523, 1489, 1473, 1394, 1362, 1242,
1142, 1083, 1056, 1010, 976, 949, 897, 784, 714. Anal. calcd. for
RuC25H33Cl2N3O2 (579.52): C, 51.75; H, 5.69; N, 7.24%. Found:
C, 51.05; H, 5.35; N, 6.81%.
Polymerization of NBD
In a typical ROMP experiment, 0.04 mmol (0.023 g) of the catalyst solution in 3 mL of dry CH2Cl2 was transferred into a
Schlenk tube under Ar and stirred for a few minutes. A total of
8 mmol of NBD monomer was injected into the catalyst solution and stirred for a few minutes before 1 mL of 0.1M TMSD
in the reaction solvent (CH2Cl2) (0.1 mmol) was added dropwise over a 30-min period. The reaction mixture was kept at
RT and was also refluxed at 40 C for different time periods
([Ru]/[NBD] 1/200). Then, the polymer was precipitated in
large amount of methanol by adding dropwise and stirring vigorously. The precipitated polymer was filtered and washed with
small portions of methanol several times. It was dried in vacuo
overnight and characterized by FTIR, GPC, and 1H-NMR spectroscopy. 1H-NMR (400 MHz, CDCl3): d 5.65.2 (m, 4H), 3.6
(s, 1H), 3.2 (s,1H), 2.3 (s, 2H), 1.2 (s, 2H). FTIR (cm1) 3380,
2962, 2915, 1681, 1425, 865.
RESULTS AND DISCUSSION

The synthesis of 1,10 -pyridine-2,6-diylbis[3-(dimethylamino)prop-2-en-1-one] (1) ligand is given in Scheme 1. The band
assignment for the FTIR spectrum of ligand was as follows: olefinic CAH stretching band in C
CH and aromatic CAH
stretching band at about 3000 cm1; CAH stretching in CH3
and CAH stretching in CH2 at about 2900 cm1; CAH stretchO diketones carbonyl
ing in NACH3 at about 2800 cm1; C

WILEYONLINELIBRARY.COM/APP

ARTICLE

Scheme 1. Synthetic route of 1,10 -pyridine-2,6-diylbis[3-(dimethylamino)prop-2-en-1-one] (1).

absorption band at 16551635 cm1; C


C in pyridine ring at
16001585 cm1; C
C in chain and C
N in pyridine ring at
16801620 cm1; CAC in pyridine ring at 15001400 cm1;
and CAN in ring at 13351020 cm1. Further bands were
shown by aromatic rings in the fingerprint region between 1225
and 950 cm1.
From the 1H-NMR spectrum of amine ligand, the peak at 8.12
ppm was assigned to CH group proton of pyridine-Hp. The
peaks at 7.917.98 ppm were assigned to four CH groups proN(CH3)2. The d value of two CH
tons of pyridine-Hm and HC
groups protons that O
CACH was shown at 6.59 ppm. The
signals on 3.212.98 ppm were assigned to four CH3 groups of
N(CH3)2.
The Ru(II) complex (3) containing 1,10 -pyridine-2,6-diylbis[3(dimethylamino)prop-2-en-1-one] ligand was synthesized by
starting from [RuCl2(p-cymene)]2 (Scheme 2). From the FTIR
spectrum of the catalyst: olefinic CAH stretching in C
CH and
aromatic CAH stretching at about 3040 cm1; CAH stretching
in CH3 and CAH stretching in CH2 at about 2900 cm1; CAH
C in chain and
stretching in NACH3 at about 2800 cm1; C
O diketones carbonyl
C
N ring at about 1620 cm1; C
C in pyridine ring at
absorption band at about 1640 cm1; C
about 1523 cm1; CAC in pyridine ring at 15001400 cm1;
and CAN in ring at 13351020 cm1. Further bands were
shown by alkanes and alkenes in the fingerprint region between
1400 and 800 cm1.
The structural analysis of 3 was determined on the basis of 1Hand 13C-NMR spectra. From the 1H-NMR spectra, the effect of
coordination of the Ru(II) to the 1,10 -pyridine-2,6-diylbis[3(dimethylamino)prop-2-en-1-one] ligand through the pyridyl
nitrogen for 3 could clearly be seen. The protons of the pyridine
ring of 3 are shifted downfield compared with free ligands (1).
The triplet peak at 8.75 ppm was assigned to CH group proton
of pyridine-Hp. The doublet peak at 8.20 ppm was assigned to
other two CH group protons of pyridine-Hm. The d value of
two CH groups protons of HC
N(CH3)2 chain was shown at
7.60 ppm. The doublet signal on 6.30 ppm was assigned to two

