Stephen A. Miller

Find an error

Name: Miller, Stephen A.
Organization: University of Florida , USA
Department: Department of Chemistry
Title: Associate(PhD)

TOPICS

Co-reporter:Ha Thi Hoang Nguyen;Gabriel N. Short;Pengxu Qi
Green Chemistry (1999-Present) 2017 vol. 19(Issue 8) pp:1877-1888
Publication Date(Web):2017/04/20
DOI:10.1039/C6GC03238A
The general and efficient copolymerization of lactones with hydroxy-acid bioaromatics was accomplished via a concurrent ring-opening polymerization (ROP) and polycondensation methodology. Suitable lactones were L-lactide or ε-caprolactone and four hydroxy-acid comonomers were prepared as hydroxyethyl variants of the bioaromatics syringic acid, vanillic acid, ferulic acid, and p-coumaric acid. Copolymerization conditions were optimized on a paradigm system with a 20 : 80 feed ratio of caprolactone : hydroxyethylsyringic acid. Among six investigated catalysts, polymer yield was optimized with 1 mol% of Sb2O3, affording eight copolymer series in good yields (32–95% for lactide; 80–95% for caprolactone). Half of the polymers were soluble in the GPC solvent hexafluoroisopropanol and analyzed to high molecular weight, with Mn = 10 500–60 700 Da. Mass spectrometry and 1H NMR analysis revealed an initial ring-opening formation of oligolactones, followed by polycondensation of these with the hydroxy-acid bioaromatic, followed by transesterification, yielding a random copolymer. By copolymerizing bioaromatics with L-lactide, the glass transition temperature (Tg) of polylactic acid (PLA, 50 °C) could be improved and tuned in the range of 62–107 °C; the thermal stability (T95%) of PLA (207 °C) could be substantially increased up to 323 °C. Similarly, bioaromatic incorporation into polycaprolactone (PCL, Tg = −60 °C) accessed an improved Tg range from −48 to 105 °C, while exchanging petroleum-based content with biobased content. Thus, this ROP/polycondensation methodology yields substantially or fully biobased polymers with thermal properties competitive with incumbent packaging thermoplastics such as polyethylene terephthalate (Tg = 67 °C) or polystyrene (Tg = 95 °C).
Co-reporter:Mayra Rostagno;Steven Shen;Ion Ghiviriga
Polymer Chemistry (2010-Present) 2017 vol. 8(Issue 34) pp:5049-5059
Publication Date(Web):2017/08/30
DOI:10.1039/C7PY00205J
A series of polyvinyl aromatic acetals was obtained from the condesation of commercially available polyvinyl alcohol (PVA) and sustainable aromatic aldehydes: 4-hydroxybenzaldehyde, vanillin, syringaldehyde, ethylvanillin, ortho-vanillin, isovanillin, salicylaldehyde, ortho-anisaldehyde, para-anisaldehyde, benzaldehyde, cinnamaldehyde, cuminaldehyde, and hydroxymethylfurfural. The degree of acetalization was determined by 1H NMR and found to range from 54 to 75%, resulting in calculated number average molecular weights of 38 200 to 46 000 Da. The glass transition temperature (Tg) of PVA (75 °C) was increased to substantially excel over that of polystyrene (PS, 100 °C), with polyvinyl aromatic acetal Tg values ranging from 114 to 157 °C. Heterogeneous degradation studies of polyvinyl vanillin acetal at room temperature indicated nearly complete hydrolysis over 24 hours in acidic aqueous media. For example, at pH = 1, the initial acetalization of 54% dropped to 0.8%, generating benign aromatic small molecules and water-soluble, biodegradable PVA.
Co-reporter:Pengxu Qi, Hsiao-Li Chen, Ha Thi Hoang Nguyen, Chu-Chieh Lin and Stephen A. Miller  
Green Chemistry 2016 vol. 18(Issue 15) pp:4170-4175
Publication Date(Web):23 Jun 2016
DOI:10.1039/C6GC01081D
Itaconic acid, a naturally occurring compound mass-produced via fermentation of glucose, was reacted with ethanolamine or ethylene diamine to afford hydroxy-acid or diacid monomers containing the 2-pyrrolidone lactam. Homopolymerization or copolymerization with diols, respectively, yielded polylactam esters with higher glass transition temperatures and faster hydrolytic degradation compared to the commercial polyester polylactic acid (PLA).
Co-reporter:Ertugrul Sahmetlioglu, Ha Thi Hoang Nguyen, Olivier Nsengiyumva, Ersen Göktürk, and Stephen A. Miller
ACS Macro Letters 2016 Volume 5(Issue 4) pp:466
Publication Date(Web):March 23, 2016
DOI:10.1021/acsmacrolett.6b00095
A kinetic study revealed that the acid-catalyzed (p-TSA) equilibration of Me2Si(OMe)2 and Me2Si(OEt)2, forming Me2Si(OEt)OMe, is established in 300 min in benzene at room temperature. This silicon acetal metathesis reaction is exploited for the step-growth polymerization of bis-silicon acetals (MeOSiMe2OROSiMe2OMe) with metathetical loss of Me2Si(OMe)2. Thus, a convenient and generalized silicon acetal metathesis polymerization (SAMP) method is introduced as the acid-catalyzed copolymerization of a diol (HOROH) and Me2Si(OMe)2, driven by elimination of methanol and/or Me2Si(OMe)2. SAMP constitutes an effective and powerful strategy for manipulating the most common bond in the Earth’s crust, the silicon–oxygen bond.
Co-reporter:Mayra Rostagno;Erik J. Price;Alexer G. Pemba;Ion Ghiriviga;Khalil A. Abboud
Journal of Applied Polymer Science 2016 Volume 133( Issue 45) pp:
Publication Date(Web):
DOI:10.1002/app.44089

ABSTRACT

The polymerization of biorenewable molecules to polymers with hydrolyzable main-chain functionality is one approach to identifying sustainable replacements for common, environmentally unsound packaging plastics. Bioaromatic polyacetals were synthesized via acid-catalyzed acetal formation from dialdehydes and tetraols. Ethylene linked dialdehyde monomers VV and SS were constructed from bioaromatics vanillin and syringaldehyde, respectively. Tetraol monomers included biogenic erythritol (E), along with pentaerythritol (P), and ditrimethylolpropane (D). Four copolymer series were prepared with varying tetraol content: E/P-VV; E/D-VV; E/P-SS; and E/D-SS. Number average molecular weights (Mn) ranged from 1,400 to 27,100 Da. Generally, the copolymerization yields were inversely proportional to the feed fraction of erythritol (E), implying that tetraols P and D react more readily. The materials were typically amorphous and exhibited glass transition temperatures (Tg) ranging from 57 to 159 °C, suitably mimicking the Tg values of several commodity plastics. The syringaldehyde-based copolymers exhibited a higher Tg range (71–159 °C) than the vanillin-based copolymers (57–110 °C). Accelerated degradation studies in aqueous HCl (3 M, 6 M, concentrated) over 24 h showed that degradation (Mn decrease) was proportional to the acid concentration. A one-year degradation study of E50/D50-SS (from 50% feed of erythritol) in seawater, deionized water, tap water, or pH 5 buffer showed no Mn decrease; but in pH 1 buffer, the decrease was 40% (18,800 to 11,200). © 2016 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2016, 133, 44089.

Co-reporter:Ha Thi Hoang Nguyen, Marcus H. Reis, Pengxu Qi and Stephen A. Miller  
Green Chemistry 2015 vol. 17(Issue 9) pp:4512-4517
Publication Date(Web):14 Jul 2015
DOI:10.1039/C5GC01104C
Ferulic acid and p-coumaric acid are abundant, biorenewable precursors for the synthesis of polyethylene ferulate (PEF) and polyethylene coumarate (PEC), as well as cognate copolymers with prescribed hydrogenation of the main-chain double bond. By controlling the comonomer feed ratios, copolymers with tunable thermal properties are obtained, including the thermal range occupied by polystyrene (PS).
Co-reporter:Ersen Göktürk, Alexander G. Pemba and Stephen A. Miller  
Polymer Chemistry 2015 vol. 6(Issue 21) pp:3918-3925
Publication Date(Web):28 Apr 2015
DOI:10.1039/C5PY00230C
We present a new approach to synthesizing polyglycolic acid (PGA) via the cationic alternating copolymerization of formaldehyde (from trioxane) and carbon monoxide (CO), sustainable C1 feedstocks obtainable from biomethanol or biogas. This method constitutes an inexpensive and efficient pathway for the synthesis of PGA, circumventing the usual route requiring glycolide. PGA was successfully synthesized with yields up to 92% from trioxane, 800 psi of CO, and 1 mol% triflic acid (TfOH) initiator at 170 °C over three days. 1H NMR, 13C NMR, and FT-IR spectra of the polymer from CO and trioxane are identical to those of commercial PGA prepared via the ring-opening polymerization of glycolide—confirming the alternating microstructure. Although high copolymerization conversions were obtained, molecular weight analysis usually suggested the formation of oligomeric glycolic acid (OGA). High molecular weight PGA can be obtained via post-polymerization polycondensation of OGA catalyzed by Zn(OAc)2·2H2O. Alternatively, increased molecular weight PGA can be achieved by inclusion of glycerol as a branching agent during the C1 copolymerization.
Co-reporter:Ryan T. Martin, Ludmila P. Camargo and Stephen A. Miller  
Green Chemistry 2014 vol. 16(Issue 4) pp:1768-1773
Publication Date(Web):26 Feb 2014
DOI:10.1039/C3GC42604A
Polyesteracetal (PEA) copolymers of 1,3-dioxolan-4-one (DOX) with L-lactide (LA) were synthesized via ring-opening polymerization (ROP). Facile degradability in distilled water or seawater was observed—presumably via hydrolysis of the main-chain acetal functionality. Over 45 days, mass loss approached 2% and molecular weight loss (Mn) exceeded 15%; control experiments with polylactic acid (PLA) indicated no degradation.
Co-reporter:Stephen A. Miller  
Polymer Chemistry 2014 vol. 5(Issue 9) pp:3117-3118
Publication Date(Web):07 Mar 2014
DOI:10.1039/C4PY90017K
A graphical abstract is available for this content
Co-reporter:Alexander G. Pemba, Mayra Rostagno, Tanner A. Lee and Stephen A. Miller  
Polymer Chemistry 2014 vol. 5(Issue 9) pp:3214-3221
Publication Date(Web):25 Feb 2014
DOI:10.1039/C4PY00178H
Monomers structurally resembling lignin were prepared by reacting 4-hydroxybenzaldehyde, vanillin, syringaldehyde, (each bio-available) or ethylvanillin (synthetic) with dibromoethane, yielding dialdehydes CHO–Ar–OCH2CH2O–Ar–CHO. Condensation copolymerization with tetraols catalyzed by para-toluene sulfonic acid yielded polyacetal ethers with cyclic acetals in the case of di-trimethylolpropane (di-TMP) and spirocyclic acetals in the case of pentaerythritol (PTOL). Number average molecular weights (Mn) were in the range of 10600 to 22200, although the insolubility of those polymers based on 4-hydroxybenzaldehyde precluded this measurement. The polymers are thermally robust and exhibit 5% mass loss via thermogravimetric analysis in the range of 307–349 °C. Those copolymers based on PTOL displayed glass transition (Tg) temperatures (108–152 °C) at least 40 °C higher than their di-TMP analogues (68–98 °C), highlighting the added rigidity conferred by spirocyclic acetals versus cyclic acetals. Preliminary degradation studies were conducted in dimethyl sulfoxide with 0.5% added aqueous HCl (concentrated or 2 M). Dynamic light scattering confirmed the facile hydrolysis of the polymers. Generally, polymer degradation was faster with a higher acid concentration and copolymers from the PTOL tetraol were more resistant to hydrolysis than those from the di-TMP tetraol.
Co-reporter:Alexander G. Pemba, Jeniree A. Flores and Stephen A. Miller  
Green Chemistry 2013 vol. 15(Issue 2) pp:325-329
Publication Date(Web):17 Dec 2012
DOI:10.1039/C2GC36588J
Polyalkylene acetals are synthesized from simple bio-derived diols and diethoxymethane via a novel technique involving interchange of acetal functional groups. This Acetal Metathesis Polymerization (AMP) method provides a route to polyacetals that are designed to degrade under abiotic conditions via simple acid-catalyzed hydrolysis.
Co-reporter:Stephen A. Miller
ACS Macro Letters 2013 Volume 2(Issue 6) pp:550
Publication Date(Web):June 5, 2013
DOI:10.1021/mz400207g
The field of sustainable polymers is growing and evolving at unprecedented rates. Researchers are increasingly concerned with the feedstock origins and the degradation behavior of, especially, large-scale commodity packaging plastics. A perspective is offered here for the design of sustainable polymers, specifically addressing opportunities for monomer development and polymer degradation. Key concepts include: water degradability instead of biodegradability; incorporation of novel main-chain functionality, such as acetals; utilization of lignin-based aromatics; and direct polymerization of biogenic C1 feedstocks.
Co-reporter:Jianfang Chai, Khalil A. Abboud and Stephen A. Miller  
Dalton Transactions 2013 vol. 42(Issue 25) pp:9139-9147
Publication Date(Web):01 Mar 2013
DOI:10.1039/C3DT50163A
Several analogues of the sterically expanded constrained geometry catalyst Me2Si(η1-C29H36)(η1-N-tBu)ZrCl2·OEt2 (2) were synthesized to assess the effect on branching and molecular weight for ethylene homopolymerization. Catalysts based on tetramethyltetrahydrobenzofluorene (TetH), ethylTetH, t-butylTetH, and octamethyloctahydrodibenzofluorenyl (OctH) bearing a diphenylsilyl bridge were prepared and characterized: Me2Si(η5-C21H22)(η1-N-tBu)ZrCl2 (3); Me2Si(η5-C23H26)(η1-N-tBu)ZrCl2 (4); and Me2Si(η5-C25H30)(η1-N-tBu)ZrCl2 (5); Me2Si(η5-C21H22)(η1-N-tBu)ZrMe2 (6); and Ph2Si(η5-C29H36)(η1-N-tBu)ZrCl2 (7). Complexes 4, 5, 6, and 7 were characterized by X-ray crystallography and displayed η5 hapticity to the carbon ring in each case, in contrast to 2. In comparison to 2, complexes 3, 4, 5, and 7 (in combination with methylaluminoxane = MAO) showed diminished branching, higher molecular weight, and higher polydispersity indices for obtained ethylene homopolymers. While 4/MAO produced the greatest molecular weight polymers, no branching was observed. Reactivity ratios were determined for the copolymerization of ethylene and 1-decene with 2/MAO. A value of rethylene = 14.9 and an exceedingly high value of r1-decene = 0.49 were found—in line with previous reports of this catalyst's unusual affinity for α-olefins.
Co-reporter:Hsuan-Ying Chen ; Laurent Mialon ; Khalil A. Abboud
Organometallics 2012 Volume 31(Issue 15) pp:5252-5261
Publication Date(Web):July 19, 2012
DOI:10.1021/om300121c
A series of Li, Na, Mg, and Ca complexes with 2,6-di-tert-butyl-4-methylphenol (BHT) as the ligand has been synthesized, and their reactivity for the ring-opening polymerization of lactide has been studied. The Ca complex with 2,2′-ethylidenebis(4,6-di-tert-butylphenol) (EDBP) as the ligand also has been synthesized to compare with the BHT systems. All complexes, in the presence of benzyl alcohol as initiator, exhibit high activity for the ring-opening polymerization of lactide. Polymerization activities follow the order of Na ≫ Li > Ca ≫ Mg. Additionally, metal–BHT systems are more efficient than metal–EDBP systems, and this superiority was pronounced in THF solution. Four model metal–BHT complexes with coordinated dimethoxyethane (DME) have been characterized by single-crystal X-ray diffraction. These complexes exhibited a wide variation in the Cipso–O–metal angle (159.3° to 180.0°), and DFT calculations suggested that this flexibility allows the metal to vary its electron density and thereby expedite the catalytic cycle that requires both monomer activation and substrate lability.
Co-reporter:Laurent Mialon;Rob Verhenst;Alexer G. Pemba
Macromolecular Rapid Communications 2011 Volume 32( Issue 17) pp:1386-1392
Publication Date(Web):
DOI:10.1002/marc.201100242
Co-reporter:Laurent Mialon;Rob Verhenst;Alexer G. Pemba
Macromolecular Rapid Communications 2011 Volume 32( Issue 17) pp:
Publication Date(Web):
DOI:10.1002/marc.201190045
Co-reporter:Laurent Mialon, Alexander G. Pemba and Stephen A. Miller  
Green Chemistry 2010 vol. 12(Issue 10) pp:1704-1706
Publication Date(Web):06 Jul 2010
DOI:10.1039/C0GC00150C
Lignin-based vanillin and acetic anhydride are subjected to the Perkin reaction and then hydrogenation to afford acetyldihydroferulic acid. Polymerization of this monomer yields poly(dihydroferulic acid), which exhibits thermal properties functionally similar to those of polyethylene terephthalate (PET).
Co-reporter:Eric D. Schwerdtfeger, Craig J. Price, Jianfang Chai and Stephen A. Miller
Macromolecules 2010 Volume 43(Issue 11) pp:4838-4842
Publication Date(Web):May 11, 2010
DOI:10.1021/ma100545q
A zirconium-based polymerization catalyst, Me2Si(η1-C29H36)(η1-N-tBu)ZrCl2·OEt2 (1), and a chromium-based oligomerization catalyst, [(tBuSCH2CH2)2NH]CrCl3 (2), in combination with methylaluminoxane (MAO), constitute a novel tandem catalyst system for converting ethylene alone to linear low-density polyethylene with short and long branches. Independent reactions demonstrated that 1/MAO converts ethylene to polyethylene with ethyl and long (≥6C) branches while 2/MAO converts ethylene to 1-hexene and 1-octene. Under certain reaction conditions (80 °C, 120 psi ethylene), 2/MAO produces 1-hexene and 1-octene from ethylene with 80−85% selectivity for 1-hexene and minimal production of polymer. Polymerization of ethylene by 1/MAO in the presence of 2/MAO yields polyethylene with 14−44 total branches per 1000 carbon atoms, with 4−16 of those branches arising from incorporation of 1-hexene. The fraction of branches arising from 1-hexene varies from 12 to 54%, depending on the relative amounts of 1, 2, and MAO present, as well as on the pressure of ethylene and the prepolymerization time. By adjusting the polymerization conditions, the branching characteristics of polyethylene could be varied in a controlled manner.
Co-reporter:James W. Ogle and Stephen A. Miller  
Chemical Communications 2009 (Issue 38) pp:5728-5730
Publication Date(Web):02 Sep 2009
DOI:10.1039/B914732B
The catalytic activity of iridium-mediated transfer hydrogenation is readily tuned by electronic variation of the ligated tetraaryl-N-heterocyclic carbene and the installation of electron donating groups on the N-aryl substituents is more important than on the C-aryl substituents for effecting catalytic enhancement.
Co-reporter:CraigJ. Price;Hsuan-Ying Chen;L.Marie Launer;StephenA. Miller
Angewandte Chemie 2009 Volume 121( Issue 5) pp:
Publication Date(Web):
DOI:10.1002/ange.200990009
Co-reporter:CraigJ. Price;Hsuan-Ying Chen;L.Marie Launer;StephenA. Miller
Angewandte Chemie 2009 Volume 121( Issue 5) pp:974-977
Publication Date(Web):
DOI:10.1002/ange.200802605
Co-reporter:CraigJ. Price;Hsuan-Ying Chen;L.Marie Launer;StephenA. Miller
Angewandte Chemie International Edition 2009 Volume 48( Issue 5) pp:
Publication Date(Web):
DOI:10.1002/anie.200990007
Co-reporter:CraigJ. Price;Hsuan-Ying Chen;L.Marie Launer;StephenA. Miller
Angewandte Chemie International Edition 2009 Volume 48( Issue 5) pp:956-959
Publication Date(Web):
DOI:10.1002/anie.200802605
Co-reporter:Ryan T. Martin
Macromolecular Symposia 2009 Volume 279( Issue 1) pp:72-78
Publication Date(Web):
DOI:10.1002/masy.200950512

Abstract

The ability of an alkyl branch to depress the melting temperature in a polyoxymethylene chain is measurably less than that in a polyethylene chain. The factors that inhibit the alkyl-branch plasticization of polyoxymethylene are considered by computational assessment of a series of model compounds at various levels of theory: DFT B3LYP 6-31+G*, DFT B3LYP 6-311++G**, MP2 cc-pVTZ, T1, and G3(MP2). Intramolecular interactions—characterized as acetal CH···O hydrogen bonds—are surprisingly strong and likely encourage conformational regularity in the vicinity of the alkyl branches, allowing maintenance of the intermolecular chain-chain interactions. The acetal CH···O hydrogen bonds in dimethylene glycol average to 2.65 kcal/mol while the non-acetal CH···O interactions in 1,3-propanediol are much weaker with an average of 0.34 kcal/mol (G3(MP2)). The related, classical OH···O hydrogen bond in ethylene glycol is found to be worth 2.12 kcal/mol. To describe this energetic ordering, an additional stabilizing anomeric effect is invoked for dimethylene glycol, a model for polyoxymethylene.

Co-reporter:Craig J. Price ; Paul D. Zeits ; Joseph H. Reibenspies
Organometallics 2008 Volume 27(Issue 15) pp:3722-3727
Publication Date(Web):July 8, 2008
DOI:10.1021/om8001526
Incorporating the octamethyloctahydrodibenzofluorenyl (Oct) ligand into metallocene and constrained geometry olefin polymerization catalysts has profound catalytic consequences. The steric influences are undoubtedly important, but it is shown herein that electronics likely also play a crucial role. The electron richness of the Oct− anion was directly measured by competitive deprotonation experiments, which reveal that the pKa of OctH is 3.9 units higher than that of fluorene. The HOMO−LUMO gap decreases by about 0.6 kcal/mol for each additional tertiary alkyl group appended to the metallocene R2C(C5H4)(C13H8)ZrCl2 (R = Me or Ph) in the 2, 3, 6, and 7 positions of the fluorenyl moiety, indicating the ability of these groups to increase the HOMO energy by electron donation. The carbonyl stretching frequencies for η5-OctMn(CO)3 (2009, 1924 cm−1) demonstrated that the Oct ligand is the most electron donating in the series of CpMn(CO)3, Cp*Mn(CO)3, and (C13H9)Mn(CO)3. DFT calculations universally corroborate these experimental findings.
Co-reporter:Jianfang Chai, Khalil A. Abboud and Stephen A. Miller
Dalton Transactions 2013 - vol. 42(Issue 25) pp:NaN9147-9147
Publication Date(Web):2013/03/01
DOI:10.1039/C3DT50163A
Several analogues of the sterically expanded constrained geometry catalyst Me2Si(η1-C29H36)(η1-N-tBu)ZrCl2·OEt2 (2) were synthesized to assess the effect on branching and molecular weight for ethylene homopolymerization. Catalysts based on tetramethyltetrahydrobenzofluorene (TetH), ethylTetH, t-butylTetH, and octamethyloctahydrodibenzofluorenyl (OctH) bearing a diphenylsilyl bridge were prepared and characterized: Me2Si(η5-C21H22)(η1-N-tBu)ZrCl2 (3); Me2Si(η5-C23H26)(η1-N-tBu)ZrCl2 (4); and Me2Si(η5-C25H30)(η1-N-tBu)ZrCl2 (5); Me2Si(η5-C21H22)(η1-N-tBu)ZrMe2 (6); and Ph2Si(η5-C29H36)(η1-N-tBu)ZrCl2 (7). Complexes 4, 5, 6, and 7 were characterized by X-ray crystallography and displayed η5 hapticity to the carbon ring in each case, in contrast to 2. In comparison to 2, complexes 3, 4, 5, and 7 (in combination with methylaluminoxane = MAO) showed diminished branching, higher molecular weight, and higher polydispersity indices for obtained ethylene homopolymers. While 4/MAO produced the greatest molecular weight polymers, no branching was observed. Reactivity ratios were determined for the copolymerization of ethylene and 1-decene with 2/MAO. A value of rethylene = 14.9 and an exceedingly high value of r1-decene = 0.49 were found—in line with previous reports of this catalyst's unusual affinity for α-olefins.
Co-reporter:James W. Ogle
Chemical Communications 2009(Issue 38) pp:NaN5730-5730
Publication Date(Web):2009/10/14
DOI:10.1039/B914732B
The catalytic activity of iridium-mediated transfer hydrogenation is readily tuned by electronic variation of the ligated tetraaryl-N-heterocyclic carbene and the installation of electron donating groups on the N-aryl substituents is more important than on the C-aryl substituents for effecting catalytic enhancement.
POLY[OXY(2-METHOXY-1,4-PHENYLENE)(3-OXO-1-PROPENE-1,3-DIYL)]
1-(2-Hydroxyethyl)-5-oxopyrrolidine-3-carboxylic acid
Poly[oxy(1-oxo-1,6-hexanediyl)]
3-Pyrrolidinecarboxylicacid, 1,1'-(1,2-ethanediyl)bis[5-oxo-
6H-Benzo[b]fluorene,2-(1,1-dimethylethyl)-7,8,9,11-tetrahydro-6,6,9,9-tetramethyl-
2-PROPENOIC ACID, 3-[4-[(6-HYDROXYHEXYL)OXY]-3-METHOXYPHENYL]-, (2E)-
Poly[oxy(1,2-dioxo-1,2-ethanediyl)oxy(2,2-dimethyl-1,3-propanediyl)]
3-[4-(2-hydroxyethoxy)phenyl]prop-2-enoic Acid
Benzoic acid, 4-(3-hydroxypropoxy)-3-methoxy-
1,1,4,4,7,7,10,10-OCTAMETHYL-2,3,4,7,8,9, 10,12-OCTAHYDRO-1H-DIBENZO[B,H]FLUORENE