Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

Subscriber access provided by Nottingham Trent University

Article
A Comprehensive Understanding of Polyester Stereocomplexation
Zhao-Qian Wan, Julie M. Longo, Li-Xin Liang, Hong-Yu Chen, Guang-Jin
Hou, Shuai Yang, Wei-Ping Zhang, Geoffrey W. Coates, and Xiao-Bing Lu
J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.9b07058 • Publication Date (Web): 28 Aug 2019
Downloaded from pubs.acs.org on August 30, 2019

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a service to the research community to expedite the dissemination
of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in
full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully
peer reviewed, but should not be considered the official version of record. They are citable by the
Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore,
the “Just Accepted” Web site may not include all articles that will be published in the journal. After
a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web
site and published as an ASAP article. Note that technical editing may introduce minor changes
to the manuscript text and/or graphics which could affect content, and all legal disclaimers and
ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or
consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W.,


Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the
course of their duties.
Page 1 of 25 Journal of the American Chemical Society

1
2
3
4 A Comprehensive Understanding of Polyester Stereocomplexation
5
6
7
8
9 Zhao-Qian Wan,† Julie M. Longo,‡ Li-Xin Liang,§ Hong-Yu Chen,§ Guang-Jin Hou,§ Shuai
10
11
12 Yang,† Wei-Ping Zhang,† Geoffrey W. Coates*,‡ and Xiao-Bing Lu*,†
13
14 † State Key Laboratory of Fine Chemicals, Dalian University of Technology, Dalian 116024,
15
16
17 China
18
19 ‡ Department
20 of Chemistry and Chemical Biology, Baker Laboratory, Cornell University, Ithaca,
21
22 New York 14853-1301, United States
23
24
§State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of
25
26
27
Sciences, Dalian 116023, China
28
29
30
31
32
33 ABSTRACT:
34
35 We report a comprehensive understanding of the stereoselective interaction between
36
37
38 two opposite enantiomeric polyesters prepared from the regioselective
39
40
copolymerization of chiral terminal epoxides and cyclic anhydrides. For many of the
41
42
43 resultant polyesters, the interactions between polymer chains of opposite chirality are
44
45
46 stronger than those of polymer chains with the same chirality, resulting in the
47
48 formation of a stereocomplex with enhanced melting point (Tm) and crystallinity. The
49
50
51 backbone, tacticity, steric hindrance of the pendant group, and molecular weight of
52
53
the polyesters have significant effects on stereocomplex formation. Bulky substituent
54
55
56 groups favor stereocomplexation, resulting in a greater rise in Tm in comparison with
57
58
59 the component enantiomeric polymers. Stereocomplex assembly of discrete (R)- and
60
1

ACS Paragon Plus Environment


Journal of the American Chemical Society Page 2 of 25

1
2
3
4 (S)-poly(phenyl glycidyl ether-alt-phthalic anhydride)s oligomers revealed that the
5
6
7 minimum degree of polymerization required for stereocomplex formation is five.
8
9 Raman spectroscopy and solid-state NMR studies indicate that stereocomplex
10
11
12 formation significantly restricts the local mobilities of C=O and C–H groups along the
13
14
backbone of chains. The reduced mobility results in the enhanced spin-lattice
15
16
17 relaxation time and both 1H and 13C downfield shifts due to the strong intermolecular
18
19
20 interactions between (R)- and (S)-chains.
21
22
23
24
25 INTRODUCTION
26
27
Chiral and steric recognitions by noncovalent interactions play critical roles in
28
29
30 selectively spontaneous assembly of molecules into stable and well-defined
31
32
33 aggregates with specific functions. For example, the intermolecular reading of genetic
34
35 code widely observed in biological systems.1-3 In the realm of macromolecules, there
36
37
38 are few enantiopure polymers that form stereocomplexes through an interlocked
39
40
orderly assembly between two opposite enantiomeric chains when mixed in
41
42
43 equivalent amounts.4-11 In comparison with the parent polymers, the stereocomplex
44
45
46 usually exhibits improved physical properties, such as higher levels of crystallinity
47
48 and melting points (Tm). One of the most widely studied examples is poly(lactic acid)
49
50
51 (PLA).4, 12-13 Stereocomplexed PLA possesses a Tm of 230 °C, which is about 50 °C
52
53
higher than that of its parent polymers, enantiopure poly(L-lactic acid) (PLLA) and
54
55
56 poly(D-lactic acid) (PDLA). Additionally, the stereocomplexation of PLLA and
57
58
59 PDLA significantly increases its hydrolytic/thermal degradation-resistance and gas
60
2

ACS Paragon Plus Environment


Page 3 of 25 Journal of the American Chemical Society

1
2
3
4 barrier properties.9
5
6
7 Epoxides have been widely studied in alternating copolymerizations with carbon
8
9 dioxide to produce polycarbonates and with cyclic anhydrides to produce
10
11
12 polyesters.14-19 The stereogenic centers in the substituted epoxides allow the
13
14
possibility of forming tactic polymers by stereospecific copolymerization.20 In recent
15
16
17 years, a variety of stereoregular polycarbonates and polyesters were prepared by
18
19
20 means of regio- and/or stereoselective ring-opening epoxide copolymerization.21-24
21
22 Most isotactic polycarbonates from CO2 and meso-epoxides are semicrystalline
23
24
25 materials, and their crystallinities depend on tacticity. Notably, blending equivalent
26
27
amounts of enantiomeric polycarbonates derived from meso-epoxides readily formed
28
29
30 crystalline stereocomplexes.25 On the contrary, cocrystallization was not observed in a
31
32
33 1:1 mixture of highly isotactic (R)- and (S)-polycarbonates from terminal epoxides,
34
35 except for styrene oxide derivatives.26 Interestingly, (R)- and (S)-poly(propylene
36
37
38 succinate), which each have a very low melting polymorph with a Tm around 70 °C,
39
40
formed a stereocomplex with improved crystallinity and an increased Tm of 120 °C.27
41
42
43 The half-life of recrystallization for the stereocomplex is approximately 3 orders of
44
45
46 magnitude faster than that of the parent enantiopure polymers. However, the critical
47
48 factors affecting polymer stereocomplexation and the relationship between molecular
49
50
51 weight and stereocomplex formation were not investigated.
52
53
Herein, we present a comprehensive study regarding the effects of the backbone
54
55
56 structure, tacticity, steric bulk of the pendant group, and molecular weight of the
57
58
59 polyesters derived from terminal epoxides on stereocomplex formation. Furthermore,
60
3