CH groups of O
CACH chain. The double doublet peaks at
5.255.45 ppm were assigned to four CH groups of benzene
ring protons. The peaks at 3.053.15 ppm were assigned to four
CH3 groups of N(CH3)2. By calculating peak areas, the d value
of CH group proton of CH(CH3)2 was assigned at 2.75 ppm,
CH3 group proton of CH(CH3)2 was assigned at 2.20 ppm, and
two CH3 group protons of CH(CH3)2 were represented at 1.15
1.20 ppm. Also, the total count of carbon peaks for 3 matched
well with the composition of the complex in 13C-NMR spectra.
When the ligand and complex carbon peaks were compared, the
d values were shifted to high values in the 13C-NMR spectra
especially at carbonyl and pyridine carbons.
The polymerization of NBD in dichloromethane at RT in the
presence of Ru(II) complex (3) and TMSD catalytic system was
achieved (Scheme 3). Furthermore, the influence of the various
experimental parameters on the polymer yield was investigated
in this study. For some conditions, however, yields were low,
but these results were convincing as they were promising
enough to launch a systematic study of the reaction.
First, NBD monomer was polymerized in various media by
ROMP. Thus, Ru(II) complex (3) (0.025 mmol, 0.014 g) was
dissolved in 3 mL of the following organic solvents under Ar
atmosphere: 1,2-dichloroethane, toluene, THF, dichloromethane,
and dimethylsulfoxide. The same controlled polymerization conditions were carried out. The results of ROMP of NBD with the
solvents are summarized in Table I. No precipitation took place
when the reaction was carried out in dimethylsulfoxide. The polymerization of NBD using the catalyst was first examined in
CH2Cl2 at 25 C (Table I, entry 1). The viscosity of the solution
increased more rapidly with CH2Cl2 in comparison to the other
solvents when it was stirred at 25 C, which means the reaction
time was shorter. The resulting mixture was diluted and poured
into MeOH by stirring continuously, giving the polymer as a
white-yellow solid precipitate in a comparatively higher yield
(55%). In all cases, the polymer was identified by 1H-NMR
spectroscopy. Because the highest yield of polymer was obtained
in dichloromethane, this particular solvent was chosen as the
reaction medium for all subsequent investigations.

Scheme 2. Synthetic route of (3).

WWW.MATERIALSVIEWS.COM

WILEYONLINELIBRARY.COM/APP

J. APPL. POLYM. SCI. 2012, DOI: 10.1002/APP.37867

ARTICLE
Table II. Polymerization of NBD with Various [C]/[M] Ratiosa

Scheme 3. ROMP of NBD. [Color figure can be viewed in the online


issue, which is available at wileyonlinelibrary.com.]
a

To further explore the catalytic potential of Ru(II) complex (3)


concerning the ROMP of NBD, the monomer/catalyst ratio was
increased (reactions at RT, in dichloromethane, and 18 h). The
reactions were carried out for 1/200, 1/500, 1/1000, 1/1500, and
1/2000 catalyst/monomer ratio (Table II). The ratios of 50/1
and 100/1 were not preferred because of their low polymerization yield. The conversion slightly increased for Ru(II) complex
(3), as the monomer/catalyst ratio was driven up to 1000/1
NBD with a maximum conversion of 55%. For the higher
monomer/catalyst ratio, the conversions did not change significantly. As a result, optimum [M]/[C] ratio for the ROMP of
NBD was found to be 1000/1.
For ROMP reactions of NBD, a linear increase of the conversion
versus time was observed within the time period of 18 h at RT.
After this period, the maximum conversion for NBD polymerization was reached. The solvent (CH2Cl2) amount was determined as 3 mL. Polymer product could not solve in any solvent
when monomer was polymerized in the lower amount of solvent (<3 mL) despite the higher polymerization yield. The
lower polymerization yield was obtained in the higher amount
of solvent.
The polymerization reactions were carried out at RT and at
40 C reflux for 18 h in 3 mL of CH2Cl2 at constant [C]/[M]
and other variables were kept constant. The polymerization
yield was raised to 80% when the reaction was performed at
reflux for 18 h. It was observed that when the reaction was carried out for 6 h at reflux, the yield was not changed. Therefore,
optimum temperature was decided as for 6 h at 40 C reflux
temperature because of shorter reaction time.
The role of the cocatalyst precursor on the polymerization process was emphasized. When the polymerization of NBD was
tested with pure TMSD (without catalyst, 1 mL of a NBD solution in dichloromethane, and [NBD]/[TMSD] 1000/1, RT
and 40 C, and 6 h), no polymerization was observed. In a
related study on the use of vinylideneRu complexes to catalyze
Table I. Polymerization of NBD in Various Solventsa

