Polymer Dynamic Article LinksC Chemistryrsc.anu.edu.au/~cylin/Publication/38.Diels-Alder.pdf · The...

12
Thermally reversible Diels–Alder-based polymerization: an experimental and theoretical assessmentJiawen Zhou, a Nathalie K. Guimard, a Andrew J. Inglis, a Mansoor Namazian, b Ching Y. Lin, b Michelle L. Coote, * b Emmanouil Spyrou, d Stefan Hilf, c Friedrich Georg Schmidt d and Christopher Barner-Kowollik * a Received 9th August 2011, Accepted 17th September 2011 DOI: 10.1039/c1py00356a A pair of monomers capable of undergoing reversible polymerization—based on reversible Diels–Alder (DA) chemistry—as a function of the applied reaction temperature is presented. Specifically, the reaction of isophorone bis(sorbic carbamate), a difunctional diene, with 1,4-phenylenebis(methylene) bis((diethoxyphosphoryl)methanedithioformate), a difunctional dithioester, was studied in detail. Various factors, including the monomer concentration, the type of solvent, and the presence of a Lewis acid, that influence this step-growth polymerization were evaluated. The solvent type was found to have a significant effect on the DA reaction rate. Under the optimized conditions, which are 1.8 g mol 1 of each monomer in acetonitrile with 1.1 equivalents of zinc chloride at 50 C for 4 h, a polymer with a peak molecular weight of 9600 g mol 1 (relative to poly(styrene) standards) was obtained. The resulting polymer was employed to investigate the correlation between time, temperature, and percentage of debonded monomers achieved during the retro DA (rDA) reaction. In addition, theoretical predictions of the rDA temperature were obtained via ab initio quantum chemical calculations. The monomeric diene and dienophile system was employed for the calculations of the equilibrium constants at various rDA reaction temperatures to correlate the percentage of bonded molecules with the applied temperature. It was calculated that 60% of the polymer becomes debonded at a temperature (T qc ) of around 220 C, a result that agrees well with that obtained experimentally (T exp ¼ 219 C). Introduction The [4 + 2] cycloaddition of an electron rich diene with an elec- tron poor dienophile was discovered by Otto Diels and Kurt Alder in 1928. 1 The simple and straightforward reaction that bears their names has revolutionized the field of organic chem- istry. 2 Consequently, Diels and Alder obtained the Nobel Prize for their achievement in 1950. Recently, a number of scientists in polymer and material chemistry have focused on the Diels–Alder (DA) reaction because it is high yielding, regio- and stereo- controlled, thermally reversible, and even—as recently demon- strated—mechanically reversible. 3 Furthermore, by employing judiciously selected diene and dienophile functional groups as reactive centers, thermally switchable systems can be readily obtained. Polymers that possess such thermal properties can be utilized to fabricate materials that are inherently capable of self- healing. 4 When a defect is introduced into the polymer matrix the material can be heated to allow for the retro DA (rDA) process to occur. Once the material is cooled, the polymer chains—which have rearranged to accommodate the defect—re-crosslink. 4 A wide variety of mendable materials have been prepared by employing such diene–dienophile-based cross-linking tech- niques. 5 However, the majority of these mendable materials are maleimide- and furan-based compounds. 6 The thermoreversible DA reaction of furan with maleimide results in a bicyclic struc- ture, which will undergo a rDA reaction above 90 C. 7 In addi- tion to approaches based on reversible covalent linkages, a Preparative Macromolecular Chemistry, Institut f ur Technische Chemie und Polymerchemie, Karlsruhe Institute of Technology (KIT), Engesser Strasse 18, 76128 Karlsruhe, Germany. E-mail: christopher. [email protected]; Fax: +49 721 608-45740 b ARC Centre of Excellence for Free-Radical Chemistry and Biotechnology, Research School of Chemistry, Australian National University, Canberra, ACT 0200, Australia. E-mail: [email protected]; Fax: +61 2 6125 0750 c Evonik Industries AG, Rodenbacher Chaussee 4, 63457 Hanau-Wolfgang, Germany d Evonik Industries AG, Paul-Baumann-Strasse 1, 45764 Marl, Germany † Electronic supplementary information (ESI) available: Tables for the detailed information about molecular weight and PDI of the polymers 7–13, experimental and theoretical m/z values for the structures corresponding to the peaks in Fig. 3b, and summary of conditions of rDA reactions of 7 in acetonitrile at 120, 140, 160, 180, and 219 C and the corresponding calculated percentage of debonding are presented. In addition, tables of the zero-point-energy (ZPE), thermal correction (TC), entropy (S), electronic energy at 0K (E(0K), at the employed level of theories). See DOI: 10.1039/c1py00356a This journal is ª The Royal Society of Chemistry 2011 Polym. Chem. Dynamic Article Links C < Polymer Chemistry Cite this: DOI: 10.1039/c1py00356a www.rsc.org/polymers PAPER Downloaded on 05 January 2012 Published on 26 October 2011 on http://pubs.rsc.org | doi:10.1039/C1PY00356A View Online / Journal Homepage

Transcript of Polymer Dynamic Article LinksC Chemistryrsc.anu.edu.au/~cylin/Publication/38.Diels-Alder.pdf · The...

Thermally reversible Diels–Alder-based polymerization: an experimental andtheoretical assessment†

Jiawen Zhou,a Nathalie K. Guimard,a Andrew J. Inglis,a Mansoor Namazian,b Ching Y. Lin,b

Michelle L. Coote,*b Emmanouil Spyrou,d Stefan Hilf,c Friedrich Georg Schmidtd andChristopher Barner-Kowollik*a

Received 9th August 2011, Accepted 17th September 2011

DOI: 10.1039/c1py00356a

A pair of monomers capable of undergoing reversible polymerization—based on reversible Diels–Alder

(DA) chemistry—as a function of the applied reaction temperature is presented. Specifically, the

reaction of isophorone bis(sorbic carbamate), a difunctional diene, with 1,4-phenylenebis(methylene)

bis((diethoxyphosphoryl)methanedithioformate), a difunctional dithioester, was studied in detail.

Various factors, including the monomer concentration, the type of solvent, and the presence of a Lewis

acid, that influence this step-growth polymerization were evaluated. The solvent type was found to have

a significant effect on the DA reaction rate. Under the optimized conditions, which are 1.8 g mol!1 of

each monomer in acetonitrile with 1.1 equivalents of zinc chloride at 50 "C for 4 h, a polymer with

a peak molecular weight of 9600 g mol!1 (relative to poly(styrene) standards) was obtained. The

resulting polymer was employed to investigate the correlation between time, temperature, and

percentage of debonded monomers achieved during the retro DA (rDA) reaction. In addition,

theoretical predictions of the rDA temperature were obtained via ab initio quantum chemical

calculations. The monomeric diene and dienophile system was employed for the calculations of the

equilibrium constants at various rDA reaction temperatures to correlate the percentage of bonded

molecules with the applied temperature. It was calculated that 60% of the polymer becomes debonded

at a temperature (Tqc) of around 220 "C, a result that agrees well with that obtained experimentally

(Texp # 219 "C).