ACS Paragon Plus Environment


Journal of the American Chemical Society Page 4 of 25

1
2
3
4 Raman spectroscopy and solid state NMR were used to investigate the nature of the
5
6
7 intermolecular interactions in a stereocomplexed polyester, poly(phenyl glycidyl
8
9 ether-alt-phthalic anhydride) (poly(PGE-alt-PA)).
10
11
12 RESULTS AND DISCUSSION
13
14
Synthesis of Enantiomeric (R)- and (S)-polyesters for Screening Stereocomplex
15
16
17 Formation.
18
19
20 Various enantiopure polyesters (Scheme 1) were prepared by the regioselective
21
22 copolymerization of a series of chiral terminal epoxides and cyclic anhydrides
23
24
25 catalyzed by the binary catalyst consisting of [PPN][NO3] and fluorine-substituted
26
27
salcyCo(ІІІ)NO3 complex, which is a highly active, regioselective Co(ІІІ) species
28
29
30 with longer catalyst life time in the copolymerization of epoxide with anhydride 28
31
32
33 (salcy = N,N’-bis(salicylidene)cyclohexanediamine, PPN =
34
35 bis(triphenylphosphine)iminium). The number average molecular weights (Mn) of
36
37
38 these isotactic polyesters ranged from 7.2 to 23.0 kDa, depending on the combination
39
40
of epoxide and cyclic anhydride. All resultant polyesters exhibited high levels of
41
42
43 enantioselectivity (≥98% (S)- or (R)-stereocenter) and regioregularity (≥98%
44
45
46 head-to-tail linkages) due to selective ring-opening at the less sterically hindered
47
48 methylene carbon of the epoxides. (see Supporting Information, Table S1) After
49
50
51 precipitation in excess methanol, most of the isolated isotactic polyesters were
52
53
semi-crystalline, although some were amorphous. However, upon soaking in
54
55
56 methanol, these amorphous polyesters gradually transformed into semi-crystalline
57
58
59 materials. This might be ascribed to their slow crystallization rates. Therefore, the
60
4

ACS Paragon Plus Environment


Page 5 of 25 Journal of the American Chemical Society

1
2
3
4 isotactic polyesters used in this study were immersed in methanol overnight after
5
6
7 precipitation to improve crystallization.
8
9 Scheme 1. Regioselective Copolymerization of Chiral Terminal Epoxides and Cyclic Anhydride
10
11
12 Using Chiral SalcyCoNO3 and [PPN][NO3]
13
14
15 SalcyCoNO3
O O O
16 R2 [PPN][NO3] O O
17 O N N
+ O
30 oC Co
18 R1 R2 R1 R2 R2 n
F O O F
19 O NO3
t t
20 Bu Bu
O O O O O SalcyCoNO3
21 O
O
22 Ph Bn Ph
O O O
23 PO TBO PGE BGE EPB
24 O O O
O O O O
25 O O O O O O O
26
PA MA SA
27 TBGE IGE EGE MGE
28
29
30
31
32
33 The copolymers of phthalic anhydride (PA) and enantiopure glycidyl ether
34
35 derivatives are all semi-crystalline with Tms between 90 and 150 °C (see Figure 1 for
36
37
38 named polymer structures). The copolymer (S)-poly(PO-alt-PA) shows a Tm of 142
39
40
°C. When (S)-EPB) or (S)-TBO) was copolymerized with PA, the corresponding
41
42
43 (S)-poly(EPB-alt-PA) or (S)-poly(TBO-alt-PA) is amorphous with a glass transition
44
45
46 temperature (Tg) of 59 °C or 94 °C, respectively. The copolymer of (S)-PGE and
47
48 maleic anhydride (MA) ((R)-poly(PGE-alt-MA)) is a semi-crystalline polymer with a
49
50
51 Tm of 123 °C. While both isotactic poly(PGE-alt-PA) and isotactic poly(PGE-alt-MA)
52
53
are semicrystalline polymers, isotactic poly(PGE-alt-SA) is amorphous with a Tg of
54
55
56 23 °C, despite its high level (99%) of isotacticity.
57
58
59 To test the stereocomplex formation, 50 mg (S)- and (R)-polyesters were dissolved
60
5

ACS Paragon Plus Environment


Journal of the American Chemical Society Page 6 of 25

1
2
3
4 in 3 mL dichloromethane respectively, and the two solutions mixed. The mixed
5
6
7 solution was added dropwise to an excess of methanol to precipitate the polymer.
8
9 After drying, the resulting blend was analyzed by differential scanning calorimetry
10
11
12 (DSC) and wide-angle X-ray diffraction (WAXD). The melting point and crystalline
13
14
diffraction peaks of each blend were compared with those of the parent polyesters to
15
16
17 confirm stereocomplex formation.
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44 Figure 1. Equivalent isotactic (R)- and (S)-polyester blends used for screening stereocomplex
45
46
47 formation by DSC and WAXD analysis (ΔTm = Tm(blend)-Tm(iso)).
48
49
50
51
52 No stereocomplexation was observed when equal masses of (S)- and
53
54
55 (R)-poly(PO-alt-PA) were blended (see Supporting Information, Figures S15 and S16)
56
57 even though enantiomeric mixtures of structurally similar poly(PO-alt-SA) form
58
59
60 stereocomplexes. In contrast, the blending of (S)- and (R)-poly(PGE-alt-PA) formed a
6

ACS Paragon Plus Environment


Page 7 of 25 Journal of the American Chemical Society

1
2
3
4 stereocomplex with a Tm of 184 °C, which is 60 °C higher than that of its parent
5
6
7 polymers (Figure 2a). While no Tg was detected in the stereocomplexed
8
9 poly(PGE-alt-PA), a Tg of 68 °C was observed in (S)-poly(PGE-alt-PA) (Figure 2a).
10
11
12 This implies that the stereocomplexed poly(PGE-alt-PA) has higher levels of
13
14
crystallinity than the component enantiomeric polymers. WAXD further confirmed
15
16
17 stereocomplex formation. The stereocomplexed poly(PGE-alt-PA) has four major
18
19
20 crystalline diffraction peaks at 6.1°, 20.2°, 22.2°, and 25.2°. In contrast,
21
22 (S)-poly(PGE-alt-PA) has three strong peaks at 11.5°, 17.2°, and 19.0° (Figure 2b).
23
24
25
26 stereocomplex
27
stereocomplex
28
29 O O
30 O O
O O
O O n
n
31
O
32 O Ph
Ph
33
34
35
36
37
38
39 Figure 2. a) DSC thermograms and b) WAXD profiles of (S)-poly(PGE-alt-PA) and
40
41
42 stereocomplexed poly(PGE-alt-PA).
43
44
45
46
47 We next sought to further investigate the effect of epoxide side chain structure on
48
49
50 stereocomplex formation. A series of (R)- and (S)- poly(glycidyl ether-alt-PA)
51
52 polyesters was synthesized and the thermal properties of the component enantiomeric
53
54
55 polyesters and corresponding blends investigated by DSC (Figure 1). Blending (R)-
56
57 and (S)-poly(BGE-alt-PA) in 1:1 ratio also resulted in stereocomplex formation (see
58
59
60 Supporting Information, Figures S17 and S18). Stereocomplexed poly(BGE-alt-PA)
7

ACS Paragon Plus Environment


Journal of the American Chemical Society Page 8 of 25

1
2
3
4 shows a Tm of 150 °C, which is 56 °C higher than that of its parent polyesters. To
5
6
7 verify whether the aromatic group in the pendant chains is essential for
8
9 stereocomplexation, the blending of enantiomeric polyesters without aromatic groups
10
11
12 in the side chains was performed. Blending (R)- and (S)-poly(TBGE-alt-PA) also
13
14
resulted in stereocomplex formation, as confirmed by DSC and WAXD analysis (see
15
16
17 Supporting Information, Figures S19 and S20). The DSC plot shows that the
18
19
20 stereocomplexed poly(TBGE-alt-PA) has a Tm of 162 °C, 68 °C higher than that of
21
22 the component enantiomeric polyesters. This result illustrates that the aromatic groups
23
24
25 in the pendant chains of enantiomeric polyesters are not essential for
26
27
stereocomplexation.
28
29
30 To test the steric effect of the side chains on stereocomplex formation,
31
32
33 enantiomeric polyesters with different substituent groups on the pendant chain were
34
35 investigated. Blending equal masses of (R)- and (S)-poly(IGE-alt-PA) formed a
36
37
38 stereocomplex with a Tm of 143 °C, 30 °C higher than that of its parent polyesters (see
39
40
Supporting Information, Figures S21 and S22). Blending equal masses of (R)- and
41
42
43 (S)-poly(EGE-alt-PA) also formed a stereocomplex with a Tm of 111 °C, which is
44
45
46 only 2 °C higher than that of the component enantiomeric copolymers (Figure 3a).
47
48 Although there is only a slight difference in Tm, the stereocomplexed
49
50
51 poly(EGE-alt-PA) shows unique crystalline behavior, distinct from its parent
52
53
polymers (Figure 3b). Five diffraction peaks at 8.1°, 12.7°, 19.5°, 21.3°, 24.3°, and
54
55
56 25.4° were found in the WAXD of (S)-poly(EGE-alt-PA), while stereocomplexed
57
58
59 poly(EGE-alt-PA) has five large diffraction peaks at 7.0°, 19.1°,
60
8

ACS Paragon Plus Environment


Page 9 of 25 Journal of the American Chemical Society

1
2
3
4
5
6 stereocomplex
7 stereocomplex
8
9 O O
O O
O O
n
10 O O
n

11 O
O

12
13
14
15
16
17
18 Figure 3. a) DSC thermograms and b) WAXD profiles of (S)-poly(EGE-alt-PA) and the blend of
19
20
21 (R)- and (S)-poly(EGE-alt-PA) in 1:1 ratio.
22
23
24
25
26 20.9°, 22.5°, and 24.4° (Figure 3b). Interestingly, further decreasing the steric
27
28
29 hindrance of the substituent groups in the pendant chain resulted in no
30
31 stereocomplexation. The 1:1 blend of (R)- and (S)-poly(MGE-alt-PA) showed the
32
33
34 same diffraction peaks at 2θ angles as (S)-poly(MGE-alt-PA) (see Supporting
35
36
37 Information, Figures S23 and S24). These results suggest that steric hindrance in the
38
39 pendant chains of enantiomeric polyesters derived from PA and epoxides plays an
40
41
42 important role in stereocomplex formation.
43
44 It should be noted that all the stereocomplexes discussed above are based on
45
46
47 glycidyl ether derivatives. Blending two opposite enantiomeric polyesters from EPB
48
49
50 or TBO also formed stereocomplexes (see Supporting Information, Figures S25, S26,
51
52 S27, and S28), confirmed by DSC and WAXD. This result suggests that an oxygen
53
54
55 atom in the pendant chains of enantiomeric polyesters is not essential to
56
57 stereocomplexation.
58
59
60 To investigate the effect of the backbone structure of enantiomeric polyesters on
9

ACS Paragon Plus Environment


Journal of the American Chemical Society Page 10 of 25

1
2
3
4 stereocomplex formation, enantiomeric poly(PGE-alt-MA) and poly(PGE-alt-SA)
5
6
7 were prepared. No stereocomplexation occurs between the two opposite
8
9 enantiomeric poly(PGE-alt-SA)s, (see Supporting Information, Figures S29 and S30)
10
11
12 while the 1:1 blend of (R)- and (S)-poly(PGE-alt-MA) formed a stereocomplex with
13
14
a Tm of 137 °C, which is 14 °C higher than that of either isotactic polymer. (see
15
16
17 Supporting Information, Figures S31 and S32). As noted above, stereocomplexed
18
19
20 poly(PGE-alt-PA) showed a Tm 60 °C higher than that of isotactic poly(PGE-alt-PA).
21
22 This demonstrates that the backbone rigidty of enantiomeric polyesters also has a
23
24
25 critical influence on stereocomplex formation, and that enantiomeric polyesters based
26
27
on PA and epoxides with bulky substituent groups were more likely to form
28
29
30 stereocomplexes.
31
32
33 The Effect of Tacticity and Mn on Stereocomplexation
34
35 Having discovered a series of crystalline polyesters and their stereocomplexes, we
36
37
38 probed the effect of tacticity on crystallinity and stereocomplex formation, using
39
40
poly(PGE-alt-PA) as a model polyester. Various poly(PGE-alt-PA)s with 60, 74, 82,
41
42
43 and 99% enantiomeric excess (ee) were prepared with Mn values between 12.0 and
44
45
46 14.0 kDa. Isotactically-enriched poly(PGE-alt-PA)s with 60% and 74% ee are
47
48 amorphous materials with Tg values of 63 °C and 70 °C, respectively (Figure 4a). In
49
50
51 contrast, an enhanced Tg of about 80 °C and a very small melting endothermic peak at
52
53
110 °C with a melting enthalpy (ΔH) of 7 J/g was observed in the poly(PGE-alt-PA)
54
55
56 sample with 82% ee, indicating a very low degree of crystallinity (Figure 4a). The
57
58
59 highly isotactic poly(PGE-alt-PA) with 99% ee exhibited a sharp and high
60
10

ACS Paragon Plus Environment


Page 11 of 25 Journal of the American Chemical Society

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18 Figure 4. DSC thermograms of a) isotactically-enriched poly(PGE-alt-PA) with
19
20
21 different ee values; and b) the 1:1 blend of two opposite enantiomeric
22
23
24 poly(PGE-alt-PA)s with different ee values.
25
26
27
28
29 crystallization endothermic peak at about 120 °C (Figure 4a). The influence of the
30
31 tacticity of the parent poly(PGE-alt-PA)s on stereocomplexation is shown in Figure
32
33
34 4b. No stereocomplexation is observed for the 1:1 mixture of the two opposite
35
36
37 configuration poly(PGE-alt-PA)s with 60% ee, while the blending of parent
38
39 copolymers with 74%, 82%, and 99% ee resulted in stereocomplex formation (Figure
40
41
42 4b). The Tm and ΔH values increase with the tacticity of the parent
43
44 poly(PGE-alt-PA)s.
45
46
47 Subsequently, the stereocomplexation of enantiomeric poly(PGE-alt-PA) with
48
49
50 different molecular weights was investigated. To control the molecular weight, benzyl
51
52 alcohol was used as a chain transfer agent. The blending of (R)- and
53
54
55 (S)-poly(PGE-alt-PA) with Mns of 3.6 kDa and dispersities (Ð) of 1.20 resulted in the
56
57 formation of a stereocomplex with a Tm of 165 °C (Figure 5a). Interestingly, the
58
59
60 blend of 1.8 kDa enantiomeric poly(PGE-alt-PA) also formed the stereocomplex,
11

ACS Paragon Plus Environment


Journal of the American Chemical Society Page 12 of 25

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18 Figure 5. a) DSC thermograms and b) WAXD profiles of (S)-poly(PGE-alt-PA)
19
20
21 oligomers with molecular weight of 1.8 kDa and 3.6 kDa, and the 1:1 blend of two
22
23
24 opposite enantiomeric poly(PGE-alt-PA)s with molecular weight of 1.8 kDa and 3.6
25
26 kDa.
27
28
29
30
31 confirmed by WAXD (Figure 5b). However, its Tm is close to that of its parent
32
33
34 polyesters (Figure 5a). This result provides evidence for the critical number of
35
36
37 ordered repeat units for forming this stereocomplex. We suspect the minimum
38
39 molecular weight required for poly(PGE-alt-PA) stereocomplexation is about 1.8 kDa.
40
41
42 Since the molecular weight of the repeat unit of poly(PGE-alt-PA) is 298.1 g/mol, the
43
44 minimum degree of polymerization (DP) for stereocomplex formation must be less
45
46
47 than 6. Inspired by the excellent researches on length-dependent structure property
48
49
50 relationship,29-32 thin-layer chromatography (150:1, chloroform:methanol) was used to
51
52 isolate discrete stereoregular oligomers from the enantiomeric poly(PGE-alt-PA) with
53
54
55 an original Mn of 1.8 kDa and a Ð of 1.21. Enantiomeric oligomers with DP’s of 4 and
56
57 5 were obtained, and confirmed by MALDI-TOF MS (Figure 6), GPC (see
58
59
60 Supporting Information, Figure S40) and 1H NMR (see Supporting Information,
12

ACS Paragon Plus Environment


Page 13 of 25 Journal of the American Chemical Society

1
2
3
4 a) Bn O
O O
O O H
Oligmer b) [M + Na]+Found = 1323
mixture
5 n
Mn = 1.8 kDa
6 O DGPC = 1.21
298.1 a.m.u
[M + K]+Found = 1339
Ph
7 DP = 4
8 chromatography
separation
9 [M + Na]+Found = 1621
10
11 O O O O 298.1 a.m.u
[M + K]+Found = 1637

12 Bn O O O H
4
Bn O O O H
5
13 O
DP = 5
O
14 Ph Ph
DP = 5
15 DP = 4

16
17
18 Figure 6. a) Schematic illustration of the separation of 1.8 kDa
19
20
21 (R)-poly(PGE-alt-PA) ; b) MALDI-TOF mass spectra of the isolated
22
23
24 (R)-poly(PGE-alt-PA) samples (DP = 4, and 5).
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43 Figure 7. a) DSC thermograms and b) WAXD profiles of (S)-poly(PGE-alt-PA) with
44
45
46 a DP of 4 and 5, the 1:1 blend of two opposite enantiomeric poly(PGE-alt-PA)
47
48 oligomers with a DP of 4 and 5.
49
50
51
52
53
Figures S13 and S14) characterization. No stereocomplexation was observed when
54
55
56 (R)- and (S)-poly(PGE-alt-PA) with a DP of 4 were mixed (Figures 7a and 7b).
57
58
59 However, when (R)- and (S)-poly(PGE-alt-PA) oligomers with DP of 5 were blended,
60
13

ACS Paragon Plus Environment


Journal of the American Chemical Society Page 14 of 25

1
2
3
4 stereocomplexation was observed by both DSC and WAXD (Figures 7a and 7b).
5
6
7 While the parent polyesters with DP of 5 were amorphous, the blend was
8
9 semicrystalline with a melting point of 81 °C and a ΔH value of 56 J/g. Therefore, the
10
11
12 minimum DP required for stereocomplexation between (R)- and (S)-poly(PGE-alt-PA)
13
14
is five.
15
16
17
18
19
20 Raman and Solid State NMR Investigation of Polyester Stereocomplexation
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Figure 8. Raman spectra (1650-1830 cm-1) of (A) isotactic and (B) stereocomplexed
35
36
poly(PGE-alt-PA).
37
38
39
40
41
42 To better understand the interaction between the two opposite enantiomeric
43
44 polymers, Raman spectroscopy was used to investigate the differences in vibrational
45
46
47 modes between isotactic and stereocomplexed poly(PGE-alt-PA) in the solid state.33
48
49
50 A
51
52 single peak at 1729.0 cm-1 assigned to the ν(C=O) appears in the isotactic
53
54
55 poly(PGE-alt-PA) (Figure 8, plot A), while in the
56
57 stereocomplexed-poly(PGE-alt-PA), the ν(C=O) signal was split into two major
58
59
60 peaks at 1706.8 cm-1 and 1735.0 cm-1 (Figure 8, plot B). The splitting of the ν(C=O)
14

ACS Paragon Plus Environment


Page 15 of 25 Journal of the American Chemical Society

1
2
3
4 signal implies that two different carbonyl environments exist after stereocomplexation.
5
6
7 The new peak at 1706.8 cm-1 can be ascribed to the ν(C=O) signal generated by the
8
9 stereocomplexation. The 22 cm-1 frequency shift of ν(C=O) after stereocomplexation
10
11
12 indicates that some carbonyl groups may participate in hydrogen bonding as proton
13
14
acceptors.34-35 The Raman shift of the vibration of the possible proton donor, ν(C-H),
15
16
17 is also observed (see Supporting Information, Figure S35).
18
19
20 Furthermore, the interactions between (R)- and (S)- chains in stereocomplexed
21
22 poly(PGE-alt-PA), were investigated using solid state 13 C cross polarization
23
24
25 (CP)/magic angle spinning (MAS) and 1 H- 13 C HETCOR NMR methods. 36-38 In
26
27
comparison to the component enantiomeric polymers, the stereocomplexed
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47 Figure 9. 13C CP/MAS spectra of (A) isotactic and (B) stereocomplexed poly(PGE-alt-PA).
48
49
50
51
52 poly(PGE-alt-PA) has a different solid state 13C CP/MAS spectrum: the carbonyl
53
54
55 carbons resonance signals in isotactic poly(PGE-alt-PA) are at 169.7 and 166.8 ppm
56
57 whereas the same signals in stereocomplexed poly(PGE-alt-PA) shift to 171.1 and
58
59
60 165.9 ppm, respectively (Figure 9). The larger splitting of these resonance signals in
15

ACS Paragon Plus Environment


Journal of the American Chemical Society Page 16 of 25

1
2
3
4 the spectrum of the stereocomplex indicates a bigger difference in the chemical
5
6
7 environment of the carbonyl carbons. Additionally, in the spectrum of isotactic
8
9 poly(PGE-alt-PA), only two peaks at 130.0 and 132.9 ppm can be observed between
10
11
12 124 and 140 ppm. As for the spectrum of stereocomplexed poly(PGE-alt-PA), five
13
14
peaks are observed in this range. In CP/MAS solid state NMR, the group or molecular
15
16
17 motions decrease the cross polarization efficiency, which reduces resolution of the
18
19
20 spectrum. The peaks from 124 to 140 ppm are mainly ascribed to the resonance of
21
22 aromatic carbons in the backbones of polymers. Therefore, the increased spectrum
23
24
25 resolution in the area implies that the motions of aromatic rings in the backbones of
26
27
stereocomplexed poly(PGE-alt-PA) are restricted.
28
29
30
31
32 Table 1. 13C NMR T1 Relaxation Times for Solid (S)- and Stereocomplexed- Poly(PGE-alt-PA)
33 at Ambient Temperature.
34
35 O O
i k a a
36 O
O
c
37 d n
j O
38 b e
39 h
40 f
41 g
42
43 Entry T1 (s)
44 C=O CH CH2
45
46 a1 a2 b c–f g h i j k
47 1a 340 120 75 52 72 19 76 84
48 2b >500 >500 77 52 72 17 11 55
49
50
0
51 a(S)-poly(PGE-alt-PA), bstereocomplexed-poly(PGE-alt-PA).
52
53
54
55
56
57
58 Spin-lattice relaxation time (T1) was measured for each carbon atom in isotactic
59
60
16

ACS Paragon Plus Environment


Page 17 of 25 Journal of the American Chemical Society

1
2
3
4 and stereocomplexed poly(PGE-alt-PA) samples under the cross polarization
5
6
7 condition by the application of the saturation recovery-based sequence. The T1 values
8
9 of the carbonyl peaks for stereocomplexed poly(PGE-alt-PA) are longer than those of
10
11
12 isotactic poly(PGE-alt-PA). For the carbonyl peak with lower chemical shift (165.9
13
14
ppm, Ca2), its T1 is over 500 s, while the T1 value of the corresponding carbonyl for
15
16
17 isotactic poly(PGE-alt-PA) is 120 s. (Table 1) The much larger T1 value of the
18
19
20 carbonyl peak at 165.9 ppm for stereocomplexed poly(PGE-alt-PA) indicates that the
21
22 motion of the carbonyl is restricted after stereocomplexation. Similar T1 values for
23
24
25 carbon atoms of the phenyl group (Cb, Cg, Ch) in the side chain were measured for
26
27
isotactic and stereocomplexed poly(PGE-alt-PA)s. The T1 value of the methylene
28
29
30 carbons (Cj, Ck) for stereocomplexed poly(PGE-alt-PA) (55 s) is shorter than that of
31
32
33 isotactic poly(PGE-alt-PA) (84 s). (Table 1) The T1 value of the methine carbon (Ci)
34
35 for stereocomplexed poly(PGE-alt-PA) (110 s) is longer than that of isotactic
36
37
38 poly(PGE-alt-PA) (76 s), (Table 1) indicating that the motion of the methine is also
39
40
restricted after stereocomplexation.
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57 Figure 10. 1H–13C HETCOR NMR spectra at the methine region of (A) isotactic and (B)
58
59
60 stereocomplexed poly(PGE-alt-PA) at a contact time of 1000 µs.
17

ACS Paragon Plus Environment


Journal of the American Chemical Society Page 18 of 25

1
2
3
4
5
6 1H–13C
7 HETCOR NMR experiments with a contact time of 1000 μs were carried
8
9 out for both isotactic and stereocomplexed poly(PGE-alt-PA) samples, and the spectra
10
11
12 of the methine region are shown in Figure 10. While the chemical shifts of the
13
14
methine hydrogen and carbon in isotactic poly(PGE-alt-PA) are 4.03 and 69.34 ppm,
15
16
17 respectively, the corresponding signals in the stereocomplex are 5.82 and 71.89 ppm,
18
19
20 respectively. The 1H downfield shift of 1.79 ppm and 13C downfield shift of 2.55 ppm
21
22 indicate that the electronic cloud density of the methine hydrogen was weakened after
23
24
25 stereocomplexation. This may be attributed to the electron-withdrawing inductive
26
27
effect of the carbonyl oxygen atom in the polymer chain with opposite configuration.
28
29
30 Taking the Raman spectroscopy and solid state NMR studies into consideration, we
31
32
33 attribute the driving force of the stereocomplexation to the weak hydrogen bond
34
35 interaction of C-H•••O=C between the oxygen atom of a carbonyl group of one
36
37
38 enantiomer and the hydrogen atom on the methine C-H group at the chiral carbon of
39
40
the other stereoisomer.
41
42
43 CONCLUSIONS
44
45
46 We report the synthesis of a variety of highly isotactic polyesters via the
47
48 regioselective ring-opening copolymerization of enantiopure terminal epoxides with
49
50
51 cyclic anhydrides mediated by the fluorine-substituted salcyCoNO3 complex in
52
53
combination with [PPN][NO3]. Most isotactic polyesters are semicrystalline materials,
54
55
56 possessing Tms between 92 and 142 °C. Notably, stereocomplexation was generally
57
58
59 observed in the 1:1 mixture of (R)- and (S)-polyesters, affording the corresponding
60
18

ACS Paragon Plus Environment


Page 19 of 25 Journal of the American Chemical Society

1
2
3
4 stereocomplexes with enhanced Tm and higher levels of crystallinity, as well as
5
6
7 distinct crystallization behavior from its parent polymers. The steric hindrance of the
8
9 pendant chain, backbone rigidity, tacticity, and molecular weight have significant
10
11
12 effects on stereocomplex formation. Bulky epoxide substituent groups appear to
13
14
benefit stereocomplexation with higher increases in Tm in comparison with the
15
16
17 component enantiomeric polymers. The study on the stereoselective interaction
18
19
20 between two opposite enantiomeric, discrete oligomers suggests that the minimum
21
22 degree of polymerization for the stereocomplex formation between enantiomeric
23
24
25 poly(PGE-alt-PA)s is 5.
26
27
Raman spectroscopy study confirmed the obvious difference at the ν(C=O) signal.
28
29
30 A single peak at 1729.0 cm-1 appears in the isotactic poly(PGE-alt-PA), while in the
31
32
33 stereocomplex, the ν(C=O) signal was split into two major peaks at 1706.8 cm-1 and
34
35 1735.0 cm-1. Solid state 13C cross polarization/magic angle spinning and 1H-13C
36
37
38 HETCOR NMR analysis demonstrated that strong intermolecular interactions
39
40
between stereocomplexed R- and S-chains significantly restricts the local mobilities of
41
42
43 C=O and C–H groups along the backbone of chains and results in the enhanced
44
45
46 spin-lattice relaxation time and both 1H and 13C downfield shifts. These studies
47
48 suggest the driving force of stereocomplexation is the weak hydrogen-bonding
49
50
51 interaction between carbonyl and methine of the opposite enantiomeric polyesters.
52
53
54
55
56 ASSOCIATED CONTENT
57
58
59 Supplementary characterization data including 1H NMR and 13C NMR spectra, DSC
60
19

ACS Paragon Plus Environment


Journal of the American Chemical Society Page 20 of 25

1
2
3
4 and XRD data, as well as GPC of various polyesters.
5
6
7
8
9
10 AUTHOR INFORMATION
11
12
13 Corresponding Author
14
15
16 *E-mail: xblu@dlut.edu.cn;
17
18 coates@cornell.edu
19
20
21 Notes
22
23
24
The authors declare no competing financial interests.
25
26
27
28 ACKNOWLEDGMENT
29
30
31 This work is supported by the National Natural Science Foundation of China (NSFC,
32
33 Grant 21690073), and Program for Changjiang Scholars and Innovative Research
34
35
36 Team in University (IRT-17R14), and the Center for Sustainable Polymers, an NSF
37
38
39 Center for Chemical Innovation (CHE-1413862).
40
41
42
43
44 REFERENCES
45
46 (1) Nakano,T.; Okamoto, Y. Synthetic helical polymers: Conformation and
47
48 function. Chem. Rev. 2001, 101, 4013–4038.
49
50 (2) Sawaya, M. R.; Sambashivan, S.; Nelson, R.; Ivanova, M. I.; Sievers, S. A.;
51
52 Apostol, M. I.; Thompson, M. J.; Balbirnie, M.; Wiltzius, J. J. W.; Mcfarlane, H. T.;
53
54 Madsen, A. Ø.; Riekel, C.; Eisenberg, D. Atomic structures of amyloid cross-beta
55
56 spines reveal varied steric zippers. Nature 2007, 447, 453–457.
57
58 (3) Uhlenheuer, D. A.; Petkau, K.; Brunsveld, L., Combining supramolecular
59
60 chemistry with biology. Chem. Soc. Rev. 2010, 39, 2817-2826.
20

ACS Paragon Plus Environment


Page 21 of 25 Journal of the American Chemical Society

1
2
3
4 (4) Ikada, Y.; Jamshidi, K.; Tsuji, H.; Hyon, S. H., Stereocomplex Formation
5
6 between Enantiomeric Poly(Lactides). Macromolecules 1987, 20, 904-906.
7
8 (5) Jiang, Z. Z.; Boyer, M. T.; Sen, A., Chiral and Steric Recognition in
9
10 Optically-Active, Isotactic, Alternating Alpha-Olefin-Carbon Monoxide Copolymers -
11
12
Effect on Physical-Properties and Chemical-Reactivity. J. Am. Chem. Soc. 1995, 117,
13
14
7037-7038.
15
(6) Wu, G.-P.; Jiang, S.-D.; Lu, X.-B.; Rena, W.-M.; Yan, S.-K., Stereoregular
16
17
poly(cyclohexene carbonate)s: Unique crystallization behavior. Chin. J. Polym. Sci.
18
19 2012, 30, 487-492.
20
21 (7) Auriemma, F.; De Rosa, C.; Di Caprio, M. R.; Di Girolamo, R.; Ellis, W. C.;
22
23 Coates, G. W., Stereocomplexed Poly(Limonene Carbonate): A Unique Example of
24
25 the Cocrystallization of Amorphous Enantiomeric Polymers. Angew. Chem. Int. Ed.
26
27 2015, 54, 1215-1218.
28
29 (8) Ren, J. M.; Lawrence, J.; Knight, A. S.; Abdilla, A.; Zerdan, R. B.; Levi, A. E.;
30
31 Oschmann, B.; Gutekunst, W. R.; Lee, S.-H.; Li, Y.; McGrath, A. J.; Bates, C. M.;
32
33 Qiao, G. G.; Hawker C. J. Controlled Formation and Binding Selectivity of Discrete
34
35 Oligo(methyl methacrylate) Stereocomplexes. J. Am. Chem. Soc. 2018, 140,
36
37 1945−1951.
38
39 (9) Bertin, A., Emergence of Polymer Stereocomplexes for Biomedical
40
41 Applications. Macromol. Chem. Phys. 2012, 213, 2329-2352.
42
43 (10) Zhu, J.-B.; Watson, E. M.; Tang, J.; Chen, E. Y. X. A synthetic polymer
44
45 system with repeatable chemical recyclability. Science 2018, 360, 398−403.
46
47 (11) Tsuji, H.; Noda, S.; Kimura, T.; Sobue, T.; Arakawa, Y. Configurational
48
49 molecular glue: one optically active polymer attracts two oppositely configured
50
51
optically active polymers. Sci. Rep. 2017, 7, 45170.
52
53
(12) Tsuji, H.; Yamasaki, M.; Arakawa, Y. Stereocomplex Formation between
54
Enantiomeric Alternating Lactic Acid-Based Copolymers as a Versatile Method for
55
56
the Preparation of High Performance Biobased Biodegradable Materials. ACS Appl.
57
58 Polym. Mater. 2019, 16, 1476-1484.
59
60 (13) Marin, P.; Tschan, M. J‐L.; Isnard, F.; Robert, C.; Haquette, P.; Trivelli, X.;
21

ACS Paragon Plus Environment


Journal of the American Chemical Society Page 22 of 25

1
2
3
4 Chamoreau, L.‐M.; Guérineau, V.; del Rosal, I.; Maron, L. Polymerization of
5
6 rac‐Lactide Using Achiral Iron Complexes: Access to Thermally Stable
7
8 Stereocomplexes. Angew. Chem. Int. Ed. 2019, DOI: 10.1002/anie.201903224.
9
10 (14) Darensbourg, D. J. Making Plastics from Carbon Dioxide: Salen Metal
11
12
Complexes as Catalysts for the Production of Polycarbonates from Epoxides and CO2.
13
14
Chem. Rev. 2007, 107, 2388−2410.
15
(15) Klaus, S.; Lehenmeier, M. W.; Anderson, C. E.; Rieger, B. Recent Advances in
16
17
CO2/Epoxide Copolymerization: New strategies and Cooperative Mechanisms.
18
19 Coord. Chem. Rev. 2011, 255, 1460−1479.
20
21 (16) Kember, M. R.; Buchard, A.; Williams, C. K. Catalysts for CO2/epoxide
22
23 Ring-Opening Copolymerization. Chem. Commun. 2011, 47, 141−163.
24
25 (17) Luo, M.; Li, Y.; Zhang, Y. Y.; Zhang, X. H. Using Carbon Dioxide and Its
26
27 Sulfur Analogues as Monomers in Polymer Synthesis. Polymer 2016, 82, 406−431.
28
29 (18) Paul, S.; Zhu, Y.; Romain, C.; Brooks, R.; Saini, P. K.; Williams, C. K.
30
31 Ring-opening copolymerization (ROCOP): synthesis and properties of polyesters and
32
33 polycarbonates. Chem. Commun. 2015, 51, 6459−6479.
34
35 (19) Longo, J. M.; Sanford, M. J.; Coates, G. W. Ring-Opening Copolymerization
36
37 of Epoxides and Cyclic Anhydrides with Discrete Metal Complexes:
38
39 Structure–Property Relationships. Chem. Rev. 2016, 116, 15167−15197.
40
41 (20) Ian Childers, M.; Longo, J. M.; Van Zee, N. J.; LaPointe, A. M.; Coates, G. W.
42
43 Stereoselective Epoxide Polymerization and Copolymerization. Chem. Rev. 2014, 114,
44
45 8129−8152.
46
47 (21) Lu, X.-B.; Ren, W.-M.; Wu, G.-P. CO2 Copolymers from Epoxides: Catalyst
48
49 Activity, Product Selectivity, and Stereochemistry Control. Acc. Chem. Res. 2012, 45,
50
51
1721−1735.
52
53
(22) Liu, Y.; Ren, W.-M.; Liu, J.; Lu, X.-B. Asymmetric Copolymerization of CO2
54
with meso-Epoxides Mediated by Dinuclear Cobalt(III) Complexes: Unprecedented
55
56
Enantioselectivity and Activity. Angew. Chem., Int. Ed. 2013, 52, 11594−11598.
57
58 (23) Liu, J.; Bao, Y.-Y.; Liu, Y.; Ren, W.-M.; Lu, X.-B. Binuclear chromium–salan
59
60 complex catalyzed alternating copolymerization of epoxides and cyclic anhydrides.
22

ACS Paragon Plus Environment


Page 23 of 25 Journal of the American Chemical Society

1
2
3
4 Polym. Chem. 2013, 4, 1439−1444.
5
6 (24) Li, J.; Liu, Y.; Ren, W.-M.; Lu, X.-B. Asymmetric Alternating
7
8 Copolymerization of Meso-epoxides and Cyclic Anhydrides: Efficient Access to
9
10 Enantiopure Polyesters. J. Am. Chem. Soc. 2016, 138, 11493−11496.
11
12
(25) Lv, X.-B. Stereoregular CO2 Copolymers: from Amorphous to Crystalline
13
14
Materials, Acta Polym. Sin. 2016, 9, 1166−1178.
15
(26) Ren, W.-M.; Yue, T.-J.; Zhang, X.; Gu, G.-G.; Liu, Y.; Lu, X.-B.
16
17
Stereoregular CO2 Copolymers from Epoxides with an Electron-Withdrawing Group:
18
19 Crystallization and Unexpected Stereocomplexation. Macromolecules 2017, 50,
20
21 7062−7069.
22
23 (27) Longo, J. M.; DiCiccio, A. M.; Coates, G. W. Poly(propylene succinate): A
24
25 New Polymer Stereocomplex. J. Am. Chem. Soc. 2014, 136, 15897−15900.
26
27 (28) DiCiccio, A. M.; Longo, J. M.; Rodríguez-Calero, G. G.; Coates, G. W.
28
29 Development of Highly Active and Regioselective Catalysts for the Copolymerization
30
31 of Epoxides with Cyclic Anhydrides: An Unanticipated Effect of Electronic Variation.
32
33 J. Am. Chem. Soc. 2016, 138, 7107−7113.
34
35 (29) Lamers, B. A G; van Genabeek, B.; Hennissen, J; de Waal, B. F M.; Palmans,
36
37 A. R A; Meijer, E W. Stereocomplexes of discrete, isotactic lactic acid oligomers
38
39 conjugated with oligodimethylsiloxanes. Macromolecules 2019, 52, 1200-1209.
40
41 (30) van Genabeek, B.; Lamers, Brigitte. A G; de Waal, B. F M; van Son, M. H C;
42
43 Palmans, A. R A; Meijer, E W. Amplifying (Im) perfection: The impact of
44
45 crystallinity in discrete and disperse block co-oligomers. J. Am. Chem. Soc. 2017, 139,
46
47 14869-14872.
48
49 (31) Lawrence, J.; Goto, E.; Ren, J. M; McDearmon, B.; Kim, D. S.; Ochiai, Y.;
50
51
Clark, P. G; Laitar, D.; Higashihara, T.; Hawker, C. J. A versatile and efficient
52
53
strategy to discrete conjugated oligomers. J. Am. Chem. Soc. 2017, 139,
54
13735-13739.
55
56
(32) Zhang, C.; Kim, D. S.; Lawrence, J.; Hawker, C. J; Whittaker, A. K.
57
58 Elucidating the Impact of Molecular Structure on the 19F NMR Dynamics and MRI
59
60
23

ACS Paragon Plus Environment


Journal of the American Chemical Society Page 24 of 25

1
2
3
4 Performance of Fluorinated Oligomers. ACS Macro. Lett. 2018, 7, 921-926.
5
6 (33) Kister, G.; Cassanas, G.; Vert, M. Effects of morphology, conformation and
7
8 configuration on the IR and Raman spectra of various poly(lactic acid)s. Polymer
9
10 1998, 39, 267-273.
11
12
(34) Zhang, J.-M.; Sato, H.; Tsuji, H.; Noda, I.; Ozaki, Y. Infrared Spectroscopic
13
14
Study of CH3···O=C Interaction during Poly(L-lactide)/Poly(D-lactide) Stereocomplex
15
Formation. Macromolecules 2005, 38, 1822-1828.
16
17
(35) Zhang, J.-M.; Sato, H.; Tsuji, H.; Noda, I.; Ozaki, Y. Differences in the CH3⋯
18
19 OC interactions among poly (L-lactide), poly (L-lactide)/poly (D-lactide)
20
21 stereocomplex, and poly (3-hydroxybutyrate) studied by infrared spectroscopy. J. Mol.
22
23 Struct. 2005, 735, 249-257.
24
25 (36) Pan, P.-J.; Yang, J.-J.; Shan, G.-R.; Bao, Y.-Z.; Weng, Z.-X.; Cao, A.-M.;
26
27 Yazawa, K.; Inoue, Y. Temperature-variable FTIR and solid-state 13C NMR
28
29 investigations on crystalline structure and molecular dynamics of polymorphic poly
30
31 (L-lactide) and poly (L-lactide)/poly (D-lactide) stereocomplex. Macromolecules 2011,
32
33 45, 189-197.
34
35 (37) Tsuji, H.; Horii, F.; Nakagawa, M.; Ikada, Y.; Odani, H.; Kitamaru, R.
36
37 Stereocomplex formation between enantiomeric poly (lactic acid)s. 7. Phase structure
38
39 of the stereocomplex crystallized from a dilute acetonitrile solution as studied by
40
high-resolution solid-state 13C NMR spectroscopy. Macromolecules 1992, 25,
41
42
43 4114-4118.
44
45 (38) Zhao, Z.-C.; Ren, P.-J.; Liu, Y.; Zhao, K.-B.; Lu, X-B; Zhang, W-P. Unveiling
46
47 chain–chain interactions in CO2-based crystalline stereocomplexed polycarbonates by
48
49 solid-state NMR spectroscopy and DFT calculations. J. Energy Chem. 2018, 27,
50
51
361-366.
52
53
54
55
56
57
58
59
60
24

ACS Paragon Plus Environment


Page 25 of 25 Journal of the American Chemical Society

1
2
3
4
5
6
7
8
9
10 SYNOPSIS TOC:
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
25

ACS Paragon Plus Environment

You might also like