Entry

Solvent

Yield %

Dichloromethane

55

Toluene

27

Tetrahydrofuran

30

1,2-Dichloroethane

20

Dimethylsulfoxide

Reactions at RT; [NBD]/[Ru] 1000; and 6 h.

J. APPL. POLYM. SCI. 2012, DOI: 10.1002/APP.37867

Entry

[C]/[M]

Yield %

1/200

25

1/500

35

1/1000

55

1/1500

56

1/2000

57

Reactions at RT, in CH2Cl2; [NBD]/[Ru] 1000; and 18 h.

the ROMP of NBD derivatives, Katayama and Ozawa reported


the unrivaled superiority of the trimethylsilylvinylidene species
over various other aliphatic and aromatic derivatives.29
Although no exact reason for this superiority was given, the
authors assumed that the particularly high stability of silylvinylidene complexes might account for their outstanding catalytic
activity.
However, when NBD was tested with Ru catalyst without being
activated with TMSD ([NBD]/([Ru] 1000/1, 3 mL dichloromethane, and 6 h), a conversion of 5% was observed. As the
reaction was not activated by TMSD yet, the initial metalcarbene complex must have been resulted from a reaction between
the catalyst and the olefin substrate. The catalyst performance is
altered significantly when the reactions are initiated by a small
amount of TMSD.
When a diazo compound was added to the reaction mixture,
polymerization took place with no apparent induction time and
high yields of polynorbornadienes (PNBDs) were obtained. We
therefore assume that the diazo compound reacts with the precatalyst Ru(II)amine complex to form a highly reactive coordinatively unsaturated Rucarbene species that initiate the polymerization of NBD with no (or very short) observable
induction time. Because of more active Rucarbene complexes,
80% conversion was reached, which was much higher than in
the case when no activator was used. The polymer exhibited a
 w =M
 n 3.53). The
broad molecular weight distribution (M
broadening of the molecular weight distribution was caused by
the generation of multireactive sites (in the case of the activation in situ and by TMSD).
The rate of addition and the amount of diazo compound added
to the monomer and catalyst solution at RT were investigated.
As a result, it seemed to be the most efficient reaction when the
original standard conditions (0.1 mmol of TMSD and 30 min
addition time) were followed. Modifying the concentration of
the TMSD solution or the rate of the syringe pusher used to
Table III. Properties of PNBD Formed with the Catalytic Systems (2) and
(3) when Activated with a Catalytic Amount of TMSDa
Mn

Mw

PDI

Yield %

[RuCl2(p-cymene)]2 (2)

22,450

48,600

2.16

60

Ru(II) complex (3)

34,300

121,300

3.53

80

Reactions at 40 C reflux temperature, in CH2Cl2; [NBD]/[Ru] 1000;


and 6 h.
a

WILEYONLINELIBRARY.COM/APP

ARTICLE

Scheme 4. Possible mechanism for the formation of the initial metalcarbene in the presence of TMSD.