Introduction

The [4 + 2] cycloaddition of an electron rich diene with an elec-

tron poor dienophile was discovered by Otto Diels and Kurt

Alder in 1928.1 The simple and straightforward reaction that

bears their names has revolutionized the field of organic chem-

istry.2 Consequently, Diels and Alder obtained the Nobel Prize

for their achievement in 1950. Recently, a number of scientists in

polymer and material chemistry have focused on the Diels–Alder

(DA) reaction because it is high yielding, regio- and stereo-

controlled, thermally reversible, and even—as recently demon-

strated—mechanically reversible.3 Furthermore, by employing

judiciously selected diene and dienophile functional groups as

reactive centers, thermally switchable systems can be readily

obtained. Polymers that possess such thermal properties can be

utilized to fabricate materials that are inherently capable of self-

healing.4 When a defect is introduced into the polymer matrix the

material can be heated to allow for the retro DA (rDA) process

to occur. Once the material is cooled, the polymer chains—which

have rearranged to accommodate the defect—re-crosslink.4

A wide variety of mendable materials have been prepared by

employing such diene–dienophile-based cross-linking tech-

niques.5 However, the majority of these mendable materials are

maleimide- and furan-based compounds.6 The thermoreversible

DA reaction of furan with maleimide results in a bicyclic struc-

ture, which will undergo a rDA reaction above 90 "C.7 In addi-

tion to approaches based on reversible covalent linkages,

aPreparative Macromolecular Chemistry, Institut f!ur Technische Chemieund Polymerchemie, Karlsruhe Institute of Technology (KIT), EngesserStrasse 18, 76128 Karlsruhe, Germany. E-mail: [email protected]; Fax: +49 721 608-45740bARCCentre of Excellence for Free-Radical Chemistry and Biotechnology,Research School of Chemistry, Australian National University, Canberra,ACT 0200, Australia. E-mail: [email protected]; Fax: +61 2 61250750cEvonik Industries AG, Rodenbacher Chaussee 4, 63457 Hanau-Wolfgang,GermanydEvonik Industries AG, Paul-Baumann-Strasse 1, 45764 Marl, Germany† Electronic supplementary information (ESI) available: Tables for thedetailed information about molecular weight and PDI of the polymers7–13, experimental and theoretical m/z values for the structurescorresponding to the peaks in Fig. 3b, and summary of conditions ofrDA reactions of 7 in acetonitrile at 120, 140, 160, 180, and 219 "C andthe corresponding calculated percentage of debonding are presented. Inaddition, tables of the zero-point-energy (ZPE), thermal correction(TC), entropy (S), electronic energy at 0K (E(0K), at the employedlevel of theories). See DOI: 10.1039/c1py00356a

This journal is ª The Royal Society of Chemistry 2011 Polym. Chem.

Dynamic Article LinksC<PolymerChemistryCite this: DOI: 10.1039/c1py00356a

www.rsc.org/polymers PAPER

Dow

nloa

ded

on 0

5 Ja

nuar

y 20

12Pu

blish

ed o

n 26

Oct

ober

201

1 on

http

://pu

bs.rs

c.or

g | d

oi:1

0.10

39/C

1PY

0035

6AView Online / Journal Homepage

microencapsulation of healing agents within a polymer matrix

has been increasingly employed for generating self-healing

polymer coatings.8 For instance, the coating on metallic

substrates can consist of a microencapsulated healing-agent and

phase-separated catalyst within a matrix. Damage to the surface

releases the healing-agent into the matrix where it comes into

contact with the catalyst and, subsequently, undergoes cross-

linking to form a new polymer layer at the site of damage.

Consequently, the metallic substrate remains protected from the

environment. Weder and colleagues recently reported on an

alternative approach in which optically healable supramolecular

metallopolymers could be mended through exposure to UV

radiation with a wavelength of 320–390 nm.9 The benefit of DA-

based healable systems over such other systems is the ability to

tether the dienophile and diene functional groups to any number

of polymer systems, as well as the ability to control the rDA

reaction via thermal and mechanical stimuli.

Previously, we reported the DA cycloaddition of dithioester

end-groups of RAFT polymers with suitable dienes.10 More

specifically, the hetero-DA (HDA) ligation technique has been

employed to generate block copolymers,11 star-shaped (co)

polymers,12,13 functionalized single-walled carbon nanotubes,14

and fullerenes,15 as well as surface modified microspheres,16 Si

wafers,17 and cellulose.18 The ability of a dithioester to undergo

a HDA cycloaddition reaction can be remarkably improved by

attaching electron deficient substituents (i.e., CF3,19 CN,20

COOR,21 PO(OR)2,22 pyridyl group,23 SO2R

24) to the a-positionof the thiocarbonyl group, which lowers the lowest unoccupied

molecular orbital (LUMO) energy of the thiocarbonyl group,

rendering it more reactive with an electron rich diene. Cyclo-

pentadiene is an example of a highly reactive, electron rich diene.

Cyclopentadienyl (Cp) functionalized polymers have been

demonstrated to undergo DA reactions with a variety of dien-

ophiles. Preliminary research in this area was realized in the

Barner-Kowollik laboratories and consisted of the ultrafast

polymer conjugation of a Cp-functionalized polymer with

several RAFT polymers.25 In addition, it was found that quan-

titative conversion of a variety of bromide end-functionalized

polymers into the corresponding Cp functional polymers was

possible at ambient temperature without any side product

formation when nickelocene was used as the substituting agent.26

A rapid bonding–debonding system consisting of a Cp-end-

capped polymer, synthesized by employing the facile nickel-

ocene-based end-group switching technique, and a pyridyl

dithioester tri-linker was recently developed and motivated the

search for additional DA-based crosslinkable systems.27

A molecule that is well known as a precursor to many indus-

trial polymeric products is 5-(isocyanato)-1-(isocyanatomethyl)-

1,3,30-trimethylcyclohexane (recognized by its trivial name

isophorone diisocyanate (IPDI)).28 Patel et al. studied the DA

reaction of a bisfuran derivative, containing a urethane linkage

formed by adding furfuryl alcohol to IPDI, with bismaleimides.29

The properties of the poly(urethane-imide)s obtained were

analyzed and it was found that these polymers undergo an irre-

versible aromatization at temperatures above 200 "C.30 Such an

irreversible aromatization is a considerable disadvantage when

employing these materials as thermally switchable systems. In

contrast, the DA reaction of a diene functionalized polymer

with a benzyl (diethoxyphosphoryl)dithioformate forms

a 3,6-dihydro-2H-thiopyran heterocycle, which was demon-

strated to be thermally stable below 100 "C and was shown to

have good stability in acidic and basic environments.31 Thus, the

DA reaction of a difunctional diene, based on isophorone dii-

socyanate, with a difunctional version of benzyl (diethoxy-

phosphoryl)dithioformate should result in a thermally

switchable polymeric material of significant industrial interest.

In the current study the DA and rDA reactions of a new bis-

diene, isophorone bis(sorbic carbamate) (IPDI-SA), with a new

bis-dithioester, 1,4-phenylenebis(methylene)bis((diethoxy-phos-

phoryl)methanedithioformate) (P-Di-linker), as the dienophile,

were analyzed. The AA–BB step-growth polymerization, depic-

ted in Scheme 1, and the debonding reaction of this system were

studied in detail. The DA reaction conditions, including the

monomer concentration, type of solvent, type and concentration

of Lewis acid, as well as reaction time, were varied in order to

analyze their influence on the reaction rate and to obtain opti-

mized conditions for the DA polymerization. It must be kept in

mind that a step-growth polymerization (in this case a poly-

addition) is governed by the Carothers equation, which reads

DPn# 1/(1! x) (where DPn is the degree of polymerization and x

is the conversion), given equal concentrations of monomers A

and B.32 Additionally, the correlation of the percentage of

debonding with rDA reaction time and temperature was assessed

and compared to ab initio quantum chemical calculations, which

were performed for the monomeric diene and dienophile struc-

tures utilized in the reaction. The results of the theoretical

calculations matched very well with the experimental results.

Therefore, future predictions of the debonding temperature for

a wide array of dienes and dienophiles, via ab initio quantum

chemical approaches, may be possible.

Experimental section

Materials

Isophorone bis(sorbic carbamate) (IPDI-SA) (Evonik Industries

AG), 1,4-bis(bromomethyl)benzene (97%, Aldrich), sodium

hydride (60% dispersion in mineral oil, Aldrich), tetrahydrofuran

(THF) (anhydrous,$99.9%, Acros), diethyl phosphate (>99.0%,

Fluka), carbon disulfide (anhydrous, $99.9%, Aldrich), zinc

Scheme 1 Step-growth polymerization of difunctional monomers based

on Diels–Alder chemistry.

Polym. Chem. This journal is ª The Royal Society of Chemistry 2011

Dow

nloa

ded

on 0

5 Ja

nuar

y 20

12Pu

blish

ed o

n 26

Oct

ober

201

1 on

http

://pu

bs.rs

c.or

g | d

oi:1

0.10

39/C

1PY

0035

6A

View Online

2-ethylhexanoate (97%, Aldrich), toluene (Fluka), 1,4-dioxane

(Aldrich), chloroform (Fluka), dimethyl sulfoxide (DMSO)

(Merck), dimethylformamide (DMF) (Fluka), N-methyl-2-pyr-

rolidone (NMP) (Aldrich), and acetonitrile (anhydrous, 99.8%,

Fluka) were used as received. Zinc chloride (Aldrich) was

vacuum dried and stored under nitrogen.

