l. wöckel,a a. seifert, c. mende, i. roth-panke, l. kroll and s. … · 2016. 11. 21. · resins...

8
FigSupplementary information L. Wöckel, a A. Seifert, a C. Mende, b I. Roth-Panke, b L. Kroll b and S. Spange* a a Polymer Chemistry, Technische Universität Chemnitz, 09107 Chemnitz, Germany. b Department of Lightweight Structures and Polymer Technology, Technische Universität Chemnitz, 09107 Chemnitz, Germany. Fig. S 1 Liquid 1 H-NMR spectra of the reaction of a) pMBC and b) mMBC with trifluoromethanesulfoic acid (TFMSA) in CDCl 3 obtained at different times. To study the reaction mechanism of the cationic polymerization of carbonates, the reaction was monitored with liquid 1 H-NMR spectroscopy. The polymerization was done at room temperature using trifluoromethanesulfonic acid (TFMSA, tenth mol regarding carbonate). Different time intervals were chosen to take a NMR sample from the reaction mixture. The 1 H-NMR signal A (-CH2- of carbonate) was integrated while calibrating the naphthaline standart and a decrease of carbonate conzentration was determined. In 1 H-NMR spectra methylol signals (-CH2OH) between 4.4 and 4.7 ppm appear. As well traces of methylether bridges (-CH2OCH2-) in the region of 4.7 and 4.9 ppm can be detected. The signals of methylenbridges (-CH2-), which can be found between 3.3 and 4.0 ppm, increases depending on time. All three carbonates can not be investigated with the same acid catalyst. The least reactive mMBC shows no polymerization with trifluoroacteic acid (TFA, pKa aq ̴ 0 1 ), methanesulfonic acid (MSA, pKa aq ̴ -2 1 ) and p-toluenesulfonic acid (pTS, pKa aq -2.8 1 ) in solution at room temperature. A strong acid trifluoromethansulfonic acid (TFMSA, pka aq ̴ -13 1 ) is nessecary to polymerized mMBC. Additionally, reaction time of several days are nessecary and low conversions can be monitored within mMBC polymerization. DFC is too reactive for polymerization with TFMSA. Extensive investigations on molecular structure formation of anisolic resins have not been reported in literature, yet. The assignment of NMR signals were performed according the published NMR data of phenolic resins because anisol as a derivative of phenol can likewise polymerize to methylene bridged anisolic resins. A overview of possible structure units and the chemical shifts in 1 H- and 13 C-NMR were shown in Table S 1 2–7 . a) pMBC + TFMSA in CDCl 3 naphthalin as standart b) mMBC + TFMSA in CDCl 3 naphthalin as standart 8 7 6 5 4 3 2 7 d 2 d 59 min Naphthalin H 2 O -CH 2 -OH [ppm] 0 min -CH 2 -O-CH 2 - A -OCH 3 -CH 2 - H Aryl H Aryl 8 7 6 5 4 3 2 2 d 16 h Naphthalin H 2 O -CH 2 -OH 360 min 59 min 1.5 min 0.3 min [ppm] 0 min -CH 2 -O-CH 2 - A -OCH 3 -CH 2 - H Aryl H Aryl Electronic Supplementary Material (ESI) for Polymer Chemistry. This journal is © The Royal Society of Chemistry 2016

Upload: others

Post on 18-Feb-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

  • FigSupplementary information

    L. Wöckel,a A. Seifert,a C. Mende,b I. Roth-Panke,b L. Kroll b and S. Spange*a

    a Polymer Chemistry, Technische Universität Chemnitz, 09107 Chemnitz, Germany.

    b Department of Lightweight Structures and Polymer Technology, Technische Universität Chemnitz, 09107 Chemnitz, Germany.

    Fig. S 1 Liquid 1H-NMR spectra of the reaction of a) pMBC and b) mMBC with trifluoromethanesulfoic acid (TFMSA) in CDCl3 obtained at different times.