add TMSD, as distinct from literature, did not result in any significant improvement and sometimes had a detrimental influence on the polymerization outcome. Instantaneous addition of
TMSD led to polymers with a more broad distribution of high
molecular weights.
As a result of the experiments committed in this study, optimum conditions for polymerization were detected. Polymerization was carried out with 1000/1 as an M/C ratio, in 3 mL of
dry dichloromethane, at 40 C reflux temperature, and in 6 h.
In a final set of experiments, a ROMP reaction of NBD was run
by [RuCl2(p-cymene)]2 (2) and Ru(II) complex (3). The results
obtained using these catalysts are summarized in Table III. The
catalytic system with amine ligands-substituted Ru complex (3)
was more active. The ligands also play an important role when
dealing with ROMP. It is well known that the ligand nature can
have profound and largely unpredictable effects on the catalytic
activity of the coordination complexes in the polymerization
reactions and present results strongly support this fact. It exhibited a higher molecular weight and a better catalytic activity
because of the steric and electronic properties of amine ligand.
The polydispersities in these catalytic systems were broad, which
indicated that the polymerization was subjected to transfer
reactions.
Possible mechanism for the formation of the catalytically active
species in ROMP reactions is given in Scheme 4.
The most likely process for the initiation of the ROMP of NBD
involves the coordination of the diene to the metal center followed by ring-opening, 1,2-hydrogen shift and alkylidene ligand
formation. Such a transformation can result from the different
types of coordination of NBD ligand to the metal center. In
the presence of an excess of NBD, the weakly coordinating
alkene bond can be substituted by NBD molecule leading to
the formation of alkylidene ligand and next metallacyclobutane
unit.

WWW.MATERIALSVIEWS.COM

Structural Characterization of the PNBD


The 1H-NMR spectrum of PNBD demonstrated five well-separated groups of signals. The formation of an unsaturated PNBD
with characteristic signals of the olefinic protons at ca. 5.6 and
5.4 ppm appeared simultaneously (protons denoted 5,6 and
2,3). From integral ratio of the signals due to protons denoted
1,4 at d 3.6 (cis double bond) and d 3.2 (trans double bond)
ppm appeared. H1/H4 50 : 50 was found from the integrated
areas under our reaction conditions. Two protons denoted 7
appeared at d 2.4 and 1.3 ppm.
Further proof for the successful synthesis of PNBD was
obtained by FTIR analysis. The absorption bands at 2960 and
2915 cm1 were attributed to olefinic CAH stretching. The
intense bands of 1681 and 1425 cm1 were attributed to aliphatic C
C and CAH stretching in the spectrum of PNBD,
unlike from NBD monomer.
The polymer formed in high yield and had a simple unimodal
molecular weight distribution as proofed by GPC analysis of the
 n ), weightpolymer. The number-average molecular weight (M



average molecular weight (Mw ), and PDI (Mw =Mn ) were deter n 34,320 g/mol,
mined by GPC. According to GPC trace, M
 w 121,380 g/mol, and M
 w =M
 n 3.53 results were obtained.
M
From the TGA of PNBD, the degradation occurred in several
stages. The first stage was observed at 100 C. The second stage
of decomposition commenced at 100300 C, and the last and
max decomposition stage was observed at 300650 C. The maximum peak temperature was about 425 C. The total weight loss
was about 70% at this temperature range. The remaining portion of the organic material is carbonized.
CONCLUSIONS

Consequently, the synthesis of a new class of Ru-based catalyst


coordinated to amine ligand that exhibits activity in ROMP
reaction of NBDs was achieved. The ROMP activity of the

WILEYONLINELIBRARY.COM/APP

J. APPL. POLYM. SCI. 2012, DOI: 10.1002/APP.37867

ARTICLE

systems in this study increased dramatically when TMSD was


added to activate the catalytic systems to form highly reactive
coordinatively unsaturated Rucarbene species. Under appropriate conditions, monomer conversion was 80% with a relatively
broad molecular weight distribution for ROMP. These can be
further narrowed by stopping the reaction at lower monomer
conversions. When the new synthesized Ru-based catalyst was
compared with the commercially Ru(II) arene dimer ([RuCl2(pcymene)]2 (2)), it exhibited a higher molecular weight and a
better catalytic activity because of the steric and electronic properties of amine ligand. The polymers with high molecular
weight and high yield were obtained with 1000/1 as a M/C ratio
when CH2Cl2 was used as solvent and TMSD was added to the
reaction at 40 C reflux temperature.
Despite the relatively broad molecular weight distribution, the
amines studied can be used as an alternative to ancillary ligands.
The resulting complexes can be used for practical purposes
because of several reasons; the amines are readily available from
commercial products and relatively cheap according to other
ligands, and the production of polymers was obtained with the
higher polymerization yield.

The Scientific and Technological Research Council of Turkey


(TUBITAK; Project No: 108T101) and Ege University Research
Foundation (Project number: 2007 FEN 046) are acknowledged
for the financial support. The authors are grateful to Salih Gunnaz
for NMR analysis.

