Vous êtes sur la page 1sur 5

Catalysis Communications 4 (2003) 33–37

www.elsevier.com/locate/catcom

Ring-opening polymerization of D ,L -lactide with bis(trimethyl


triazacyclohexane) praseodymium triflate
a,*
Randolf D. K€
ohn , Zhida Pan a, Junquan Sun b, Chengfeng Liang b

a
Department of Chemistry, University of Bath, Claverton Down, Bath BA2 7AY, UK
b
College of Material Science and Chemical Engineering, Zhejiang University, Hangzhou 310027, PR China
Received 16 August 2002; received in revised form 29 October 2002; accepted 29 October 2002

Abstract

Ring-opening polymerization of D ,L -lactide (LA) was catalyzed by the well-defined bis(trimethyl triazacyclohexane)
praseodymium triflate ½ðMe3 TACÞ2 PrðOTfÞ3  (Cat) for the first time. The effect of the [LA]/[Cat] molar ratio, solvents,
temperature, and reaction time were investigated in detail. The results show that the praseodymium catalyst is very
efficient in the melt polymerization, giving D ,L -PLA with molecular weights of 104 in good conversion. The importance
of the triazacyclohexane was shown by comparison to the poor catalytic activity of PrðOTfÞ3 .
Ó 2002 Elsevier Science B.V. All rights reserved.

Keywords: Ring-opening polymerization; Polylactide; Praseodymium; Lanthanide triflates; Lewis acid catalyst

1. Introduction investigated and among the numerous polyesters,


polylactides are no doubt the most attractive and
Biodegradable polymers have raised much in- hopeful class for biomedical materials.
terest over the last two decades not only in the Because there is only L -lactic acid enzyme ex-
fundamental research, but also in clinical practice isting in human bodies, poly(L -lactide) has more
[1–3]. As biodegradable materials, they should advantages than poly(D ,L -lactide) in the clinical
have good biocompatibility, and be easily hydro- application. Optically pure poly(L -lactide) is a
lyzed in creature bodies, and their non-toxic deg- crystalline, hard and tough engineering plastic,
radation products will be removed during normal possibly used in the internal fixation of bone
metabolism. With the aid of biological substitutes, fractures. However, poly(D ,L -lactide) having a
creatures like human beings are able to rebuild random stereo sequence is an amorphous trans-
their injured tissue. In view of these features, parent material, which can be used for the pro-
mainly aliphatic polyesters have been extensively duction of transparent films and glues [3].
Besides the enzymatic polymerization, the po-
*
Corresponding author. Tel.: +44-1225-383305; fax: +44-
lymerization of lactide can be promoted by three
1225-386231. kinds of initiators with various mechanisms. First
E-mail address: r.d.kohn@bath.ac.uk (R.D. K€
ohn). of all, strong acids or some carbenium ion donors,

1566-7367/02/$ - see front matter Ó 2002 Elsevier Science B.V. All rights reserved.
PII: S 1 5 6 6 - 7 3 6 7 ( 0 2 ) 0 0 2 4 4 - 3
34 R.D. K€
ohn et al. / Catalysis Communications 4 (2003) 33–37

for examples, triflic acid [4] and its esters, can LnðOTfÞ3 Lewis acid catalysts. Here we report the
initiate a cationic polymerization. Secondly, alkali preliminary results of ring-opening polymerization
metal alkoxides, phenoxides and carboxylates [3] of D ,L -LA by this catalyst (see Fig. 2).
are active to initiate an anionic polymerization of
lactide. The last but not the least, so called ‘‘co-
ordination–insertion’’ catalysts, based on organo- 2. Results and discussion
metallic complexes such as ½ðn-BuOÞ3 Al2 and
SnðOctÞ2 [5]. The lactide plays temporarily the role ½ðMe3 TACÞ2 PrðOTfÞ3  was found to be able to
of a ligand coordinated to the metal atom by the catalyze the polymerization of D ,L -lactide in vari-
carbonyl O atom activating the lactide for inser- ous solvents without the requirement of any addi-
tion according to Fig. 1. Some lanthanide catalysts tional reagent. The optimal yield (95% at
belonging to this category, for instance LnðOiPrÞ3 [LA]:[Cat] ¼ 1000) and molecular weight of the
[6], LnðacacÞ3 [7], LnðnaphÞ3 ; LnðOctÞ3 and others polymer (18,000) was obtained in melt polymeri-
[8], were found to be efficient in the polymerization zation at 170 °C (18 h). Under the same conditions,
of lactide. Lewis acid catalysts without OR- or PrðOTfÞ3 was a poor catalyst giving low molecular
similar nucleophilic groups, e.g. SnðOTfÞ2 and weight D ,L -PLA (5700) in medium yields (74%).
ScðOTfÞ3 [9], can be used when an alcohol is added Thus, the additional ligands have significantly im-
as the initiating group. proved the catalyst. GPC showed a much lower Mn
During the last 10 years, N-substituted 1,3,5- (6200) and a PDI of 2.72 indicating fast inter- and
triazacyclohexane ( R3 TAC ), which can act as a six- intramolecular transesterification [12] leading to an
electron donor ligand, is causing more and more average of 21 polymer chains per metal. This was
interest in coordination chemistry and catalysis. also confirmed by NMR analysis of the polymer.
This ligand is easily accessible and highly variable The 13 C NMR region for the methane carbon at-
due to its simple preparation from primary amines oms showed two major signals at 69.1 and 68.9
and formaldehyde. The acute N-metal-N angle ppm (ratio 1:3) expected for poly-rac-D ,L -lactide
of around 60° makes triazacyclohexanes sterically [13] but also signals at 69.4, 69.3 and 69.0 that can
little demanding facially coordinating ligands. The only derive from transesterification. The major
ring strain upon j3 -coordination facilitates a observable end group is the alcohol HOCMeHC
potential hemi-lability towards j1 -coordination. ð@OÞA with signals at 4.4 (CMeH) ð1 HÞ and 20.5
Thus triazacyclohexanes should be interesting an- (CH3 ), 66.6 (HOC), and 174.7 ppm (C@O) ð13 CÞ
cillary ligands for catalysis. Its chromium(III) [14] at intensities suggesting an NMR-determined
complexes activated by MAO are highly active ho- Mn of only 3200. No additional signals of similar
mogeneous ethylene polymerization catalysts that intensity for the other end group could be detected.
resemble the Phillips catalyst in many important However, much smaller signals at 2.8 ð1 HÞ and 37.1
properties [10]. The first lanthanide complex with and 72.4 ppm ð13 CÞ may correspond to protonated
triazacyclohexane, ½ðMe3 TACÞ2 PrðOTfÞ3 , was Me3 TAC of approximately the amount added with
synthesized recently [11] as modification to existing the catalyst. Thus, we suspect that hydroxide from
deprotonation of trace amounts of water in the
lactide with Me3 TAC can act as the initiator

Fig. 1. Mechanism of the lactide insertion into metal alkoxides. Fig. 2. Molecular structure of ðj3 -Me3 TACÞ2 PrðOTfÞ3 .
R.D. K€
ohn et al. / Catalysis Communications 4 (2003) 33–37 35

leading to carboxylate end groups. Introduction of


nitrogen containing some oxygen to the system
leads to a sharp drop in activity and a deeply col-
ored polymer from degradation of the polylactide.
We have studied the influence of the polymeriza-
tion conditions on the catalyst performance in an
attempt to optimize the polymerization conditions
for high molecular weight polymers with high
monomer conversion.

2.1. The effect of solvent on the polymerization

The polymerization has been carried for 24 hr


Fig. 3. Effect of polymerization temperature (conditions: [LA]/
in refluxing THF, dichloromethane, ethyl acetate [Cat] ¼ 1000, 18 h, vacuum).
and toluene (at 80 °C), respectively (Table 1).
Compared with the melt polymerization at 170 °C
the yields and the molecular weights were lower
but still better than for PrðOTfÞ3 .

2.2. The effect of polymerization temperature and


time

Figs. 3 and 4 show the polymerization results


obtained at different temperature and time. The
increase in yield follows first-order kinetics up to
about 20 hr followed by a drop in yield at long
reaction time due to transesterification reactions
leading to soluble low molecular weight oligomers
(see e.g. [12]). The molecular weight follows a
similar trend. GPC results of the polymers ob-
Fig. 4. Effect of polymerization time (conditions: [LA]/
tained after 5 (PDI ¼ 1.77) and 18 hr (PDI ¼ 2.72) [Cat] ¼ 1000, 170 °C, vacuum).
showed nearly identical numbers of polymer
chains per metal (21 and 22) indicating that the
production of cyclic polymer chains from transe- initial stage, and reached the maximum values at
sterification and reinsertion of the cyclic esters into 170 °C, then followed by a rapid decrease with a
the polymer chains are in equilibrium after 5 h and broader distribution of molecular weight after-
the slowing increase and eventual drop in wards. This is very similar to the polymerizations
Mg ð Mw Þ with time is due to increasing PDI. of lactide catalyzed by stannous octanoate or
The molecular weight of the resulting poly(D , NdðacacÞ3 where the optimal temperature is near
L -lactide) increased with the temperature at the 140 °C [8,15]. GPC results for the polymers ob-

Table 1
Effect of various solvents on the polymerization of D ,L -lactide

a
Solvent THF Dichloromethanea Ethyl acetatea Toluenea No solventb
Yield (%) 88 82 79 84 95
Mg ð104 Þ 1.03 0.83 1.26 1.18 1.80
a
[LA]/[Cat] ¼ 1000, reflux or 80 °C, 24 h.
b
[LA]/[Cat] ¼ 1000, 170 °C, 18 h, vacuum.
36 R.D. K€
ohn et al. / Catalysis Communications 4 (2003) 33–37

tains at 120° (PDI ¼ 1.85), 180° (2.72) and 200 °C of the apparent first-order rate constant kapp
(3.46) showed average number of polymer chains against the catalyst concentration is shown in Fig.
per metal is 14, 21 and 45, respectively. 6. A fractional order of 1/2 in the catalyst con-
centration is found. This is not uncommon for ring
2.3. The effect of the molar ratio [LA]/[Cat] on the opening polymerization of lactones ([16] and ref-
polymerization erences therein) and can be explained by a fast
equilibrium between an active mononuclear com-
Fig. 5 shows both the polymer yield and mo- plex with an inactive dinuclear complex as the
lecular weight as a function of the molar ratio resting state of the catalyst. The solution behavior
[LA]/[Cat] when polymerization was carried out at of the precatalyst ½ðMe3 TACÞ2 PrðOTfÞ3  has al-
170 °C for 18 h. The molecular weight increased ready shown a tendency to dimerize with loss of
gradually but not linearly with the [LA]/[Cat] ratio one of the two Me3 TAC [11]. Formation of alk-
as a result of transesterification. A logarithmic plot oxide groups as good bridging ligands during the
catalysis should favor this dimerization.

3. Conclusions

The melt polymerization of D ,L -lactide with


½ðMe3 TACÞ2 PrðOTfÞ3  can efficiently be carried
out at 170 °C and is first order in lactide and 1/2
order in catalyst concentration and is accompa-
nied by a large degree of transesterification. The
improvement of the catalyst performance over
PrðOTfÞ3 by the triazacyclohexane ligands may be
due to improved initiation (via deprotonation of
water) and stabilization of active mononuclear
complexes.
Fig. 5. Effect of [LA]/[Cat] on the polymerization (conditions:
170 °C, 18 h, vacuum).

4. Experimental

Materials. Solvents were dried according to


standard methods and collected by distillation
before use. D ,L -lactide used in this work was pre-
pared from lactic acid according to the literature
procedures [17], and was recrystallized from ethyl
acetate for at least three times to remove as much
meso-lactide as possible, which intensively absorb
moisture and therefore affect the polymerization.
The melting point of the monomer was determined
as 123 °C, in good agreement with those reported
in the literature. The lanthanide catalyst was syn-
thesized as previously described [11]. The complex
does not show thermal decomposition up to 170
Fig. 6. Plot of lnðkapp Þ versus ln([Cat]) for the polymerization at °C (the melting point of 154 °C is retained after
170 °C. The solid line represents the linear fit of experimental cooling) but decomposes above 180 °C and turns
data (slope ¼ 0.48(2), R ¼ 0:998). brown. All manipulations of air- and moisture-
R.D. K€
ohn et al. / Catalysis Communications 4 (2003) 33–37 37

sensitive compounds were carried out under an National Natural Science Foundation of China
atmosphere of nitrogen or argon using standard (20074028).
Schlenk line or glove box techniques.
Polymerization. The polymerization was per-
formed in a 20 ml ampoule that was thoroughly References
dried and purged with argon for several cycles.
After addition of D ,L -LA, catalyst (and solvent, if [1] R.K. Kulkarni, C. Pani, C. Neuman, F. Leonard, Arch.
in solution polymerization) sequentially, the am- Surg. 93 (1966) 839–843.
[2] F.E. Kohn, J.W.A. Vandenberg, G. van de Ridder, J.
poule was sealed under high vacuum (except for Appl. Polym. Sci. 29 (1984) 4265–4272.
solution polymerization), and then placed in an [3] H.R. Kricheldorf, Chemosphere 43 (2001) 49–54, and
oven at constant temperature. The melt polymer- references therein.
ization was quenched by adding some ethanol, the [4] H.R. Kricheldorf, R. Dunsing, Makromol. Chem. 187 (7)
(1986) 1611–1625.
polymer was dissolved in chloroform, and precip-
[5] J.W. Leenslag, A.J. Pennings, Makromol. Chem. 188
itated by ethanol containing 5% hydrochloric acid, (1987) 1809–1814.
and washed with ethanol for several times. [6] Y.Q. Shen, F.Y. Zhang, Y.F. Zhang, Z.Q. Shen, Chin. J.
Poly(D ,L -lactide) was dried at 50 °C under vac- Polym. Sci. 2 (1995) 222–227.
uum. [7] J.F. Liu, Z.Q. Shen, J.Q. Sun, Chin. J. Appl. Chem. 12
Measurements. The intrinsic viscosity of PLA in (1995) 59–61.
[8] J.Q. Sun, L.T. Wu, Chin. J. Polym. Sci. 14 (1996) 4, see
CHCl3 was determined by Ubbelohde viscometer at also 324–329.
25 °C. The viscosity-averaged molecular weight was [9] M. Moller, F. Nederberg, L.S. Lim, R. Kange, C.J.
calculated according to the following equation [18]: Hawker, J.L. Hedrick, Y.D. Gu, R. Shah, N.L. Abbott,
0:77
J. Polym. Sci., Part A: Polym. Chem. 39 (2001) 3529–
½g ¼ 2:21  104 M g ðdl=gÞ: 3538.
[10] R.D. K€ ohn, M. Haufe, S. Mihan, D. Lilge, Chem.
Additionally, molecular weights and their distri- Commun. (2000) 1927–1928.
bution of D ,L -PLA was measured on a Waters As- [11] R.D. K€ ohn, Z. Pan, G. Kociok-K€ ohn, M.F. Mahon, J.
sociates M208 Gel Permeation Chromatograph Chem. Soc., Dalton Trans. (2002) 2344–2347.
[12] P. Dubois, C. Jacobs, R. Jer^
ome, P. Teyssie, Macromol-
with THF as an eluent at 37 °C for some samples.
ecules 24 (1991) 2266–2270.
Generally, good agreement was found between Mw [13] M.H. Chisholm, S.S. Iyer, D.G. McCollum, M. Pagel,
(GPC) and Mg . Average numbers of polymer chains U. Werner-Zwanziger, Macromolecules 32 (1999) 963–
per metal were calculated as yield  ð½LA=½CatÞ  973.
MLA =Mn . DSC study of typical polymer sample was [14] W.M. Stevels, M.J.K. Ankone, P.J. Dijkstra, J. Feijen,
Macromolecules 29 (1996) 6132–6138.
carried out on a Perkin–Elmer DSC7 machine, with
[15] P.Z. Hu, Master Thesis, Chemistry Department, Zhejiang
a glass transition temperature of 51.5 °C. University, 1988.
[16] B.M. Chamberlain, M. Cheng, D.R. Moore, T.M. Ovitt,
E.B. Lobkovsky, G.W. Coates, J. Am. Chem. Soc. 123
Acknowledgements (2001) 3229–3238.
[17] R.K. Kulkarni, E.G. Moore, A.F. Hegyeli, Biomed. Mater.
Res. 5 (1971) 169–175.
This work has been financially supported by the [18] A. Schindler, D. Harper, J. Polym. Sci., Part A: Polym.
Royal Society of the United Kingdom and the Chem. 17 (1979) 2593–2599.

Vous aimerez peut-être aussi