Characterization

1H Nuclear Magnetic Resonance (NMR) spectroscopy was per-

formed on a Bruker AM 250 spectrometer operating at 250 MHz

for hydrogen nuclei. All samples were dissolved in either CDCl3or DMSO-d6. The d-scale was calibrated to the internal standard

trimethylsilane (TMS, d # 0.00 ppm).13C NMR spectroscopy was performed on a Bruker 500 spec-

trometer operating at 125 MHz for carbon nuclei. Samples were

dissolved in CDCl3. The d-scale was calibrated to the residual

chloroform peak (d # 77.0 ppm).

Attenuated Total Reflectance-Infrared (ATR-IR) spectroscopy

was performed on a Bruker Vertex 80 FT-IR/NIR spectrometer

possessing a standard DTGS detector and a KBr beam splitter,

a heatable sample cell holder, as well as an ATR unit. The

parameters employed for the collection of each spectrum were

a resolutionof 4 cm!1, 128background scans, and64 sample scans.

Gel Permeation Chromatography (GPC) measurements were

performed on a Polymer Laboratories PL–GPC 50 Plus Inte-

grated System (Varian), comprising an autosampler, a PLgel

5 mm bead-size guard column (50 $ 7.5 mm), one PLgel 5 mmmixed E column (300 $ 7.5 mm), three PLgel 5 mm mixed C

columns (300 $ 7.5 mm), and a differential refractive index

detector, using THF as the eluant at 35 "C with a flow rate of

1 mL min!1. The SEC system was calibrated using linear poly

(styrene) (PS) standards ranging in molecular weight from 160 to

6 $ 106 g mol!1. The Mark–Houwink–Kuhn–Sakurada param-

eters utilized were K # 14.1 $ 10!5 dL g!1 and a # 0.70.33 The

peak molecular weight (Mp) and number average molecular

weight (Mn) were calculated relative to PS standards.

Mass Spectrometry was performed on a LXQ mass spectrom-

eter (ThermoFisher Scientific) equipped with an atmospheric

pressure ionization source operating in the nebulizer-assisted

electrospray mode, which was used in the positive ion mode. The

instrument was calibrated in the m/z range between 195–1822

using a standard comprising of caffeine, Met–Arg–Phe–Ala

acetate (MRFA), and a mixture of fluorinated phosphazenes

(Ultramark 1621) (Aldrich). Samples (c# 0.1–0.2 mg mL!1) were

dissolved in a 3 : 2 (v/v)mixture of THFandmethanol dopedwith

sodium trifluoroacetate (0.014 mg mL!1). All spectra were

acquired within them/z range of 150–2000 with a spray voltage of

5 kV and a capillary temperature of 275 "C. Nitrogen was used as

the sheath gas (flow: 45%ofmaximum) andheliumwasused as the

auxiliary gas (flow: 5% of maximum). Theoretical mass calcula-

tions were performed using the XCalibur 1.0 software.

Synthesis of 1,4-phenylenebis(methylene)bis

((diethoxyphosphoryl)methanedithioformate) (P-Di-linker) (1)

P-Di-linker 1 was synthesized according to a method adapted

from the procedure for the synthesis of benzyl (diethoxy-

phosphoryl)dithioformate described in the literature.34 A

solution of diethyl phosphite (5.3 mL, 41.2 mmol) in anhydrous

THF (20 mL) was added dropwise under nitrogen to a suspen-

sion of NaH (1.64 g, 41.2 mmol) in anhydrous THF (40 mL) in

a 2-neck flask equipped with a reflux condenser and a magnetic

stir bar. After the evolution of hydrogen had ceased, the mixture

was refluxed at 65 "C for 10 min. After cooling to ambient

temperature, the mixture was further chilled in a liquid nitrogen

bath. Subsequently, CS2 (12.26 mL, 203.6 mmol) was added

dropwise, after which the mixture was allowed to warm to

ambient temperature. Stirring was continued for an additional 30

min and then a solution of 1,4-bis(bromomethyl benzene) (5.44 g,

20.6 mmol) in anhydrous THF (40 mL) was added dropwise to

the reaction mixture. Stirring was continued at ambient

temperature for 3 h and then 200 mL of hexane was added and

the reaction mixture was filtered. The fuchsia filtrate was

collected and the solvent removed under reduced pressure. The

crude product was purified by column chromatography over

silica gel using hexane as the eluant initially to remove impurities

and subsequently using 100% ethyl acetate as the eluant to collect

the product 1. After removal of the solvent under reduced

pressure, the P-Di-linker 1 was obtained as a dark fuchsia solid

(60% yield). 1H NMR (250 MHz, DMSO-d6, 25"C): d (ppm) #

7.36 (s, 4H, ArH), 4.57 (s, 4H, –CH2S–), 4.21–4.08 (m, 8H,

–OCH2CH3), 1.30–1.21 (t, J # 7 Hz, 12H, –CH2CH3).13C NMR

(125 MHz, CDCl3, 25"C): d (ppm) # 228.28 (d, J # 173.75 Hz,

–S(CS)–), 133.89 (s, ArC), 129.83 (d, J# 32.5 Hz, ArC), 65.01 (d,

J # 6.25 Hz, –CH2S–), 40.23 (d, J # 2.5 Hz, –OCH2CH3), 16.50

(d, J # 6.25 Hz, –CH2CH3). ATR-IR: n # 2983 (aromatic C–H),

2908 (alkane C–H), 1639 (aromatic C]C), 1512 (aromatic C]C), 1472, 1387, 1242 (P]O), 1161 (C–O), 1082 (C]S), 1036 (C–

O), 1007 (P–OR), 978, 947 (P–OR), 750 cm!1. ESI-MS (+Na):

m/z calculated # 553.01 amu; m/z found # 553.12 amu.

Step-growth polymerization of P-Di-linker 1 and IPDI-SA 2

(3–13)

A typical polymerization procedure was as follows: both IPDI-

SA and P-Di-linker were each separately dissolved in acetonitrile

(0.25 mL) and were subsequently mixed in a 1 : 1 ratio with

respect to functional groups such that the concentration of each

monomer in the resulting solution was 1.8 mol L!1. Subse-

quently, 1.1 equivalents of zinc chloride (ZnCl2) were added to

the reaction mixture. After heating the mixture to 50 "C for 4 h,

the resulting viscous mixture was diluted in 1 mL chloroform,

washed with water to remove ZnCl2, and dried (85–96% yield).

This procedure for step-growth polymerization was performed in

the presence of THF, acetonitrile, toluene, 1,4-dioxane, chloro-

form, DMSO, NMP, and DMF to give the products 4, 7, 8, 9, 10,

11, 12, and 13, respectively. The variations of the reaction

conditions and the associated products are summarized in Tables

1 and 2. The polymer products were analyzed by GPC relative to

PS standards. Depending on the reaction conditions, the

following results were obtained: Mn # 1200–4800 g mol!1,Mp #2100–9600 g mol!1, and PDI # 1.41–2.60.

Retro Diels–Alder reaction of 7

A typical reaction procedure was as follows: 100 mg of the

polymer 7 was dissolved in 5 mL acetonitrile. While stirring, the

This journal is ª The Royal Society of Chemistry 2011 Polym. Chem.

Dow

nloa

ded

on 0

5 Ja

nuar

y 20

12Pu

blish

ed o

n 26

Oct

ober

201

1 on

http

://pu

bs.rs

c.or

g | d

oi:1

0.10

39/C

1PY

0035

6A

View Online

solution was heated in a pressure tube to 140 "C for 40 min. This

process was repeated for temperatures between 120 and 219 "C

and reaction times ranging from 10 to 160 min. After heating the

mixture for the allocated time, the reaction was rapidly termi-

nated by cooling with liquid nitrogen. A sample (0.1 mL) of the

rDA product was diluted in THF (0.4 mL) and analyzed by GPC

and mass spectrometry. Depending on the reaction conditions,

the rDA products were found to have Mn # 800–4400 g mol!1

and PDI # 1.42–2.14.

Theoretical procedures

Standard ab initio molecular orbital theory and density func-

tional theory calculations were performed using Gaussian 03.35

Calculations were performed at a high level of theory, as used

successfully in studies of several other classes of chemical reac-

tions.36–40 Further benchmarking of this methodology against

experiments for this particular application is carried out below.

Geometries of all species were optimized at the B3-LYP/6-31G

(d) level of theory and frequency calculations were also per-

formed at that level and scaled by their recommended factors.41

All calculations were based on the lowest energy conformers of

each species, as obtained via systematic conformational search-

ing at this level using our Energy-Directed Tree Search algo-

rithm.42 In cases where multiple Diels–Alder products were

possible, calculations were based on the lowest energy isomers

(i.e., the thermodynamic products). Full details of this screening

process are provided in the ESI†.