9. Schrock, R. R.; Murdzek, J. S.; Bazan, G. C.; Robbins, J.;


DiMare, M.; ORegan, M. J. Am. Chem. Soc. 1990, 112,
3875.
10. Mark, H. F. Encyclopedia of Polymer Science, 3rd ed.;
Wiley: Hoboken, NJ, 2004; Vol. 11, p 547..
11. Ivin, K. J.; Mol, J. C. Olefin Metathesis and Metathesis Polymerization; Academic Press: London, 1997; p 317.
12. Weck, M.; Schwab, P.; Grubbs, R. H. Macromolecules 1996,
29, 1789.
13. Dias, E. L.; Grubbs, R. H. Organometallics 1998, 17, 2758.
14. Schneider, M. F.; Lucas, N.; Velder, J.; Blechert, S. Angew.
Chem. Int. Ed. Engl. 1997, 36, 257.
15. Furstner, A.; Langemann, K. J. Am. Chem. Soc. 1997, 119,
9130.
16. Delaude, L.; Demonceau, A.; Noels, A. F. Macromolecules
1999, 32, 2091.

18. Demonceau, A.; Stumpf, A. W.; Saive, E.; Noels, A. F. Macromolecules 1997, 30, 3127.
19. Alvarez, P.; Gimeno, J.; Lastra, E. Organometallics 2002, 21,
5678.
20. Tolman, C.; Parshall, G. W. J. Chem. Ed. 1999, 76, 177.
21. Noels, A. F.; Graziani, M.; Hubert, A. J., Eds. Metals Promoted Selectivity in Organic Synthesis; Kluwer Academic
Publishers: Dordrecht, 1991.

REFERENCES

1. Grubbs, R. H.; Chang, S. Tetrahedron 1998, 54, 4413.


2. Ivin, K. J.; Laverty, D. T.; ODonnell, J. H.; Rooney, J. J.;
Steward, C. Makromol. Chem. 1979, 180, 1989.
3. Makovetsky, K. L.; Finkelshtein, E. S.; Ostrovskaya, I. Y.;
Portnykh, E. B.; Gorbacheva, L. I.; Golberg, A. I.; Ushakov,
N. V.; Yampolskii, Y. P. J. Mol. Catal. 1992, 76, 107.
4. Finkelshtein, E. S.; Makovetskii, K. L.; Yampolskii, Y. P.;
Portnykh, E. B.; Ostrovskaya, I. Y.; Kaliuzhnyi, N. E.; Pritula, N. A.; Golberg, M. S.; Yatsenko, M. S.; Plate, N. A.
Makromol. Chem. 1991, 192, 1.
5. Anderson, A. W.; Merkling, N. G. U. S. Pat. 2,721,189 (1955).
6. Nguyen, S. T.; Johnson, L. K.; Grubbs, R. H. J. Am. Chem.
Soc. 1992, 114, 3974.

J. APPL. POLYM. SCI. 2012, DOI: 10.1002/APP.37867

8. Schwab, P.; Grubbs, R. H.; Ziller, J. W. J. Am. Chem. Soc.


1996, 118, 100.

17. Stumpf, A. W.; Saive, E.; Demonceau, A.; Noels, A. F. J.


Chem. Soc. Chem. Commun. 1995, 1127.

ACKNOWLEDGMENTS

7. Schwab, P.; France, M. B.; Ziller, J. W.; Grubbs, R. H.


Angew. Chem. Int. Ed. Engl. 1995, 34, 2039.

22. Collman, J. P.; Hegedus, L. S.; Norton, J. R.; Finke, R. G.


Principles and Applications of Organotransition Metal
Chemistry; University Science Books: Mill Valley, 1987.
23. Matos, J. M. E.; Lina-Neto, B. S. J. Mol. Catal. A 2004, 22,
81.
24. Colton, R. Coord. Chem. Rev. 1971, 6, 269.
25. Drew, M. G. B. Prog. Inorg. Chem. 1977, 23, 67.
26. Melnik, M.; Sharrock, P. Coord. Chem. Rev. 1985, 65, 49.
27. Bennet, M. A.; Smith, A. K. J. Chem. Soc. Dalton Trans.
1974, 2, 233.
28. Gamez, P.; Steensma, R. H.; Driessen, W. L.; Reedijk, J.
Inorg. Chim. Acta 2002, 333, 51.
29. Katayama, H.; Ozawa, F. Chem. Lett. 1998, 67.

WILEYONLINELIBRARY.COM/APP

You might also like