    To study the reaction mechanism of the cationic polymerization of carbonates, the reaction was

    monitored with liquid 1H-NMR spectroscopy. The polymerization was done at room temperature using

    trifluoromethanesulfonic acid (TFMSA, tenth mol regarding carbonate). Different time intervals were

    chosen to take a NMR sample from the reaction mixture. The 1H-NMR signal A (-CH2- of carbonate)

    was integrated while calibrating the naphthaline standart and a decrease of carbonate conzentration

    was determined. In 1H-NMR spectra methylol signals (-CH2OH) between 4.4 and 4.7 ppm appear. As

    well traces of methylether bridges (-CH2OCH2-) in the region of 4.7 and 4.9 ppm can be detected. The

    signals of methylenbridges (-CH2-), which can be found between 3.3 and 4.0 ppm, increases depending

    on time.

    All three carbonates can not be investigated with the same acid catalyst. The least reactive mMBC

    shows no polymerization with trifluoroacteic acid (TFA, pKaaq ̴ 01), methanesulfonic acid (MSA,

    pKaaq ̴ -21) and p-toluenesulfonic acid (pTS, pKaaq -2.81) in solution at room temperature. A strong acid

    trifluoromethansulfonic acid (TFMSA, pkaaq ̴ -131) is nessecary to polymerized mMBC. Additionally,

    reaction time of several days are nessecary and low conversions can be monitored within mMBC

    polymerization. DFC is too reactive for polymerization with TFMSA.

    Extensive investigations on molecular structure formation of anisolic resins have not been reported in

    literature, yet. The assignment of NMR signals were performed according the published NMR data of

    phenolic resins because anisol as a derivative of phenol can likewise polymerize to methylene bridged

    anisolic resins. A overview of possible structure units and the chemical shifts in 1H- and 13C-NMR were

    shown in Table S 1 2–7.

    a) pMBC + TFMSA in CDCl3 naphthalin as standart b) mMBC + TFMSA in CDCl3 naphthalin as standart

    8 7 6 5 4 3 2

    7 d

    2 d

    59 min

    Naphthalin H2O-C

    H2-O

    H

    [ppm]

    0 min

    -CH

    2-O

    -CH

    2-

    A

    -OCH3

    -CH2-

    HAryl

    HAryl

    8 7 6 5 4 3 2

    2 d 16 h

    Naphthalin H2O

    -CH

    2-O

    H

    360 min

    59 min

    1.5 min

    0.3 min

    [ppm]

    0 min

    -CH

    2-O

    -CH

    2-

    A

    -OCH3

    -CH2-

    HAryl

    HAryl

    Electronic Supplementary Material (ESI) for Polymer Chemistry.This journal is © The Royal Society of Chemistry 2016

  • Table S 1 Molecular structure units in phenolic resin and the chemical shifts in 1H and 13C-NMR.

    Structure unit Chemical shift 13C 1H-NMR

    [ppm]

    Structure unit Chemical shift 1H -NMR [ppm]

    CArylCH3 18 CArylCH3 2.3

    CArylCH2CAryl- o,o’- o,p’- p,p’-

    30 35 40

    CArylCH2CAryl- 3.5-4.0

    CArylCH2-OH 58-65 CArylCH2-OH 4.4-4.7

    CArylCH2-O-CH2CAryl 66-72 CArylCH2-O-

    CH2CAryl

    4.8

    CArylOCH3 149-156 CArylOCH3 3.7

    CArylCOOH 171 CArylCOOH 11

    CArylCHO 194 CArylCHO 10

    Fig. S 2 ATR-FTIR spectra of anisolic resins and polyfurfuryl alcohol resin obtained by cationic polymerization of pMBC, mMBC and DFC.

    Fig. S 3 DSC measurement a) and ATR-FTIR spectra b) of polymers obtained by cationic polymerization of pMBC and mMBC with a surfactant (S) and trioxane (TO) at 90 °C.

    3500 3000 2500 2000 1500 1000 500

    wave number [cm1

    ]

    DFC_60

    pMBC_150

    pMBC/TO_150

    mMBC/TO_100

    mMBC_150

    (C-H)aromatic

    (CH2)

    (C=C)aromatic

    (C=O)

    a) b)

    50 100 150 200 250 300

    heat flow

    [m

    W]

    temperature [°C]

    exo

    mMBC/1TO/S_90

    pMBC/1TO/S_90

    3500 3000 2500 2000 1500 1000 500

    mMBC/1TO/S_90

    pMBC/1TO/S_90

    wave number [cm-1]

    (C=O)

  • Fig. S 4 Liquid 1H-NMR of methoxybenzyl carbonates a) pMBC and b) mMBC compared to solid state 1H-NMR of polymers without and under addition of 1,3,5-trioxane.