Having obtained the correct geometries, improved energies

were calculated using an ONIOM approximation to the

G3MP2//B3LYP method.38 G3MP2//B3LYP is a high-level

composite ab initio procedure that approximates QCISD(T)

calculations with a large triple zeta basis as the sum of the

lower-cost QCISD(T)/6-31G(d) calculations and a basis set

correction term, calculated at the MP2 level of theory. This

method has been shown to reproduce a large test set of ther-

mochemical data to within chemical accuracy (ca. 5 kJ mol!1),38

and this procedure or variants of it have been successfully

employed to predict the kinetics and thermodynamics of

a broad range of reactions, including computationally difficult

processes such as radical reactions,37 bond dissociation ener-

gies,38 redox potentials,39 and pKa values.40 Owing to the large

size of the molecules employed, in the present work an ONIOM

approximation to G3MP2//B3LYP was employed in which the

core of the reaction (designed to include the reaction center and

all primary substituents) was studied at G3MP2//B3LYP and

the remaining remote substituent effects were studied at

progressively lower levels of theory. For the model reaction

(14 + 15) the remote substituent effects were studied at the

MP2/GTMP2 Large level. It has been previously shown that,

provided all a and conjugated substituents are included in the

core, this technique can reproduce the corresponding full G3

(MP2)//B3LYP calculations to within chemical accuracy.43 For

the full reaction (1 + 2), a further ONIOM layer was utilized in

which the results for the model system were treated as the core,

and the remaining substituent effect was modeled with

B3-LYP/6-311 + G(3df,2p).

Gas-phase enthalpies (H), entropies (S), and Gibb’s free

energies (G) were calculated using the standard textbook

formulae, based on the statistical thermodynamics of an ideal gas

under the harmonic oscillator and rigid-rotor approximation.44,45

Calculations of partition functions were carried out using our in-

house program T-Chem.46,47 The free energies of each species in

solution were obtained as the sum of the corresponding gas-

phase free energy, the calculated free energy of solvation, and

a correction term, DnRTln (V), to take into account that the

solvation energy was computed for the passage from 1 atm (g) to

1 mol L!1 (soln).48 The solvation energies were modeled using the

COSMO-RS49 method at the BP/TZVP level of theory for which

it was parameterized. The COSMO-RS calculations were per-

formed in ADF,50 using the implementation of Pye et al.51 Unless

otherwise noted, the solvent was ethyl acetate, chosen to mimic

the polarity of the polymer present in bulk. The overall solution

phase free energy was then used to calculate the equilibrium

constants as for the gas-phase calculations:

Table 1 Influence of the Lewis acid (1.1 equivalents) and the monomerconcentration on the HDA reaction between 1 and 2 in THF at 50 "C for4 h

Product Reagent

Monomerconcentration/

mol L!1 Mpa/g mol!1 Mn

a/g mol!1 PDI

3 ZnCl2 0.2 4500 2100 2.194 ZnCl2 1.8 8500 3100 2.135 Zn(O2C8H15)2 0.2 2600 1200 1.676 — 1.8 3300 2200 1.88

a Relative to linear poly(styrene) standards.

Table 2 Influence of different organic solvents and reaction time on the HDA reaction between 1 (1.8 mol L!1) and 2 (1.8 mol L!1) in the presence of 1.1equivalents of zinc chloride at 50 "C. The PDI of each polymer is provided in Table S1 in the ESI† section

Product SolventMp

a/g mol!1

after 2 hMn

a/g mol!1

after 2 hMp

a/g mol!1

after 4 hMn

a/g mol!1

after 4 h

7 Acetonitrile 8100 3200 9600 48004 THF 3700 1700 8500 31008 Toluene 3000 2200 8100 30009 Dioxane 3800 2300 6000 230010 Chloroform 3800 2200 6700 270011 DMSO 2500 1900 2700 230012 NMP 2600 2500 2700 230013 DMF 2700 2300 2600 2200

a Relative to linear poly(styrene) standards.

Polym. Chem. This journal is ª The Royal Society of Chemistry 2011

Dow

nloa

ded

on 0

5 Ja

nuar

y 20

12Pu

blish

ed o

n 26

Oct

ober

201

1 on

http

://pu

bs.rs

c.or

g | d

oi:1

0.10

39/C

1PY

0035

6A

View Online

K(T) # (co)Dnexp (!DGsoln/RT) (1)

In this equation, the standard unit of concentration (co) has

a value of 1 mol L!1 for solution-phase free energies.

Results and discussion

Synthesis of the P-Di-linker 1

In the work presented here, a new diene–dienophile pairing is

developed and characterized for its potential application to self-

healing polymer systems that are based on DA and rDA chem-

istry. Additionally, it is demonstrated that theoretical ab initio

quantum chemistry predictions adhere well to experimentally

obtained rDA reaction data. The system under investigation is

based on the AA–BB step-growth polymerization of a bis-diene

and a bis-dienophile. The bis-dienophile 1 (P-Di-linker) was

designed to be difunctionalized with a phosphonate-substituted

dithioester due to the high reactivity of dithioesters as dien-

ophiles and the improved reactivity of these groups when

a-substituted with an electron-withdrawing group, such as

a phosphonate, sulfonyl, or pyridyl moiety. More specifically, the

DA chemistry of the monofunctional version of the difunctional

dithioester 1 was studied previously and found to be highly

reactive.31

The preparation of the difunctional linker was based on

a modified procedure for the synthesis of this monofunctional

dithioester, which can be found in the literature.34 In brief, in the

presence of sodium hydride, carbon disulfide, and diethyl phos-

phite a highly reactive nucleophilic intermediate, diethoxy-

phosphorylmethanedithioformate, is formed. Subjecting

dibromo-xylene to this intermediate resulted in the complete

substitution of both bromides to form the product 1 (Scheme 2).

The product was purified via column chromatography and 1H

NMR and ESI-MS analyses (Fig. 1) confirm that the RAFT

agent 1 was successfully formed.

HDA cycloaddition of the P-Di-linker 1 with diene 2

In an effort to determine the optimal reaction conditions for the

HDA reaction, the conditions utilized in the polymerization of 1

and 2 were varied to study the influence of the Lewis acid,

solvent, monomer concentration, and reaction time on the

formation of polymer (Tables 1 and 2). The resulting polymer

was analyzed viaGPC to determine its relative molecular weight.

Peak molecular weights (Mp) are presented to show the molec-

ular weight of the fraction with the longest polymer chains that

were achieved under the conditions studied. The bonds that are

formed via the HDA reaction are shown in Scheme 3. In order to

simplify the structure, not all possible isomers are depicted. In

the section pertaining to the theoretical predictions of the

Scheme 2 Synthesis of phosphonate dithioester di-linker 1.

Fig. 1 (a) 1H NMR spectrum of the difunctional linker 1 and (b) ESI-

MS spectrum of [1 + Na]+ determined experimentally (solid line) over-

lapping the theoretically predicted isotopic distribution for [1 + Na]+

(dotted line).

Scheme 3 The cycloaddition product (simplified structure of one

possible regio-/stereo-isomer is depicted) formed via the HDA reaction of

1 and 2.

This journal is ª The Royal Society of Chemistry 2011 Polym. Chem.

Dow

nloa

ded

on 0

5 Ja

nuar

y 20

12Pu

blish

ed o

n 26

Oct

ober

201

1 on

http

://pu

bs.rs

c.or

g | d

oi:1

0.10

39/C

1PY

0035

6A

View Online

debonding (i.e., rDA reaction) temperature the various possible

stereo-/regioisomers are discussed in detail.

Initially, the effect of the Lewis acid type and concentration

and monomer concentration on the DA reaction between 1 and 2

was assessed. It should be noted that in a number of studies

a Lewis acid, such as zinc chloride and trifluoroacetic acid, is

described as having catalyzed the HDA reaction.52 Although the

Lewis acid is referred to as a catalyst, in these studies an equi-

molar ratio of the catalyst and the monomer is added to the

reaction.52 Based on the significantly higher than catalytic

amounts of Lewis acid employed in these previous studies, and

the results found in the present study (see below), it is hypothe-

sized that the Lewis acid complexes with the phosphonate group

of compound 1, which results in a more electron-withdrawing

group, affording a more reactive thiocarbonyl heterodienophile.

Therefore, the true role of the Lewis acid is thought to be as

a reagent rather than a catalyst. Thus, in the current study, the

Lewis acid is referred to as a reagent.

Zinc chloride was selected as the Lewis acid of choice based on

literature reports suggesting that it enhances DA chemistry effi-

ciency.52 Comparing the molecular weight of the HDA products

4 and 6 formed in the presence and absence of zinc chloride,

respectively, (Table 1) reveals that adding a Lewis acid indeed

increases the reaction rate. At the same concentration of

monomers (1.8 mol L!1), the HDA reaction in the presence of

zinc chloride resulted in a polymer (4) with aMp of 8500 g mol!1,

which is 2.6 times larger than the polymer formed under the same

reaction conditions, but in the absence of zinc chloride (6; Mp #3300 g mol!1). In addition, the type of Lewis acid was found to

influence the efficacy of the HDA reaction. Zinc chloride resulted

in improved polymer conversion compared to that achieved in

the presence of zinc 2-ethylhexanoate. Specifically, at the same

concentration of monomers (0.2 mol L!1), the Mp of polymer 3

(4500 g mol!1), obtained in the presence of zinc chloride, is

almost two times greater than the Mp of polymer 5 (2600 g

mol!1), obtained in the presence of zinc 2-ethylhexanoate

(Fig. 2a). The lowmolecular weight of the DA product 5 suggests

that zinc 2-ethylhexanoate does not change the reactivity of the

dienophile significantly. An additional—yet not overly

surprising—factor that was found to affect the polymerization

reaction is the concentration of the monomers. As is shown in

Table 1, the Mp of the DA product almost doubled from

4500 g mol!1 (polymer 3) to 8500 g mol!1 (polymer 4) in the same

reaction time and at the same temperature by increasing the

concentration of the monomers 1 and 2 from 0.2 to 1.8 mol L!1.

Unfortunately, the effects of monomer concentrations greater

than 1.8 mol L!1 could not be explored because of the limited

solubility of IPDI-SA (i.e., 1.8 mol L!1 was the highest concen-

tration that could be achieved).

The HDA cycloaddition reaction was also performed in

various organic solvents at 50 "C for 4 h using zinc chloride as

a reagent to enhance the cycloaddition. Solvents were found to

have a significant effect on the reaction rate of the cycloaddition

of the diene/dienophile pair studied (Table 2). The organic

solvents utilized for the HDA reaction can be divided into three

classes according to their influence on the HDA reaction and,

therefore, the polymer molecular weight achieved. The HDA

reactions carried out independently in DMSO, NMP, and DMF

resulted in products (11–13) of relatively low Mp after a 4 h

reaction time at 50 "C (i.e., Mp z 2600 g mol!1). Considerably

higherMp values, between 6000 and 7000 g mol!1, were achieved

when the polymer was synthesized in dioxane (9) and chloroform

(10) after a 4 h reaction time at 50 "C. In contrast, performing the

reaction in toluene (8), THF (4), or acetonitrile (7) for the same

reaction time and temperature gave rise to a polymer with a peak

molecular weight greater than 9000 g mol!1. Specifically, aceto-

nitrile resulted in the polymer (7) with the greatest peak molec-

ular weight (9600 g mol!1). When contemplating the implication

of the molecular weight, it must be kept in mind that a step-

growth polymerization, such as DA-based polymerization, is

governed by the Carothers equation. Employing the Carothers

equation, the theoretically calculatedMn for 99.9% conversion is

474 000 g mol!1, whereas if the conversion decreases by just

a small fraction to either 99% or 95%, then a molecular weight of

47 400 or 9480 g mol!1, respectively, is expected. Thus, the

highest conversion achieved (polymer 7,Mn # 4800 g mol!1) can

be estimated to be approximately 90%. It was concluded that

Fig. 2 (a) Gel permeation chromatograms of polymers 4 (1.1 equiva-

lents ZnCl2) and 6 (without ZnCl2) obtained after HDA reaction of 1

(1.8 mol L!1) with 2 (1.8 mol L!1) in THF at 50 "C for 4 h and (b) gel

permeation chromatograms of polymers 7 (in acetonitrile), 10 (in

dioxane), 13 (in DMF) after HDA reaction at 50 "C for 4 h with 1.1

equivalents of ZnCl2 and a monomer concentration of 1.8 mol L!1.

Polym. Chem. This journal is ª The Royal Society of Chemistry 2011

Dow

nloa

ded

on 0

5 Ja

nuar

y 20

12Pu

blish

ed o

n 26

Oct

ober

201

1 on

http

://pu

bs.rs

c.or

g | d

oi:1

0.10

39/C

1PY

0035

6A

View Online

DMSO, NMP, and DMF are poor solvents for the HDA reac-

tion, whereas acetonitrile is an excellent solvent for this reaction.

However, as of yet no specific trend associated with the effect of

solvent properties on the HDA reaction has been established.

Retro Diels–Alder reaction of 7

The rDA reaction of polymer 7 was studied at various temper-

atures and reaction times to assess the onset temperature of

significant rDA reaction and the efficiency of the debonding step.

Polymer 7 was chosen, among all HDA products, as the starting

material for these studies as it was found to have the highest

number average molecular weight (Mn # 4800 g mol!1).

Regarding the GPC analysis of the rDA products, it is very

important to establish a molecular weight calculation method

and apply it consistently from sample to sample because the

calculated Mn is strongly dependent on the molecular weight

range selected for the characterization of the GPC chromato-

gram. In order to establish the calculation method it is important

to consider that in a step-growth polymerization, where 100%

conversion is unattainable, there are polymer chains of different

lengths and, therefore, of widely variable molecular weights. In

the rDA reaction the small oligomers are of particular interest.

An increase in debonding results in an increase in the low

molecular weight polymers, which implies that the molecular

weight range that should be considered for the calculation ofMn

should include the low molecular weight region of the GPC

chromatogram. Therefore, in order to determine the changes in

Mn relative to the Mn of the starting polymer 7 and each rDA

polymer sample, the molecular weight range used for the calcu-

lations was set to include all molecular weight peaks occurring in

the GPC chromatogram and kept constant for all samples having

undergone rDA reactions.

For the purposes of calculating the percentage of debonding of

the rDA products, it was assumed that the conversion achieved

in the DA reaction is equal to the percent bonding. Thus, the

percentage of unconverted monomers is equal to percentage of

debonding. Utilizing the Carothers equation, DPn # 1/(1 ! x),

polymer conversion, x, can be calculated from the degree of

polymerization, DPn. In order to simplify the calculation of DPn,

the average molecular weight of the two monomers ( "m # 474 g

mol!1) was employed because it is complicated to determine

which monomer is debonded as the molecular weight decreases.

It follows that the percentage of debonding could be calculated

using eqn (2), where DPzn#Mz/ "m is the degree of polymerization,

z is the reaction time, and Mz is the number average molecular

weight of the rDA product at time z. All rDA conditions and DPn

values are summarized in Table S3 provided in the ESI† section.

Percent debonding # 1

DP zn

$ 100% (2)

In an effort to ascertain if the rDA reaction was even possible,

the polymer 7 was initially subjected to 219 "C, which is the

highest temperature that is attainable in a silicon oil bath. Fig. 3

depicts the decrease in Mn from 4800 to 800 g mol!1 of 7 after

30 min at 219 "C, as determined by GPC. ESI-MS results addi-

tionally suggest that the rDA reaction was successful since peaks

corresponding to the molecular weight of the monomer and

small oligomers are observed. It is noteworthy to mention that

some oligomers have zinc chloride incorporated (e.g., peak D in

Fig. 3b). Zn-incorporation is assumed to result from the

complexation of the Lewis acid with the phosphonate group, as

discussed at the beginning of the HDA cycloaddition section.

The experimental and theoretical m/z values are summarized in

Table S2 provided in the ESI† section. It was also concluded that

the polymer underwent minimal decomposition during the rDA

reaction since the major ESI-MS peaks corresponded to the

molecular weight of the DA polymer building blocks.

To achieve an improved understanding of the rDA reaction,

the time and temperature dependence of the debonding reaction

was analyzed at 120, 140, and 160 "C for reaction times ranging

from 10 min to 160 min. The extent of the rDA reaction was

assessed from the Mn, which was determined via GPC, of the

rDA product as a function of time at the selected temperatures

(Fig. 4). Since the polymer molecular weight values measured are

only calculated relative to poly(styrene) standards, the molecular

weight obtained cannot be considered absolute. However, as

previously explained, the percentage of debonding can be

calculated for the different reaction conditions, using eqn (2), to

represent the extent of the rDA reaction. First, eqn (2) was

utilized to determine the percentage of debonding of the starting

material, polymer 7, which has aMn of 4800 g mol!1. Given that

polymer 7 has a degree of polymerization close to 10, it was

determined that 10% of the monomers were not bonded.

Applying eqn (2) to the molecular weight of each of the rDA

products revealed that the decrease in the molecular weight of the

DA polymer strongly depends—as expected—on the rDA reac-

tion conditions.

As depicted in Fig. 4, the rate of the debonding reaction

depends on the reaction temperature. In addition, after 1 h of

reaction time the percentage of polymer debonding begins to

plateau at a value that varies depending on the reaction

temperature. The percentage of debonding seems to plateau at

%20% at 120 "C and at %30% at 160 "C. Subsequently, the

correlation between the percentage of debonding achieved and

the reaction temperature was studied in more detail. For a fixed

rDA reaction time of 30 min the percentage of the debonded

polymer was determined at various temperatures. At tempera-

tures lower than 100 "C the rDA reaction was barely observed to

proceed. At temperatures ranging from 100 to 200 "C a sharp

increase in the rate of debonding was observed. Increasing the

temperature beyond 200 "C, however, did not appear to enhance

the rDA reaction rate further (Fig. 5). The curve obtained in

Fig. 5 has a typical S-shape, similar to the reaction rate curves

determined theoretically in the following section.

Theoretical predictions of bonding and debonding temperature

The DA reaction is an equilibrium reaction of bonding and

debonding of the diene and dienophile (see eqn (3)). The bonding

process (the forward DA reaction) is exothermic and mostly

proceeds at low temperatures. However, as the reaction is

exentropic, the DA reaction becomes progressively less favorable

as the temperature increases. Therefore, the most interesting

aspect of the theoretical predictions is the temperature at which

the reverse process, the endothermic rDA reaction, becomes

significant. To answer this question, it is necessary to establish

how the equilibrium constant changes as a function of

This journal is ª The Royal Society of Chemistry 2011 Polym. Chem.

Dow

nloa

ded

on 0

5 Ja

nuar

y 20

12Pu

blish

ed o

n 26

Oct

ober

201

1 on

http

://pu

bs.rs

c.or

g | d

oi:1

0.10

39/C

1PY

0035

6A

View Online

temperature, and it is also necessary to have a sense for how the

extent of cross-linking depends on the equilibrium constant itself.

Firstly, this latter question is addressed. For a DA reaction of the

form

A + B # C (3)

let [A], [B], and [C] refer to the molar concentrations of the

respective functional groups (diene, dienophile, linkage). If one

assumes that the equilibrium constant for each individual DA

reaction in the process is constant—irrespective of the degree of

bonding—then an equilibrium expression relating the respective

functional groups can be written as:

Fig. 3 (a) Gel permeation chromatogram depicting the decrease in

molecular weight of 7 (20 mg mL!1) after the rDA reaction in acetonitrile

at 219 "C after 30 min and (b) ESI mass spectrum corresponding to the

rDA product. The labels for the individual peak assignments in (b) (A–H)

are employed in the ESI† section for purposes of peak assignment. A:

441.32 # IPDI-SA + Na+; *: 621.46 # impurity of IPDI-SA; B: 829.16 #P-Di-linker + 2(ZnCl2) + Na+ + 4H+; C: 971.20 # IPDI-SA +

P-Di-linker + Na+; D: 1109.12 # IPDI-SA + P-Di-linker + ZnCl2 +

Na+ + 4H+; E: 1243.36 # IPDI-SA + P-Di-linker + 2(ZnCl2) + Na+; F:

1389.32 # 2(IPDI-SA) + P-Di-linker + Na+; G: 1569.32 # 2(IPDI-SA) +

P-Di-linker + ZnCl2 + acetonitrile + Na+ + 5H+; H: 1919.28 # 2(IPDI-

SA) + 2(P-Di-linker + Na+).

Fig. 4 Decrease of Mn with increasing reaction time (a) and the corre-

lation of percentage of debonding of polymer with reaction time (b)

for the rDA reaction of 7 (20 mg mL!1) in acetonitrile at 120, 140,

and 160 "C.

Fig. 5 Percentage of debonded monomers as a function of the temper-

ature (25, 120, 140, 160, 180 and 219 "C) of the rDA reaction of 7

(20 mg mL!1) in acetonitrile after 30 min reaction time.

Polym. Chem. This journal is ª The Royal Society of Chemistry 2011

Dow

nloa

ded

on 0

5 Ja

nuar

y 20

12Pu

blish

ed o

n 26

Oct

ober

201

1 on

http

://pu

bs.rs

c.or

g | d

oi:1

0.10

39/C

1PY

0035

6A

View Online

K # &C'&A'&B'

# &C'!&A'0!&C'

"!&B'0!&C'

" (4)

It should be noted here that since reagents A and B are both

difunctional, [A]0 is twice the initial concentration of reagent A

and [B]0 is twice the starting concentration of reagent B, and that

initially there is no polymer. In the special case that [A]0 # [B]0,

this expression can be solved for [C] as follows:

K # &C'!&A'0!&C'

"2

K&A'20 !!2K &A'0(1

"&C' ( K&C'2# 0

&C' #!2K&A'0(1

")

###############################################!2K &A'0(1

"2!4K2&A'20q

2K

#!2K&A'0(1

")

#####################4K &A'0(1

p

2K(5)

Since [C] cannot exceed [A]0 (due to the law of conservation of

mass), the following solution is obtained.

&C' #!2K &A'0(1

"!

#####################4K &A'0(1

p

2K(6)

Finally, [C] can be expressed as a percentage of bonding as

follows:

Percent bonding # &C'&A'0

$ 100%

#!2K &A'0(1

"!

#####################4K&A'0(1

p

2K &A'0$ 100% (7)

It is worth noting that due to chain length effects on the

entropy, K is unlikely to be constant for every individual reaction

in the process. In addition, increasing viscosity is likely to prevent

an equilibrium concentration of cross-links from forming when

the equilibrium concentration of C is high. Nonetheless, the

expression and the use of a single K value (as calculated for the

first reaction of 1 and 2) should be useful for establishing the

conditions required for the complete debonding of the cross-

linked polymer (e.g., the point at which, say, 90% of the chains

are debonded). Taking polymer 7 as an example, this polymer

was prepared using [A]# [B]# 1.8 mol L!1, which translates into

a molar concentration of active groups of [A]0 # [B]0 #3.6 mol L!1. Using eqn (7), the percentage of debonding can be

calculated as a function of K (Fig. 6). From this graph, it can be

observed that 90% debonding corresponds to a log K value

of !1.5.

To study the temperature dependence of the equilibrium

constant, ab initio molecular orbital theory was used to calculate

the gas-phase equilibrium constant of a small model reaction as

a functionof temperature, as shown in Scheme 4.TheDAreaction

can result in a number of different isomers (16–23) (Scheme 5).

Equilibrium constants were calculated for all of these and the

lowest energy structure (16) was selected for further consider-

ation. In other words, the thermodynamic calculations are based

on the thermodynamic products of the DA reaction. Having

obtained the gas-phase results, the COSMO-RS procedure was

subsequently employed to convert these gas-phase values to the

condensed phase. Finally—since entropy provides the driving

force for the rDAreaction—the gas-phase and solution-phase free

energies of the first step in the actual monomer-linking process

(i.e., the reaction of 1 and 2) were additionally calculated to ensure

that the effects of chain lengthwere correctly taken into account in

the calculations. Table 3 depicts the enthalpies, entropies, Gibbs’

free energies, and the corresponding equilibrium constants of

both the model reaction (14 + 15) and the step 1 of the full system

(1 + 2) in the gas phase as a function of temperature. To make the

results more indicative of the real system, these gas phase results

were subsequently converted to the solution phase using calcu-

lated free energies of solvation in ethyl acetate, chosen to reflect

the polarity of the bulk polymer environment. The use of ‘‘solu-

tion-phase’’ data to model the reactions in a polymer film is

expected to be reasonable at the higher temperatures where the

rDA reaction is close to complete. At that point, the size of the

molecules is close to the size of those employed in the actual ab

initio calculations on the full reaction, the molecules themselves

are relatively mobile, and the assumptions used to derive eqn (7)

are expected to be valid (i.e., equilibrium is achieved, and each of

the relevant microscopic equilibrium constants are likely to be

similar to one another).

Employing the data in Table 3, the solution-phase equilib-

rium constants are plotted as a function of temperature for

Fig. 6 Percentage of debonding of functional groups as a function of the

equilibrium constant K, for starting concentrations of the diene and

dienophile functional groups of 3.6 mol L!1 each (i.e., corresponding to

a 1.8 mol L!1 concentration of the diene and dienophile monomers),

calculated from eqn (7).

Scheme 4 The reaction of 14 and 15, as computationally studied, was

the model DA reaction for the cycloaddition of 1 with 2.

This journal is ª The Royal Society of Chemistry 2011 Polym. Chem.

Dow

nloa

ded

on 0

5 Ja

nuar

y 20

12Pu

blish

ed o

n 26

Oct

ober

201

1 on

http

://pu

bs.rs

c.or

g | d

oi:1

0.10

39/C

1PY

0035

6A

View Online

both the model reaction and step 1 of the full reaction (Fig. 7).

From Fig. 7, it can be deduced—as expected—that the DA

equilibrium constant decreases as a function of temperature

and the reverse DA (i.e., rDA) reaction becomes progressively

more favorable. It can also be observed that at any given

temperature, the equilibrium constant of the full reaction is

approximately 2 orders of magnitude lower than for the small

model system. Such a difference is largely caused by entropic

effects, and is due to the fact that when two large molecules of

similar sizes combine to form a single molecule a loss of

translational entropy occurs that scales with the mass of the

species involved. For such a bimolecular association reaction in

an ideal gas, the mass effect on the translational entropy is

given by the following simple formula:

DScorrection #3R

2ln

$*M1 (M2+M1*M2*

*M1*(M2*+M1M2

%(8)

where R is the universal gas constant (8.314472 J mol!1 K!1),

M1 andM2 are the actual molecular weights of the two reactants,

whilst their sum is the molecular weight of the product; the

corresponding starred terms are the masses of the lower molec-

ular weight model species. The ideal gas formula (eqn (8)) slightly

overestimates the chain length effect compared with that in

solution, where molecules have less independent movement to

start with, and is thus countered to a small extent by the solvation

energy contributions. Nonetheless, the observed chain length

effects imply that the representation of the various microscopic

cross-linking reactions by a single equilibrium constant is likely

to be unrealistic at high degrees of monomer-linking. Thus, the

theoretical prediction focuses on identifying the point at which

the majority of chains have become debonded (e.g., 90% or

more). Combining the data in Fig. 6 and 7, one can plot the

percentage of debonded molecules as a function of temperature

for the first step of the full reaction (Fig. 8).

The theoretically calculated percentage of debonding utilizes

concentrations for the calculation, while the experimentally

obtained percentage of debonding is calculated from molecular

weights and the degree of polymerization utilizing the Carothers

Scheme 5 Possible isomers of 16 and their relative energies (kJ mol!1).

Table 3 Thermodynamics of the reaction of IPDI-SA and P-Di-linkera

T/"C

Model system (14 + 15)b Step 1 in the full system (1 + 2)c

DH gas DS gas DG gas DDG solvlog Ksoln DH gas DS gas DG gas DDG solv

log Ksoln

0 !122.2 !243.5 !55.6 6.0 10.9 !123.9 !288.9 !45.0 5.1 9.025 !122.2 !243.6 !49.6 5.3 9.1 !124.0 !289.2 !37.8 4.3 7.350 !122.2 !243.5 !43.5 4.6 7.7 !124.1 !289.3 !30.6 3.6 5.775 !122.1 !243.4 !37.4 3.9 6.4 !124.0 !289.3 !23.3 2.9 4.4100 !122.1 !243.4 !31.3 3.3 5.3 !124.0 !289.1 !16.1 2.3 3.3150 !122.0 !243.1 !19.1 2.1 3.5 !123.8 !288.7 !1.7 1.1 1.5200 !121.4 !241.6 !7.1 1.1 2.1 !123.5 !288.0 12.8 0.0 0.0250 !120.9 !240.6 5.0 0.1 0.9 !123.1 !287.2 27.1 !1.0 !1.2300 !120.4 !239.7 17.0 !0.9 !0.1 !122.6 !286.4 41.5 !1.9 !2.2

a Calculated with an ONIOM approximation to the G3(MP2)//B3-LYP/6-31G(d) level of theory using COSMO-RS/BP/TZVP solvation energies (seetext). b Reaction of 14 and 15 is shown in Scheme 4. c Reaction of 1 and 2 is shown in Scheme 3.

Polym. Chem. This journal is ª The Royal Society of Chemistry 2011

Dow

nloa

ded

on 0

5 Ja

nuar

y 20

12Pu

blish

ed o

n 26

Oct

ober

201

1 on

http

://pu

bs.rs

c.or

g | d

oi:1

0.10

39/C

1PY

0035

6A

View Online

equation. A comparison of the experimental and theoretical

results is depicted in Fig. 8. Upon inspection of Fig. 8, it can be

deduced that the polymer becomes 60% debonded at around

220 "C and this is indeed in excellent agreement with the exper-

imental data in Fig. 5, where the polymer becomes 59% debon-

ded at around 219 "C. The discrepancy between the theoretical

predictions and the experimental data at lower temperatures is

expected and stems from a combination of factors. Firstly,

eqn (8) shows clearly that chain length effects on the equilibrium

constant are expected; these are not taken into account in the

quantum chemical calculations, which are based on the reaction

of the monomeric species (1 and 2). As a result, the equilibrium

constants are only likely to be indicative of the experimental

system at high percentage values of debonding (i.e., high

temperature), when the reaction of monomeric species is the

dominant process. Secondly, as noted above, the derivation of

eqn (7) assumes that these chain length effects are absent (and

thus all of the microscopic equilibrium constants are equal), and

that the system is at thermal equilibrium (rather than being

kinetically limited by viscosity and other effects). These

assumptions, and hence the relationship between the calculated

equilibrium constants and the measured percentage of debond-

ing data, are also likely to hold true only at high percentage

values of debonding. Therefore, it is no surprise that in Fig. 8

there is a discrepancy between the theoretical data and the

experimental data at low temperatures, whereas the theoretical

and experimental debonding values agree well at high tempera-

tures, where molecules have low molecular weight.

Conclusions

The step-growth polymerization, based on DA chemistry, of bis-

diene isophorone bis(sorbic carbamate) with the bis-RAFT agent

1,4-phenylenebis(methylene)bis((diethoxy-phosphoryl)methane-

dithioformate) was studied in detail. Solvents were found to have

a considerable influence on the reaction rate. Under optimized

conditions (i.e., 1.8 M of each monomer in acetonitrile with 1.1

equivalents of zinc chloride at 50 "C for 4 h) a polymer with

a peak molecular weight of 9600 g mol!1 relative to poly(styrene)

standards was obtained. Employing this polymer, a series of rDA

reactions, where the reaction temperature and time were varied,

were carried out in order to study the debonding process. A

maximum of 59% debonding was achieved at 219 "C after

30 min. It was found that ab initio quantum chemical calculations

of the debonding process of the system studied are in excellent

agreement with the data obtained experimentally, thus a basis

has been established for the future employment of ab initio

predictions of bonding/debonding temperatures of different

diene–dienophile pairings in the design of self-healing materials.

Acknowledgements

C. B.-K. is grateful for continued support from Evonik Indus-

tries AG. In addition, C. B.-K. acknowledges funding from the

Karlsruhe Institute of Technology (KIT) in the context of the

Excellence Initiative for leading German universities, the

German Research Council (DFG) as well as the Ministry of

Science and Arts of the state of Baden-W!uttemberg. M. L. C.

gratefully acknowledges generous allocations of supercomputing

time on the National Facility of the Australian National

Computational Infrastructure, support from the Australian

Research Council (ARC) under its Centers of Excellence

program, and receipt of an ARC Future Fellowship.

References and notes

1 O.Diels andK. Justus Alder, Justus Liebigs Ann. Chem., 1928, 460, 98.2 K. C. Nicolaou, S. A. Snyder, T. Montagnon andG. Vaailikogiannakis, Angew. Chem., Int. Ed., 2002, 41, 1668.

3 K. M.Wiggins, J. A. Syrett, D. M. Haddleton and C. W. Bielawski, J.Am. Chem. Soc., 2011, 133, 7180.

4 (a) A. Sanyal, Macromol. Chem. Phys., 2010, 211, 1417; (b)E. B. Murphy and F. Wudl, Prog. Polym. Sci., 2010, 35, 223; (c)X. Chen, M. A. Dam, K. Ono, A. Mal, H. Shen, S. R. Nutt,K. Sheran and F. Wudl, Science, 2002, 295, 1698.

5 (a) S. D. Bergmann and F. Wudl, J. Mater. Chem., 2008, 18, 41; (b)A. M. Peterson and G. R. Palmese, Reversible Diels–AlderCycloaddition for the Design of Multifunctional NetworkPolymers, in Click Chemistry for Biotechnology and MaterialsScience, ed. J. Lahann, John Wiley & Sons, Ltd, Chichester, UK,2009, DOI: 10.1002/9780470748862.ch9; (c) A. Gandini, Polym.

Fig. 7 Temperature dependence of the equilibrium constant for the DA

reactionof themodel system(14+15) andfirst stepof the full reaction (1+2).

Fig. 8 Temperature dependence of the percentage of debonding for the

DA reaction as determined by ab initio quantum chemistry calculations

for the first step of the full reaction (1 + 2) (diamonds) and experimental

data (triangles).

This journal is ª The Royal Society of Chemistry 2011 Polym. Chem.

Dow

nloa

ded

on 0

5 Ja

nuar

y 20

12Pu

blish

ed o

n 26

Oct

ober

201

1 on

http

://pu

bs.rs

c.or

g | d

oi:1

0.10

39/C

1PY

0035

6A

View Online

Chem., 2010, 1, 245; (d) B. Voit, New J. Chem., 2007, 31, 1139; (e)A. K.-Y. Jen, Polym. Prepr., 2006, 47, 988.

6 (a) A. A. Kavitha and N. K. Singha, J. Polym. Sci., Part A: Polym.Chem., 2007, 45, 4441; (b) M. Watanabe and N. Yoshie, Polymer,2006, 47, 4946; (c) A. M. Peterson, R. E. Jensen and G. R. Palmese,ACS Appl. Mater. Interfaces, 2010, 2, 1141; (d) J. R. McElhanon andD. R. Wheeler, Org. Lett., 2001, 3, 2681; (e) Y. Zhang,A. A. Broekhuis and F. Picchioni, Macromolecules, 2009, 42, 1906;(f) J. R. McElhanon, T. Zifer, S. R. Kline, D. R. Wheeler,D. A. Loy, G. M. Jamison, T. M. Long, K. Rahimian andB. A. Simmons, Langmuir, 2005, 21, 3259; (g) N. W. Polaske,D. V. McGrath and J. R. McElhanon, Macromolecules, 2010, 43,1270; (h) H. S. Patel, B. P. Patel and D. B. Patel, Int. J. Polym.Mater., 2009, 58, 625; (i) H. S. Patel, B. P. Patel and D. B. Patel,Polym.-Plast. Technol. Eng., 2010, 49, 394; (j) H. S. Patel andH. S. Vyas, Eur. Polym. J., 1991, 27, 93; (k) B. Jiang, J. Hao,W. Wang, L. Jiang and X. Cai, Eur. Polym. J., 2001, 37, 463; (l)H. S. Patel, H. D. Patel andH. S. Vyas,Macromol. Rep., 1994, 31, 511.

7 (a) P. J. Costanzo, J. D. Demaree and F. L. Beyer,Langmuir, 2006, 22,10251; (b) A.M. Peterson, R. E. Jensen andG. R. Palmese,ACSAppl.Mater. Interfaces, 2009, 1, 992; (c) A. M. Peterson and G. R. Palmese,Reversible Diels–Alder Cycloaddition for the Design of MultifunctionalNetwork Polymers, Wiley, New York, 2009.

8 (a) S. R. White, N. R. Sottos, P. H. Geubelle, J. S. Moore,M. R. Kessler, S. R. Sriram, E. N. Brown and S. Viswanathan,Nature, 2001, 409, 794; (b) S. H. Cho, S. R. White and P. V. Braun,Adv. Mater., 2009, 21, 645.

9 M. Burnworth, L. Tang, J. R. Kumpfer, A. J. Duncan, F. L. Beyer,G. L. Fiore, S. J. Rowan and C. Weder, Nature, 2011, 472, 334.

10 (a) A. J. Inglis and C. Barner-Kowollik, Macromol. Rapid Commun.,2010, 31, 1247; (b) H. Willcock and R. K. O’Reilly, Polym. Chem.,2010, 1, 149.

11 S. Sinnwell, A. J. Inglis, T. P. Davis, M. H. Stenzel and C. Barner-Kowollik, Chem. Commun., 2008, 17, 2052.

12 A. J. Inglis, S. Sinnwell, T. P. Davis, M. H. Stenzel and C. Barner-Kowollik, Macromolecules, 2008, 41, 4120.

13 (a) S. Sinnwell, A. J. Inglis, T. P. Davis, M. H. Stenzel and C. Barner-Kowollik, Macromol. Rapid Commun., 2008, 29, 1090; (b) G. Hart-Smith, M. Lammens, F. E. Du Prez, M. Guilhaus and C. Barner-Kowollik, Polymer, 2009, 50, 1986.

14 N. Zydziak, C. H!ubner, M. Bruns and C. Barner-Kowollik,Macromolecules, 2011, 44, 3374.

15 L. Nebhani and C. Barner-Kowollik, Macromol. Rapid Commun.,2010, 31, 1298.

16 L. Nebhani., S. Sinnwell, A. J. Inglis, T. P. Davis, M. H. Stenzel,C. Barner-Kowollik and L. Barner, Macromol. Rapid Commun.,2008, 29, 1431.

17 L. Nebhani, P. Gerstel, P. Atanasova, M. Bruns and C. Barner-Kowollik, J. Polym. Sci., Part A: Polym. Chem., 2009, 47, 7090.

18 A. S. Goldmann, T. Tischer, L. Barner, M. Bruns and C. Barner-Kowollik, Biomacromolecules, 2011, 12, 1137.

19 C. Portella, Y. G. Shermolovich and O. Tschenn, Bull. Soc. Chim. Fr.,1997, 134, 697.

20 (a) D.M.Vyas andG.W.Hay,Can. J. Chem., 1971, 49, 3755–3758; (b)D.M. Vyas and G.W. Hay, J. Chem. Soc., Perkin Trans. 1, 1975, 180.

21 (a) E. Vedejs, M. J. Arnost, J. M. Dolphin and J. Eustache, J. Org.Chem., 1980, 45, 2601; (b) G. W. Kirby, A. W. Lochead andS. Williamson, J. Chem. Soc., Perkin Trans. 1, 1996, 977.

22 (a) B. Heuze, R. Gasparova, M. Heras and S. Masson, TetrahedronLett., 2000, 41, 7327; (b) M. Heras, M. Gulea and S. Masson,Chem. Commun., 2001, 611; (c) M. Heras, M. Gulea, S. Massonand C. Philouze, Eur. J. Org. Chem., 2004, 160.

23 R. Bastin, H. Albadri, A. C. Gaumont and M. Gulea, Org. Lett.,2006, 8, 1033.

24 (a) J. A. Boerma, N. H. Nilsson and A. Senning, Tetrahedron, 1974,30, 2735; (b) I. El-Sayed, O. M. Ali, A. Fischer and A. Senning,Heteroat. Chem., 2003, 14, 170.

25 (a) A. J. Inglis, S. Sinnwell, M. H. Stenzel and C. Barner-Kowollik,Angew. Chem., Int. Ed., 2009, 48, 2411; (b) A. J. Inglis,M. H. Stenzel and C. Barner-Kowollik, Macromol. Rapid Commun.,2009, 30, 1792.

26 A. J. Inglis, T. Paul!ohrl and C. Barner-Kowollik, Macromolecules,2010, 43, 33.

27 A. J. Inglis, L. Nebhani, O. Altintas, F. G. Schmidt and C. Barner-Kowollik, Macromolecules, 2010, 43, 5515–5520.

28 Encyclopedia of Polymer Science and Engineering, Wiley, New York,2nd edn, 1988.

29 H. S. Patel, A. B. Mathur and I. S. Bhardwaj, J. Macromol. Sci., PartA: Pure Appl. Chem., 1995, 32, 2025.

30 V. Gaina and C. Gaina, Polym.-Plast. Technol. Eng., 2002, 41, 523.31 S. Sinnwell, C. V. Synatschke, T. