ArticlePDF Available

Insertion Copolymerization of Difunctional Polar Vinyl Monomers with Ethylene

Authors:

Abstract

A single-step synthesis, structural characterization and application of a neutral, acetonitrile ligated, palladium−phosphinesulfonate complex [{P ∧ O}PdMe(L)] (P ∧ O = κ 2-P,O−Ar 2 PC 6 H 4 SO 2 O with Ar = 2-MeOC 6 H 4 ; L = CH 3 CN) (3) in coordination/insertion copolymerization of ethylene with difunctional olefin is investigated. In a significant development, complex 3 was found to catalyze insertion copolymerization of industrially relevant 1,1-disubstituted difunctional vinyl monomers for the first time. Thus, insertion copolymerization of ethyl-2-cyanoacrylate (ECA or super glue) and trifluoromethyl acrylic acid (TFMAA) with ethylene produced the corresponding copolymers with 6.5% ECA and 3% TFMAA incorporation. Increasing the concentration of difunctional olefins led to higher incorporation but at the expense of lower activities. These observations indicate that complex 3 tolerates difunctional vinyl monomers and provides direct access to difunctional polyolefins that have not been attempted before.
Insertion Copolymerization of Difunctional Polar Vinyl Monomers
with Ethylene
Shahaji R. Gaikwad,
Satej S. Deshmukh,
Rajesh G. Gonnade,
P. R. Rajamohanan,
§
and Samir H. Chikkali*
,,
Polyolen Lab, Polymer Science and Engineering Division,
Center for Material Characterization, and
§
Central NMR Facility,
CSIR-National Chemical Laboratory, Dr. Homi Bhabha Road, Pune 411008, India
Academy of Scientic and Innovative Research, Anusandhan Bhawan, 2 RaMarg, New Delhi 110001, India
*
SSupporting Information
ABSTRACT: A single-step synthesis, structural characterization and
application of a neutral, acetonitrile ligated, palladiumphosphinesulfonate
complex [{PO}PdMe(L)] (PO=κ2-P,OAr2PC6H4SO2Owith Ar = 2-
MeOC6H4;L=CH
3CN) (3) in coordination/insertion copolymerization of
ethylene with difunctional olen is investigated. In a signicant development,
complex 3was found to catalyze insertion copolymerization of industrially
relevant 1,1-disubstituted difunctional vinyl monomers for the rst time.
Thus, insertion copolymerization of ethyl-2-cyanoacrylate (ECA or super
glue) and triuoromethyl acrylic acid (TFMAA) with ethylene produced the corresponding copolymers with 6.5% ECA and 3%
TFMAA incorporation. Increasing the concentration of difunctional olens led to higher incorporation but at the expense of
lower activities. These observations indicate that complex 3tolerates difunctional vinyl monomers and provides direct access to
difunctional polyolens that have not been attempted before.
The seminal work of Karl Ziegler and Giulio Natta laid the
foundation of coordination/insertion polymerization of
olens.
1
Tremendous progress has been made, and today the
world produces roughly 145 million tons of polyolens
annually.
2
Polyethylene (PE) occupies the top position (in
terms of production) among the polyolens, and various grades
of PE are commercially produced.
3
PE is inherently a long
chain of hydrophobic methylene repeat units without any
functional group on the backbone. This partly limits the
potential application of PE in adhesives, binders, paints,
printing ink, dying, etc. Incorporation of even a small amount
of functional groups in PE can signicantly enhance these
material properties and can further broaden the PE application
window. However, due to the high oxophilicity of early
transition metal based ZieglerNatta type catalysts, the
functional group on the polar vinyl monomer coordinates to
the metal (occupying the vacant site) that generally leads to
catalyst poisoning. This limitation has been partly addressed by
postfunctionalization of PE,
4
ADMET polymerization of
functionalized dienes,
5
or OMRP of polar vinyl monomers
with ethylene.
6
However, these strategies suer from typical
issues associated with free radical polymerization. Therefore, it
has been a long cherished dream of organometallic chemists to
synthesize functionalized PE in a single step via insertion
(co)polymerization of ethylene with industrially relevant polar
vinyl monomers.
7,8
Despite the signicant progress in olen
polymerization, insertion (co)polymerization of functional
olens remained inaccessible until recently.
9
Apart from a few random attempts, the insertion (co)-
polymerization of functional olens mainly explored two ligand
classes, (a) α-diimine ligands
10
and (b) the phosphinesulfonate
ligand system.
11
The Brookhart system mainly copolymerizes
acrylates with high branching density due to extensive chain-
walking.
12
However, the phosphinesulfonate system displays
broad functional group tolerance, and various polar vinyl
monomers such as acrylate,
13
acrylonitrile,
14
vinyl acetate,
15
vinyl ethers,
16
acrylic acid,
17
and vinyl chloride
18
could be
incorporated. The excellent performance of the phosphinesul-
fonate system further rose the scientic aspirations, and
subsequent investigations focused on structural ne-tuning of
the catalyst.
19
Three crucial positions as depicted in Figure 1
were identied, among which the eect of the donor group D
has been extensively studied (see Figure 1: C1).
20
The above investigations mainly focused on insertion
copolymerization of monosubstituted polar vinyl monomers,
wherein one functional group is introduced into the polymer
Received: August 11, 2015
Accepted: August 13, 2015
Published: August 17, 2015
Figure 1. Structural tuning of C1 at designated positions and insertion
copolymerization to (a) monofunctional polyolen (FPO) and (b)
1,1-disubstituted difunctional polyolen (dFPO).
Letter
pubs.acs.org/macroletters
© 2015 American Chemical Society 933 DOI: 10.1021/acsmacrolett.5b00562
ACS Macro Lett. 2015, 4, 933937
chain per insertion. A paradigm shift could be achieved if two
functional groups can be introduced in a polymer chain per
insertion (of a functional olen), resulting in a doubly polar
polymer. Although highly desirable, there are no reports on
insertion copolymerization of 1,1-disubstituted difunctional
olens to date.
Herein we report synthesis of the Pdphosphinesulfonate
acetonitrile complex (3) and its implications in insertion
copolymerization of ethylene with industrially relevant 1,1-
disubstituted difunctional vinyl monomers (see Figure 1;
dFPO-difunctional polyolen) for the rst time.
In our pursuit to realize this goal, we attempted synthesis of
acetonitrile complex 3(Scheme 1). Ligand 1was treated with
[(COD)PdMeCl], and within 10 min the acetone dimer 2
(78%) was obtained.
21
Treatment of AgBF4in the presence of
acetonitrile (10 equiv) aorded the anticipated complex 3
which was isolated in good yield (92%) (Scheme 1, P-I).
An alternative, one-step synthesis was also explored. Reaction
of 1with [(COD)PdMeCl] in acetonitrile directly produced
complex 3(Scheme 1, P-II) in excellent yield (91%). The
existence of palladium complex 3was unambiguously
ascertained from spectroscopic and analytical data.
22
31P
NMR of complex 3displayed a characteristic resonance at
21.1 ppm (SI, S1). In a typical proton NMR, the palladium-
bound methyl protons (PdCH3) appeared at 0.18 ppm,
whereas the corresponding methyl carbon (PdCH3) appeared
at 2.4 ppm in a 13C NMR spectrum.
23
The NMR ndings
were further corroborated by electrospray ionization mass
spectrum (ESI-MS +ve mode) which displayed a pseudomo-
lecular ion peak at m/z= 544.97 [M-ACN + Na]+(SI S6). A
single-crystal X-ray structure of complex 3displayed slightly
distorted square planar geometry at palladium (Figure 2) which
is crystallized in the orthorhombic Pbca space group. The
phosphine and the methyl group are mutually cis to each other,
whereas the acetonitrile is situated trans to the phosphine.
24
To shed light on the relative binding strength of donor
solvents, 1 equiv of DMSO was added to complex 3, and the
changes were tracked using proton and phosphorus NMR.
Addition of 1 equiv of DMSO led to complete disappearance of
acharacteristicPdMe (in 3) resonance at 0.18 ppm;
concomitantly a new signal at 0.39 ppm appeared in the 1H
NMR spectrum (SI S8 and Table S1). The new resonance (at
0.39) can be readily assigned to a previously reported DMSO-
coordinated PdMe complex.
13
These experiments indicate
that acetonitrile binding strength is lower or as good as DMSO
binding strength.
The performance of 3in ethylene-functional olen
copolymerization was evaluated, and representative polymer-
ization experiments are summarized in Table 1. The Pd-
complex 3catalyzes copolymerization of acrylonitrile and
ethylene at 95 °C. After polymerization, the volatiles were
evaporated under reduced pressure to obtain solid material.
Insertion copolymerization of acrylonitrile and ethylene
produced the desired copolymer with 9.2% acrylonitrile
incorporation (Table 1, run 11). The performance of 3in
another industrially important monomer, methyl acrylate, was
tested. An incorporation of 9.6% was observed at 0.6 mol/L
concentration, and the polymer molecular weight was 3100 g/
mol (run 12). Thus, catalyst 3tolerated the cyano- as well as
acrylate functional groups and produced functional copolymers
with reasonable molecular weights. Ethyl-2-cyanoacrylate
(ECA) was chosen as a representative 1,1-disubstituted
difunctional olen as both the cyano- and acrylate group are
installed within the same olen. It should be noted that
cyanoacrylates are commonly sold under the trade name Super
Glue or Krazy Glue, and the polymers thereof are
commercially produced and nd applications in various
adhesives. It is one of the largest adhesives produced
worldwide.
25
Interestingly though, very little information exists
on the reactivity of such an industrially highly relevant 1,1-
disubstituted difunctional olen in insertion (co)-
polymerization reaction.
26
Having known that the probability of functional group
coordination (acrylate or cyano) to the metal and subsequent
catalyst deactivation is increased by 2-fold in ethyl-2-
cyanoacrylate, we begin our exploration with a very low (0.03
mol/L) ECA concentration at 5 bar ethylene pressure (Table 1,
run 13). After polymerization, the reactor content was
transferred to a Schlenk ask; volatiles were stripped o; and
the solid mass was washed with excess chloroform to yield solid
polymeric material.
27,28
The proton NMR of the solid was
recorded in deuterated tetrachloroethane at 130 °C in a 10 mm
NMR tube, and the amount of incorporation was determined
(see Figure 3).
29
Typically, 1H NMR of the copolymer
displayed distinct signals at 4.47 (Hc), 2.872.30 (Ha), 1.49,
and 1.38 (Hb) ppm suggesting ECA incorporation. The proton
NMR ndings were further corroborated by 13C NMR, which
Scheme 1. Synthesis of the Acetonitrile-Ligated Pd
Phosphinesulfonate Acetonitrile Complex (3)
Figure 2. Molecular structure of complex 3. Solvent molecules and H
atoms are omitted for clarity, and thermal ellipsoids are drawn at the
50% probability level.
ACS Macro Letters Letter
DOI: 10.1021/acsmacrolett.5b00562
ACS Macro Lett. 2015, 4, 933937
934
revealed a very characteristic signal at 113.7117.3 ppm that
can be assigned to in-chain CN incorporation.