    Fig. S 5 Liquid 1H-NMR of difurfurylcarbonate (DFC) compared to solid state 1H-NMR of the polymer.

    Low condensated polymers of pMBC and mMBC show sharp peaks in 1H solid state NMR because of

    the flexibility of structure moieties. In contrast polymers crosslinked with 1,3,5-trioxane (TO) show

    broad signals. The polymerized DFC exhibit as well a braod signal even without adding a cross linking

    agent. The reason for this is that furfuryl moiety undergoes many side reactions where crosslinked

    structures obtained. The low condensated polymers of pMBC and mMBC show sharp signals for aryl,

    methoxy, methylene and water protons. A signal at 8.89.2 ppm may attributed to carboxylic acid

    protons resulteded from a dehydrogenation and oxidation process of methylol species. An exact

    assignment of signal is in 1H solid state NMR spectra not possible because a peak broadening and peak

    shift in solid state.

    a) b)

    12 10 8 6 4 2 0 -2

    A

    -CO

    OH

    HAryl

    mMBC

    [ppm]

    mMBC/TO_100

    mMBC_150

    H2O

    -OCH3

    -CH2-

    B

    12 10 8 6 4 2 0 -2

    B

    -CO

    OH

    pMBC

    [ppm]

    pMBC/TO_150

    pMBC_150

    HAryl

    H2O

    -OCH3

    -CH2-

    A

    12 10 8 6 4 2 0 -2

    -CH2-

    HAryl

    DFC

    [ppm]

    DFC_60

    A

  • Fig. S 6 c) Liquid 13C-{1H}-NMR spectra and e) 1H-NMR spectra of organic carbonates mMBC and pMBC compared to anisolic resins polymerized with 1,3,5-trioxane at 90 °C. a), b)and d) show the molecular structures of polymers and monomers with the assignment of atoms to signals in spectra.

    Fig. S 7 Pictures of polyfurfuryl alcohol resins obtained by the polymerization of DFC.

    Fig. S 8 Pictures of the polyfurfuryl alcohol resins obtained by the polymerization of each one-gram DFC with 10 mol% pTS, at 60 °C under utilization of different dispersing agents. One percent by weight of a PDMS with 0.8-0.9 % silanol groups, Aerosil Ox_50 with a BET surface of 50 m2/g, Aerosil 300 with a BET surface of 300 m2/g and the polydimethylsiloxane-based surfactant (S) with the designation Dabco® DC 193 was used in each experiment. As well, a polymerization with five percent by weight of Dabco® DC 193 was performed.

    b)

    a)

    220 200 180 160 140 120 100 80 60 40 20 0

    -CH2-

    -CH2- -CH3

    AC

    764 3

    21

    TO

    pTS

    pMBC

    mMBC

    5

    3'6

    A7

    4 3

    2

    [ppm]

    1

    -CH2-

    BmMBC/1TO/S_90

    pMBC/1TO/S_90p,p´

    o,p´

    o,o´

    *

    c)

    d)

    e)

    10 8 6 4 2 0

    H2O

    pMBC

    -OCH3

    HAryl

    surfactant

    [ppm]

    mMBC

    mMBC/1TO/S_90

    pMBC/1TO/S_90

    -CH3

    pTS

    Si-(CH3)3

    -CH2-

    A

    B

    -CH2-

    TO

    DFC/PDMS_60 DFC/50A®_60 DFC/300A®_60 DFC/1S_60DFC/5S_60

  • Fig. S 9 TG-MS measurements of polymers obtained by cationic polymerization of pMBC and mMBC at 100 °C. The released CO2 was detected with mass spectrometry.

    Table S 2. Performed experiments under variation of amount of 1,3,5-trioxane with the molar ratios of monomer (nM) to acid catalyst (npTS) and curing agent (nTO) and the used mass ratios of the surfactant (mS). Furthermore, the endorsement of the label indicated the final polymerization temperature. The mass loss of obtained anisolic resins after extraction, curing and pyrolysis.

    Sample

    molar ratios [mol%]

    mass ratios [wt%]

    conditions T [°C] / t [h]

    mass loss Δm [%]

    extraction thermal treatment

    nM : npTS : nTO mM : mS T1 DCM, 3d 250 °C,

    Ar 800 °C, Ar

    pMBC/4TO/S_100 1 : 0.1 : 1 1 : 0.01 100 / 4 89.2 22.7 50.0 mMBC/1TO/S_100 1 : 0.1 : 0.3 1 : 0.01 100 / 4 14.3 6.9 46.3 mMBC/2TO/S_100 1 : 0.1 : 0.7 1 : 0.01 100 / 4 14.1 12.5 38.5 mMBC/4TO/S_100 1 : 0.1 : 1 1 : 0.01 100 / 4 12.9 15.6 37.3 mMBC/6TO/S_100 1 : 0.1 : 1.3 1 : 0.01 100 / 4 10.1 16.6 36.3 pMBC/1TO/S_90 1 : 0.1 : 0.3 1 : 0.01 90 / 4 * 15.0 55.9 mMBC/1TO/S_90 1 : 0.1 : 0.3 1 : 0.01 90 / 4 * 17.1 55.1

    * soluble in DCM

    Fig. S 10 Pictures of the anisolic resins obtained by polymerization of mMBC and pMBC with 1,3,5-trioxane.

    50 100 150 200 250 300

    80

    85

    90

    95

    100

    mMBC/4TO/S_100

    mMBC/6TO/S_100

    mMBC/2TO/S_100

    mMBC/1TO/S_100

    mass [

    %]

    temperature [°C]

    0

    1

    2

    3

    4

    5

    QM

    ID *

    10

    -9 [

    A]

    TG m/z = 44 (CO2)

    mMBC/2TO/S_100 mMBC/4TO/S_100mMBC/1TO/S_100 mMBC/6TO/S_100pMBC/4TO/S_100

    pMBC/1TO/S_90

  • Fig. S 11 a) DSC measurement b) and ATR-FTIR spectrum of polymers obtained by cationic polymerization of pMBC and mMBC with a surfactant (S) and different trioxane (TO) contents at 100 °C.

    Fig. S 12 a) Liquid 13C-{1H}-NMR spectra and b) 1H-NMR spectra of extracts (labeled with E) obtained after extraction with dichlormethane.

    The extracts of polymers were investigated by means of 1H- and 13C-{1H}-NMR spectroscopy dissolving

    them in CDCl3. The 13C-{1H}-NMR spectra (Fig. S 12a) show signals for methylene (A) (29–42 ppm),

    methylol (B) (58–63 ppm), methylenether groups (C) (65–75 ppm) and aromatic carbons2,3. The extract

    of pMBC/4TO/S contains of o,o’-, o,p’- and p,p’-methylene linked anisolic resins which can be

    attributed to signals at 29.8, 35.1 and 40.6 ppm in 13C-{1H}-NMR, respectively. The mMBC/4TO/S_100E

    and mMBC/1TO/S_100E constists almost of pTS which can be seen in strong signals in the 1H- and 13C-NMR spectra (Fig. S 12a, b). In 1H-NMR the protons of methylene groups can be found between

    3.5-4.0 ppm whereas methylol and methylenether groups situated between 4.4 and 4.8 ppm4. The

    sample pMBC/4TO/S_100E shows no ether signals in NMR. The amount of methyleneether species in

    extracts of mMBC can not be determined because the signals were overlaid with methylolprotons in 1H-NMR. Traces of carbonate monomer and surfactant found, too. Additional, in extract of polymers

    produced with mMBC carbonic acid signal were found at 10.7 ppm. Signals for dichlormethan the

    extraction solvent were found in 1H-NMR, too.

    a) b)

    25 50 75 100 125 150 175 200 225 250 275 300

    mMBC/6TO/S_100

    mMBC/4TO/S_100

    mMBC/2TO/S_100

    mMBC/1TO/S_100

    heat flow

    [m

    W]

    temperature [°C]

    exo

    pMBC/4TO/S_100

    3500 3000 2500 2000 1500 1000 500

    mMBC/6TO/S_100

    mMBC/4TO/S_100

    mMBC/2TO/S_100

    mMBC/1TO/S_100

    pMBC/4TO/S_100

    wave number [cm-1]

    (C=O)

    a) b)

    10 8 6 4 2 0

    -OCH3

    HAryl

    A

    CH2Cl

    2

    surfactant

    [ppm]

    mMBC/1TO/S_100E

    mMBC/4TO/S_100E

    pMBC/4TO/S_100E

    -CO

    OH

    HAryl

    -pTS

    -CH3

    pTS

    Si-(CH3)3

    -CH2-

    B

    200 180 160 140 120 100 80 60 40 20 0

    -OC

    H3

    -CH

    2O

    CH

    2-CAryl

    mMBC/4TO/S_100E

    CAryl

    -pTS

    mMBC/1TO/S_100E

    [ppm]

    pMBC/4TO/S_100E

    3'

    6

    A

    7

    432

    1 -CH2-

    B

    -CH3

    pTS

    p,p´

    o,p´

    o,o´

  • Fig. S 13 a) The mechanism for levulinic acid formation from the reaction of furfuryl alcohol with acids8. b) Liquid 13C-{1H}-NMR spectra and b) 1H-NMR spectra of extracts (labeled with E) obtained after extraction of DFC/S_60 with dichlormethane.

    The polymer DFC/S_60 shows a extractable content of 24.6 %. The liquid 13C-{1H}-NMR spectra (Fig. S

    13b) and 1H-NMR spectra (Fig. S 13c) of the extract exhibit mainly signals of pTS and levulinic acid.

    Leviulinic acid is a side product of the cationic polymerization of furfuryl alcohol due to a hydrolytic

    furan ring opening8. The mechanism is shown in Fig. S 13a.

    Fig. S 14 SEM images of the anisolic resins obtained by cationic polymerization mMBC at 100 °C with 1,3,5-trioxane. SEM images of the annealed (A) anisolic resins at 250 °C. The inset shows the picture of the polymer foam.

    10 8 6 4 2 0

    a

    c

    CH2Cl

    2

    A

    [ppm]

    DFC/S_60E

    HAryl

    -pTS

    -CH3

    pTS

    -CH2-

    B

    d

    200 150 100 50 0

    b

    acetone-d6 a

    CAryl

    -pTS

    e

    c

    acetone-d6

    [ppm]

    DFC/S_60E

    -CH3

    pTS

    -CH2-

    d

    a)

    b) c)

    mMBC/2TO/S_100 mMBC/2TO/S_100-A

    mMBC/6TO/S_100-AmMBC/6TO/S_100

    a) b)

    c) d)

  • Fig. S 15 SEM images of the anisolic resins obtained by cationic polymerization mMBC at 100 °C with 1,3,5-trioxane. SEM images of the annealed (A) anisolic resins at 250 °C. The inset shows the picture of the polymer foam.

    Fig. S 16 The cumulative volume and pore size distribution of DFC/S_60-C and mMBC/1TO/S_90-A obtained by mercury porosimetry measurement.

    1 D. R. MacFarlane, J. M. Pringle, K. M. Johansson, S. A. Forsyth and M. Forsyth, Chem. Commun., 2006, 1905.

    2 X. Zhang, A. C. Potter and D. H. Solomon, Polymer, 1998, 39, 399–404. 3 G. E. Maciel, I. S. Chuang and L. Gollob, Macromolecules, 1984, 17, 1081–1087. 4 P. W. Kopf and E. R. Wagner, J. Polym. Sci. Polym. Chem. Ed., 1973, 11, 939–960. 5 I. S. Chuang and G. E. Maciel, Macromolecules, 1991, 24, 1025–1032. 6 I. S. Chuang and G. E. Maciel, Annu. Rep. NMR Spectrosc., 1994, 29, 169–286. 7 A. Gardziella, L. A. Pilato and A. Knop, Phenolic resins: Chemistry, Application, Standardization,

    Safety and Ecology, Springer, 2nd Completely Revised Edition., 2000. 8 A. M. Nathanael Guigo, Polym. Degrad. Stab., 2009, 94, 908–913.

    mMBC/1TO/S_100 mMBC/1TO/S_100-A

    a) b)

    mMBC/4TO/S_100-AmMBC/4TO/S_100

    c) d)

    0.010.11101001000

    0.0

    1.0

    2.0

    3.0

    4.0

    5.0

    cum. volume

    cum

    ula

    tive p

    ore

    volu

    me [cm

    3/g

    ]

    pore diameter [m]

    0.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    mMBC/1TO/S_90A-C

    rel. pore volume

    rela

    tive p

    ore

    volu

    me [%

    ]

    DFC/S_60-C