30
A representative HMBC spectrum (SI, S16D) of a low
molecular weight copolymer fraction revealed cross-peaks that
can be assigned to CH2protons (diastereotopic in nature) of
ECA (2.202.80 ppm) and CH2carbons originating from
ethylene repeat units (2830 ppm) (SI sect. 5.2).
22
The 2D
CH correlation spectra displayed cross peaks that established
the connectivity between atype protons to acarbons; a
protons to btype carbon; and vice versa (see Figures S16B
D). In addition to these 2D-NMR experiments, MALDI-ToF-
MS (SI S17AE) also indicated ECA incorporation, and
various copolymer fragments could be identied. This was
further conrmed by comparing the reported ETECA
copolymer produced by the radical polymerization method
30
as well as control experiments.
31
Thus, 1H NMR of a
copolymer obtained in run 13 revealed a very low
incorporation of 0.32%. Increasing the concentration of ECA
resulted in increased incorporation (run 13to15). On the
other hand, increasing ethylene pressure (from 1 to 10 bar) led
to decreased ECA incorporation (run 16, 14, and 17).
Thus, the highest incorporation of 6.5% could be achieved at
ambient ethylene pressure and 0.06 mol/L ECA concentration.
The rigorous washing protocol may wash out the low molecular
weight fraction and might lead to lower polydispersity in some
cases. The DMSO complex of type 3displayed similar reactivity
in the insertion copolymerization of 1,1-disubstituted functional
olen, along with ca. 2.% incorporation (run 18). It is most
likely that ethylene insertion in the PdMe bond leads to chain
propagation, or the 2,1-/1,2-insertion of ECA in the PdMe
bond initiates the polymerization, as the three initiating groups
(IG-1, IG-2, and IG-3) could be detected by 13C NMR (SI
section 6). To further strengthen our hypothesis, we
investigated insertion copolymerization of another 1,1-disub-
stituted functional olen(triuoromethyl acrylic acid:
TFMAA). About 0.7% TFMAA incorporation was observed
at a very low TFMAA concentration (0.06 mol/L run 19).
However, increased incorporation of 3% could be achieved at
higher TFMAA concentration (3.0 mol/L run 110). A
characteristic proton resonance at 2.1 ppm indicated formation
of an ethyleneTFMAA copolymer (SI S48). Furthermore, the
absence of characteristic proton resonance at 3.4 ppm ruled out
the presence of a TFMAA homopolymer (SI S50).
32
An 19F
NMR spectrum (Figure 4) of the oligomeric fraction (accessed
by suspending the copolymer in excess chloroform) revealed
characteristic signals at 65.1, 66.1, 66.7, and 68.4 ppm.
These chemical shifts are similar to those reported for
ethylenetriuoropropene copolymers (the closest copolymer
known)
33
and can be readily assigned to ETTFMAA
oligomers. The absence of a characteristic splitting pattern
and signal broadening hampered further analysis, and the
resonances can be tentatively assigned to structure AD(SI
S51). The NMR ndings were further supported by IR, which
revealed a characteristic (CO) band at 1710 cm1(SI S49),
that can be ascribed to acidic CO and is in line with earlier
Table 1. Insertion Copolymerization of Ethylene with Functional Olens Catalyzed by Complex 3
a
run FO (mol/L) C2H4(bar) % incor.
b
yield (g) Mn (103g/mol)
c
Mw/Mn
c
11 AN (1.2) 5 9.2
d
0.13 ND ND
12 MA (0.60) 5 9.6
d
0.39 3.1 1.3
13 ECA (0.03) 5 0.3 0.90 4.7 1.5
14 ECA (0.06) 5 2.1 0.79 8.2 1.4
15 ECA (0.12) 5 4.9 0.87 6.4 1.7
16 ECA (0.06) 1 6.5 0.20 5.8 1.6
17 ECA (0.06) 10 1.9 1.09 8.3 1.4
18
d
ECA (0.06) 1 2.01 0.21 4.9 1.2
19
e
TFMAA (0.06) 1 0.7 0.37 ND ND
110
e
TFMAA (3.0) 1 3.0 0.07 2.8 1.2
a
Reaction conditions: 3=20μmol in DCM, toluene = 50 mL (toluene + functional olen); temperature = 95 °C, time = 1 h, AN = acrylonitrile,
MA = methyl acrylate, ECA = ethyl-2-cyanoacrylate, TFMAA = triuoromethyl acrylic acid.
b
Incorporation was determined by high-temperature 1H
NMR in C2D2Cl4at 130 °C.
c
Determined by high-temperature GPC at 160 °C in trichlorobenzene against PS standard; ND = not determined.
d
DMSO complex of type 3was used as a catalyst.
e
Toluene = 5/3 mL, TFMAA incorporation was determined by high-temperature 1H NMR in
C2D2Cl4at 130 °C (for run 19) and C6D6+ TCB (10:90) mixture at 120 °C (for run 110).
Figure 3. 1H NMR of the ETECA copolymer in C2D2Cl4at 403 K
(Table 1, run 16).
ACS Macro Letters Letter
DOI: 10.1021/acsmacrolett.5b00562
ACS Macro Lett. 2015, 4, 933937
935
reports. Furthermore, a typical OH band was observed at
3368 cm1, which, together with CO band, indicates the
existence of a carboxylic group in the copolymer. MALDI-ToF-
MS spectra of the copolymer also indicated TFMAA
incorporation and various copolymer fragments could be
identied (see Table S6 and SI S52AC). Interestingly,
TFMAA does not homopolymerize under the current polymer-
ization conditions and thus rules out the possibility of
homopolymer contamination.
22
The reduced polymer yields
with increasing ECA/TFMAA concentration and low molecular
weights indicate reversible inhibition of polymerization, due to
the coordination of functional groups located on the 1,1-
disubstituted olen to the metal. However, the fact that the
copolymers could be obtained suggests that the presence of 1,1-
disubstituted functional olens does not necessarily lead to
detrimental catalyst decomposition.
In summary, a single-step synthetic protocol to access an
acetonitrile-ligated Pdphosphinesulfonate complex (3) was
established. Our investigations demonstrate that insertion
copolymerization of ethylene with 1,1-disubstituted difunc-
tional olen is possible. Complex 3tolerates the two functional
groups located within the same polar vinyl monomer and
successfully copolymerizes ECA and TFMAA with ethylene. A
combination of multinuclear NMR, MALDI-ToF-MS, and
control experiments suggest in-chain ECA incorporation. High-
temperature NMR investigations revealed an unprecedented
6.5% incorporation of ethyl-2-cyanoacrylate and 3% incorpo-
ration of TFMAA in a copolymer. These copolymers represent
a novel class of functional polyolens with two functional
moieties (cyanoacrylate or uoro-acrylic acid) incorporated at a
time. Investigations on the insertion copolymerization of other
industrially relevant 1,1-disubstituted polar vinyl monomers
and the mechanism are currently in progress.
ASSOCIATED CONTENT
*
SSupporting Information
The Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/acsmacro-
lett.5b00562.
Synthetic protocols, text, gures, tables, method to
determine percentage functional olen incorporation,
spectroscopic and analytical data, CIF les, and crystallo-
graphic data/processing method for 3(CCDC 1012218)
(PDF)
AUTHOR INFORMATION
Corresponding Author
*E-mail: s.chikkali@ncl.res.in. Phone: +91 20 25903145.
Author Contributions
All authors have given approval to the nal version of the
manuscript.
Notes
The authors declare no competing nancial interest.
ACKNOWLEDGMENTS
Financial support from DST-India (Ramanujan Fellowship:
SR/S2/RJN-11/2012 and SR/S1/IC-60/2012) is gratefully
acknowledged. We thank Prof. Stefan Mecking for proof
reading and useful discussion; Mrs. D. Dhoble is acknowledged
for HT-GPC analysis. SRG and SSD would like to thank CSIR
and UGC, respectively, for the junior research fellowship.
CSIR-National Chemical Laboratory and SPIRIT (DCPC) is
gratefully acknowledged for additional support. The Authors
acknowledge NMR Center for Advanced Research
(NMRCAR) at CSIR-NCL funded by CSIR, New Delhi,
through 12th FYP project, CSC 0405.
REFERENCES
(1) Ziegler, K.; Holzkamp, E.; Breil, H.; Martin, H. Angew. Chem.
1955,67, 541. Natta, G.; et al. J. Am. Chem. Soc. 1955,77, 1708.
(2) Kamnisky, W. Macromolecules 2012,45, 3289.
(3) Bhoor, J., Jr. Ziegler-Natta Catalysts and Polymerization; Academic
Press: New York, 1979.
(4) Boaen, N. K.; Hillmyer, M. A. Chem. Soc. Rev. 2005,34, 267.
(5) Watson, M. D.; Wagener, K. B. Macromolecules 2000,33, 8963
and the references therein..
(6) Kermagoret, A.; Debuigne, A.; Jerome, C.; Detrembleur, C. Nat.
Chem. 2014,6, 179 and the references therein..
(7) Chung, T. C. Functionalisation of polyolens; Academic Press:
London, 2002.
(8) Franssen, N. M. G.; Reek, J. N. H.; De Bruin, B. Chem. Soc. Rev.
2013,42, 5809.
(9) The topic has been summarized in various articles, see:
(a) Coates, G. W. Chem. Rev. 2000,100, 1223. (b) Boffa, L. S.;
Novak, B. M. Chem. Rev. 2000,100, 1479. (c) Nakamura, A.; Ito, S.;
Nozaki, K. Chem. Rev. 2009,109, 5215. (d) Gaikwad, S. R.;
Deshmukh, S. S.; Chikkali, S. H. J. Polym. Sci., Part A: Polym. Chem.
2014,52, 1. (e) Carrow, B. P.; Nozaki, K. Macromolecules 2014,47,
2541.
(10) Ittel, S. D.; Johnson, L. K.; Brookhart, M. Chem. Rev. 2000,100,
1169.
(11) (a) Drent, E.; van Dijk, R.; van Ginkel, R.; van Oort, B.; Pugh,
R. I. Chem. Commun. 2002, 744. (b) Nakamura, A.; Anselment, T. M.
J.; Claverie, J.; Goodall, B.; Jordan, R. F.; Mecking, S.; Rieger, B.; Sen,
A.; van Leeuwen, P. W. N. M.; Nozaki, K. Acc. Chem. Res. 2013,46,
1438.
Figure 4. 19F NMR (1H decoupled) of the ETTFMAA oligomeric
fraction recorded at 343 K in C2D2Cl4.
ACS Macro Letters Letter
DOI: 10.1021/acsmacrolett.5b00562
ACS Macro Lett. 2015, 4, 933937
936
(12) (a) Johnson, L. K.; Mecking, S.; Brookhart, M. J. Am. Chem. Soc.
1996,118, 267. (b) Mecking, S.; Johnson, L. K.; Wang, L.; Brookhart,
M. J. Am. Chem. Soc. 1998,120, 888.
(13) Guironnet, D.; Roesle, P.; Runzi, T.; Gottker-Schnetmann, I.;
Mecking, S. J. Am. Chem. Soc. 2009,131, 422.
(14) (a) Kochi, T.; Noda, S.; Yoshimura, K.; Nozaki, K. J. Am. Chem.
Soc. 2007,129, 8948. (b) Nozaki, K.; Kusumoto, S.; Noda, S.; Kochi,
T.; Chung, L. W.; Morokuma, K. J. Am. Chem. Soc. 2010,132, 16030.
(15) Ito, S.; Munakata, K.; Nakamura, A.; Nozaki, K. J. Am. Chem.
Soc. 2009,131, 14606.
(16) Luo, S.; Vela, J.; Lief, G. R.; Jordan, R. F. J. Am. Chem. Soc. 2007,
129, 8946.
(17) Runzi, T.; Frohlich, D.; Mecking, S. J. Am. Chem. Soc. 2010,132,
17690.
(18) Low incorporation, mostly only at the chain end, see: Leicht, H.;
Gottker-Schnetmann, I.; Mecking, S. Angew. Chem., Int. Ed. 2013,52,
3963.
(19) For structural modication of the phosphinesulfonate ligand
system see: (a) Carrow, B. P.; Nozaki, K. J. Am. Chem. Soc. 2012,134,
8802. (b) Piche, L.; Daigle, J. C.; Rehse, G.; Claverie, J. P. Chem. - Eur.
J. 2012,18, 3277. (c) Zhou, X.; Jordan, R. F. Organometallics 2011,30,
4632. (d) Kim, Y.; Jordan, R. F. Organometallics 2011,30, 4250.
(20) (a) Neuwald, B.; Olscher, F.; Gottker-Schnetmann, I.; Mecking,
S. Organometallics 2012,31, 3128. (b) Ito, S.; Kanazawa, M.;
Munakata, K.; Kuroda, J. I.; Okumura, Y.; Nozaki, K. J. Am. Chem.
Soc. 2011,133, 1232. (c) References 13 and 14a.
(21) The synthetic protocol for acetonedimer synthesis was
adopted from the following report: Runzi, T.; Guironnet, D.;
Goettker-Schnetmann, I.; Mecking, S. J. Am. Chem. Soc. 2010,132,
16623.
(22) See Supporting Information for spectroscopic, analytical data,
and more details.
(23) Similar chemical shift values (31P, 1H, 13C NMR) have been
observed for pyridine, lutidine, DMSO complexes; see for example
reference 13.
(24) The PdN (from the acetonitrile donor) bond length is 2.091
Å. A CSD bond length search for palladium(II)-bound acetonitrile
complexes suggested PdN bond length in the range of 2.0182.111
Å. See: (a) Groux, L. F.; Weiss, T.; Reddy, D. N.; Chase, P. A.; Piers,
W. E.; Ziegler, T.; Parvez, M.; Benet-Buchholz, J. J. Am. Chem. Soc.
2005,127, 1854. (b) Falvello, L. R.; Fernandez, S.; Navarro, R.; Rueda,
A.; Urriolabeitia, E. P. Organometallics 1998,17, 5887.
(25) (a) Coover, H. W., Jr. Handbook of adhesives; Skeist, J., Eds.;
Reinhold Publishing Corporation Chapman and Hall Ltd.: London,
1962; pp 409414. (b) Coover, H. W., Jr.; Shearer, N. H., Jr. Adhesive
compositions containing alkyl esters of cyanoacrylic acid. U. S. Patent
2,794,788, June 4, 1957. (c) Badejo, I. T.; Warren, T. M.; Mccrum, B.
U.S. Patent 2009/0326095A1, Dec. 31, 2009. (d) Schueneman, G. T.;
Attarwala, S.; Brantl, K. R. Cyanoacrylate compositions incorporating
organic micropulp, U. S. Patent 8,287,687, Oct. 16, 2012.
(26) No reports could be found on insertion copolymerization of 1,1-
disubstituted difunctional olens. A scinder search with Insertion
polymerization of 1,1-disubstituted functional olenas key word
resulted in 0hits. Few reports on the insertion copolymerization of
methyl methacrylate were noticed See: (a) Li, X. F.; Li, Y. G.; Li, Y. S.;
Chen, Y. X.; Hu, N. H. Organometallics 2005,24, 2502. (b) See
reference 21.
(27) Ethyl-2-cyanoacrylate is known to polymerize in the presence of
traces of air. Hence, polymerization was carried out under inert gas
conditions, and after polymerization, workup was done under positive
argon ow; for moisture-induced homopolymerization of ethyl-2-
cyanoacrylate see: (a) Mankidy, P. J.; Rajagopalan, R.; Foley, H. C.
Chem. Commun. 2006, 1139. (b) Weiss, C. K.; Ziener, U.; Landfester,
K. Macromolecules 2007,40, 928. (c) Liu, S. L.; Long, Y. Z.; Huang, Y.
Y.; Zhang, H. D.; He, H. W.; Sun, B.; Sui, Y. Q.; Xia, L. H. Polym.
Chem. 2013,4, 5696 and the references therein..
(28) To avoid radical polymerization of ethyl-2-cyanoacrylate, radical
scavenger BHT was added to the polymerization reactor. See
Supporting Information.
(29) The functional olen incorporation was determined by 1H
NMR; a detailed calculation method is presented in the Supporting
Information (SI, section 5).
(30) Poly(ethyleneECA) copolymers were prepared by Eisenbach
et al. using radical initiators and were fully characterized by 13C NMR.
See: (a) Sperlich, B.; Eisenbach, C. D. Acta Polym. 1996,47, 280.
(b) Eisenbach, C. D.; Lieberth, W.; Sperlich, B. Angew. Makromol.
Chem. 1994,223, 81.
(31) Apart from NMR, various control experiments ruled out the
possibility of existence of an ECA homopolymer in the ETECA
copolymer. See SI section 7 for control experiments.
(32) First radical homopolymerization of TFMAA was reported in
2013 by Patil et al. The chemical shift for -CH2- protons next to the
quaternary carbon was reported to be 3.4 ppm. See: Patil, Y.; Hori, H.;
Tanaka, H.; Sakamoto, T.; Ameduri, B. Chem. Commun. 2013,49,
6662.
(33) Ethylenetriuoropropene copolymerization has been recently
reported by Lanzinger et al. 19F NMR of the copolymer displayed a
resonance at 71.4 ppm, which is assigned to in-chain CF3groups.
Given the similarity between this copolymer and the herein
investigated ETTFMAA oligomers, the observed chemical shift can
be assigned to groups AD. See: Lanzinger, D.; Giuman, M. M.;
Anselment, T. M. J.; Rieger, B. ACS Macro Lett. 2014,3, 931.
ACS Macro Letters Letter
DOI: 10.1021/acsmacrolett.5b00562
ACS Macro Lett. 2015, 4, 933937
937
... This direct method of incorporating functional groups during polymerization is considered as the most important and powerful technique for the preparation of functionalized polyolethylene [23][24][25][26][27][28]. Among the reported catalysts capable of incorporating functional olefins with ethylene, palladium phosphine-sulfonate system appears to be the most successful and has demonstrated broad functional group tolerance towards vinyl ether [29], vinyl acetate [30], acrylate [31], vinyl fluoride [32], acrylonitrile [33], vinyl chloride [34], acrylic acid [35] and 1,1-disubstituted difunctional olefins [36,37] (Fig. 1, top). In addition to vinylic functional olefins, insertion copolymerization of allylic monomers with ethylene using Pd/Ni catalysts has been also reported [38][39][40][41][42]. ...
... In addition to vinylic functional olefins, insertion copolymerization of allylic monomers with ethylene using Pd/Ni catalysts has been also reported [38][39][40][41][42]. However, almost all of the above functional olefins are sourced from fossil reserves [23][24][25][26][27][28][29][30][31][32][33][34][35][36][37][38][39][40][41][42]. In addition; the fossil reserves are depleting with a much faster rate and are inadequate to cater to our future demands [43][44][45]. ...
... Although highly desirable, to the best of our knowledge, there is hardly any report on insertion copolymerization of ethylene with renewable resource derived functional olefins [14]. Fig. 1 best describes the past scenario and the current work [29][30][31][32][33][34][35][36][37]. ...
... Recently, the transition-metal-catalyzed copolymerization of ethylene with 1,1-disubstituted ethylenes bearing two polar functional groups has been reported [28]. However, mechanistic studies indicated the formation of blocky microstructures through successive 1,1-disubstituted ethylene incorporation via radical or ionic mechanisms [29]. ...
Article
The transition-metal-catalyzed copolymerization of olefins with polar comonomers is a direct strategy to access polar-functionalized polyolefins in an economical manner. Due to the intrinsic poisoning effect of polar groups towards Lewis acidic metal centers and the drastic reactivity differences of polar comonomers versus non-polar olefins, it is challenging to develop catalysts that provide the desired polymer molecular weight, comonomer incorporation, and activity. In this contribution, we tackle this issue from a comonomer perspective using 5,6-disubstituted norbornenes, which are highly versatile, easily accessible, inexpensive, and capable of introducing two functional groups in a single insertion. More importantly, they are only mildly poisoning due to the presence of long spacers between double bonds and polar groups, and are not prone to β-hydride elimination due to their cyclic structures. As strong π-donors, they can competitively bind to metal centers versus olefins. Indeed, phosphine-sulfonate palladium catalysts can catalyze the copolymerization of ethylene with 5,6-disubstituted norbornenes and simultaneously achieve a high polymerization activity, copolymer molecular weight, and comonomer incorporation. The practicality of this system was demonstrated by studying the properties of the resulting polymers, copolymerization in hydrocarbon solvents or in bulk, recovery/utilization of unreacted comonomer, molecular weight modulation, and large-scale synthesis.
... Catalytic olefin (co)polymerization has been a pivotal area of research since the foundation of most commercialized coordination/insertion polymerization of olefins using a combination of trace impurities of nickel salts with alkylaluminum compounds as discovered by Karl Ziegler and Giulio Natta [1]. ...
Article
Full-text available
Commercially, polyethylene is a highly produced commodity polymer. But due to its hydrophobic nature, it limits its applications, where utility is based on surface properties, such as adhesion, printability, wettability, and miscibility with other polar polymers. Therefore, the presence of a hydrophilic polar functionality is desired to overcome such difficulties. With its high exigency, synthesis of functionalized polyethylene with unique surface properties is now a very challenging task to be accomplished. In this perspective, developments on palladium and nickel (mainly based on ligands containing neutral α-diimine and anionic phosphine–sulfonate derivative)-mediated coordination/insertion copolymerization of ethylene with polar functionalized co-monomers are discussed herein.
... There is only one report on metal-catalyzed copolymerization difunctional olefin with ethylene. 3 If these simple polar functionalities can pose such a huge problem, even larger synthetic challenges are anticipated if more complex functionalities such as energy-or electron-transporting functionalities have to be incorporated. Thus, a polymerization method that can tolerate a wide range of functional groups and incorporate them in a controllable amount and at designated positions will be highly sought. ...
Article
Granular hydroxyl-functionalized UHMWPE was successfully prepared through copolymerization of ethylene and 10-undecen-1-ol protected by tri-iso-butylaluminum using a titanium complex. [tBuNSiMe2(2,7-tBu2Flu)]TiMe2 was activated in hexane by silica-supported modified-methylaluminoxane. Without any additional pretreatments, the obtained polymer powders were transformed into fibers through a gel-spinning and heat drawing process. When compared to fibers generated from the equivalent unfunctionalized UHMWPE or the commercial HUMWPE, both creep resistance and hydrophilic properties were improved in the hydroxy-functionalized UHMWPE fibers without losing tensile strength.
Article
Full-text available
Scale-up of Fischer-Tropsch (F-T) synthesis using microreactors is very important for a paradigm shift in the production of fuels and chemicals. The scalability of microreactors for F-T Synthesis was experimentally evaluated using 3D printed stainless steel microreactors, containing seven microchannels of dimensions 1000 µm × 1000 µm × 5cms. Mesoporous silica (KIT-6), with high surface area, containing ordered mesoporous structure was used to incorporate 10% cobalt and 5% ruthenium using a one-pot hydrothermal method. Bimetallic Co-Ru-KIT-6 catalyst was used for scale-up of F-T Synthesis. The performance of the catalysts was evaluated and examined for three different scale-up configurations (stand-alone, two, and four microreactors assembled in parallel) at both atmospheric pressure and 20 bar at F-T operating temperature of 240 ˚C using a syngas molar ratio (H2:CO) of 2. All three configurations of microreactors yielded not only comparable CO conversion (85.6% to 88.4%) and methane selectivity (~14%) but also similar selectivity towards lower gaseous hydrocarbons like ethane, propane, and butane (6.23% to 9.4%) observed in atmospheric F-T Synthesis. The overall selectivity to higher hydrocarbons, C5+ is in the range of 75% to 82% at 20 bars. A CFD model was used to investigate the effect of different design features and numbering up approaches on the performance of the microchannel reactor. The effect of the reactor inlet, the mixing internals and the channel designs on the dead zone %, the quality index factor, the cooling requirement and the maximum dimensionless temperature within the microreactor were quantified. There is no significant effect of increasing the channel width on the microreactor performance and operation of the microchannel reactor at lower Nusselt number that results in higher CO conversion. Increasing the channel width reduced the maximum temperature exhibited in the channel. Finally, the effect of increasing the y/x stacking ratio, i.e. having more reactor units in parallel compared to series, was investigated. Increasing the y/x ratio increased the cooling requirement and the maximum dimensionless temperature increase within the unit will decrease the productivity. To minimize the productivity losses, numbering up in series is the better approach; however further analysis must be done to delineate heat removal requirements.
Chapter
Fossil fuel combustion is often considered as one of the major threats to the environment, because of the carbon dioxide (CO2) release in the atmosphere. Such an accumulation of CO2 in the atmosphere leads to drastic climate change in the environment. The control in the discharge of CO2 into the atmosphere and the effective utilization of CO2 are great global challenges behind us. The recent research works show there are reasonable technologies developed on the CO2 capture, and utilization leaves us to relieve little. The recent progresses in the organometallic chemistry and catalysis afford the effective chemical transformation of CO2 and its incorporation into synthetic organic molecules under mild reaction conditions. The catalytic conversion of CO2 into small and beneficial molecules such as carbonates, methylamines, methanol, formic acid, etc., by molecular catalysts, is an interesting topic that has significantly developed in recent years. The aim of this chapter is to reveal the recent advancement in the CO2 capture and its utilization in the synthesis of commodity chemicals. In addition, this also converses various homogenous metal complexes, catalyzed fine chemicals synthesis, and their challenges.
Chapter
Full-text available
In recent years, environmental challenges have led to a focus on the production of clean synthetic fuels from different carbon sources using Fischer-Tropsch (FT) synthesis. Catalyst development and reactor improvements are the major points of interest to obtain high selectivity toward desired hydrocarbons in FT synthesis. The first part of this chapter summarizes the fundamentals of FT synthesis, catalysts, and possible reaction mechanisms, the drawbacks of present synthesis reactors, and how microchannel microreactor (specified as microreactor in this chapter) technology addresses them with its unique characteristics. Two case studies are presented to describe catalyst screening for FT synthesis in two types of microreactors: Silicon (Si) microreactors are fabricated using conventional microfabrication techniques with dimensions 1.6 cm × 50 μm × 100 μm. Stainless steel (SS) 3D printed microreactors of dimensions 2.4 cm × 500 μm × 500 μm are fabricated by direct metal laser sintering method. The FT studies with Si and SS microreactors coated with different catalysts/supports and temperature-programmed reduction (TPR) experiments with H2 not only provide insight into metal-support interactions but also catalyst performance in terms of kinetics, selectivity, CO conversion, and stability. Conversion of syngas enriched with CO2 and CO2 utilization in FT synthesis are the key factors in the production of next-generation biofuels. A case study on the effect of silica and alumina promoters on Co-Fe-K precipitated catalysts in a lab-scale reactor to enhance CO2 utilization in FT synthesis is also included.
Book
This book is part of a two-volume work that offers a unique blend of information on realistic evaluations of catalyst-based synthesis processes using green chemistry principles and the environmental sustainability applications of such processes for biomass conversion, refining, and petrochemical production. The volumes provide a comprehensive resource of state-of-the-art technologies and green chemistry methodologies from researchers, academics, and chemical and manufacturing industrial scientists. The work will be of interest to professors, researchers, and practitioners in clean energy catalysis, green chemistry, chemical engineering and manufacturing, and environmental sustainability. This volume focuses on catalyst synthesis and green chemistry applications for petrochemical and refining processes. While most books on the subject focus on catalyst use for conventional crude, fuel-oriented refineries, this book emphasizes recent transitions to petrochemical refineries with the goal of evaluating how green chemistry applications can produce clean energy through petrochemical industrial means. The majority of the chapters are contributed by industrial researchers and technicians and address various petrochemical processes, including hydrotreating, hydrocracking, flue gas treatment and isomerization catalysts.
Article
Full-text available
A reaction between sodium 2-formylbenzenesulfonate and aniline revealed the near-quantitative (91%) formation of sodium-2-((phenylimino)methyl)benzenesulfonate L1. The identity of L1 was unambiguously ascertained using spectroscopic and analytical methods. The scope of this methodology was widened and various electron-donating amines were treated with sodium 2-formylbenzenesulfonate, and a small library of 6 imine ligands L2–L6 was generated. When L2 was treated with [(COD)PdMeCl], instead of the anticipated [L2PdMe(DMSO)] complex, the formation of [(DMSO)2Pd2Cl2Me2] Pd-Dim was observed. Nevertheless, the desired imino-methyl benzenesulfonate-ligated palladium complex [L2PdMe(Lu)] C1 was obtained by in situ abstraction of chloride and addition of bulky 2,6-lutidine as the donor group. The observation of characteristic Pd–Me protons at 0.06 ppm and the corresponding carbon at −8.1 ppm indicated the formation of C1. These 1D NMR observations were corroborated by 2D C–H correlation spectra and mass analysis, and the existence of C1 was unambiguously ascertained. Along the same lines, L4 and L5 were treated with a palladium precursor to produce [L4/5PdMe(Lu)]-type complexes C2–C3 in 55–84% yield, and their identity was established by using a combination of spectroscopic tools, analytical methods, and single-crystal X-ray diffraction. The synthetic utility of C1–C3 has been demonstrated by utilizing these complexes in the insertion polymerization of ethylene to polyethylene.
Article
Polyolefins are macromolecular alkanes and include the most familiar and most commercially produced plastic, polyethylene. The low cost of these materials combined with their diverse and desirable property profiles drive such large-scale production. One property that renders polyolefins so attractive is their resistance to harsh chemical environments. However, this attribute becomes a severe limitation when attempting to chemically convert these plastics into value-added materials. Functionalization of polymers is a useful methodology for the generation of new materials with wide ranging applications, and this tutorial review describes both new and established methods for the post-polymerization modification of polyolefins.
  • A Kermagoret
  • A Debuigne
  • C Jerome
  • C Detrembleur
Kermagoret, A.; Debuigne, A.; Jerome, C.; Detrembleur, C. Nat. Chem. 2014, 6, 179 and the references therein..
Ziegler-Natta Catalysts and Polymerization
  • J Bhoor
  • Jr
Bhoor, J., Jr. Ziegler-Natta Catalysts and Polymerization; Academic Press: New York, 1979.
Functionalisation of polyolefins
  • T C Chung
Chung, T. C. Functionalisation of polyolefins; Academic Press: London, 2002.
  • N M G Franssen
  • J N H Reek
  • B De Bruin
Franssen, N. M. G.; Reek, J. N. H.; De Bruin, B. Chem. Soc. Rev. 2013, 42, 5809. (9) The topic has been summarized in various articles, see:
  • G W Coates
  • L S Boffa
  • B M Novak
  • A Nakamura
  • S Ito
  • K Nozaki
  • S R Gaikwad
  • S S Deshmukh
  • S H Chikkali
Coates, G. W. Chem. Rev. 2000, 100, 1223. (b) Boffa, L. S.; Novak, B. M. Chem. Rev. 2000, 100, 1479. (c) Nakamura, A.; Ito, S.; Nozaki, K. Chem. Rev. 2009, 109, 5215. (d) Gaikwad, S. R.; Deshmukh, S. S.; Chikkali, S. H. J. Polym. Sci., Part A: Polym. Chem. 2014, 52, 1. (e) Carrow, B. P.; Nozaki, K. Macromolecules 2014, 47, 2541.
  • D Guironnet
  • P Roesle
  • T Runzi
  • I Gottker-Schnetmann
  • S Mecking
  • T Kochi
  • S Noda
  • K Yoshimura
  • K J Nozaki
  • Am
  • Chem
Guironnet, D.; Roesle, P.; Runzi, T.; Gottker-Schnetmann, I.; Mecking, S. J. Am. Chem. Soc. 2009, 131, 422. (14) (a) Kochi, T.; Noda, S.; Yoshimura, K.; Nozaki, K. J. Am. Chem.
  • S Ito
  • K Munakata
  • A Nakamura
  • K Nozaki
Ito, S.; Munakata, K.; Nakamura, A.; Nozaki, K. J. Am. Chem. Soc. 2009, 131, 14606.
  • S Luo
  • J Vela
  • G R Lief
  • R F Jordan
Luo, S.; Vela, J.; Lief, G. R.; Jordan, R. F. J. Am. Chem. Soc. 2007, 129, 8946.
For structural modification of the phosphine−sulfonate ligand system see: (a) Carrow, B. P.; Nozaki, K
  • T Runzi
  • D Frohlich
  • S J Mecking
  • Am
  • H Leicht
  • I Gottker-Schnetmann
  • S Mecking
  • L Piche
  • J C Daigle
  • G Rehse
  • J P Claverie
  • Chem
Runzi, T.; Frohlich, D.; Mecking, S. J. Am. Chem. Soc. 2010, 132, 17690. (18) Low incorporation, mostly only at the chain end, see: Leicht, H.; Gottker-Schnetmann, I.; Mecking, S. Angew. Chem., Int. Ed. 2013, 52, 3963. (19) For structural modification of the phosphine−sulfonate ligand system see: (a) Carrow, B. P.; Nozaki, K. J. Am. Chem. Soc. 2012, 134, 8802. (b) Piche, L.; Daigle, J. C.; Rehse, G.; Claverie, J. P. Chem. -Eur.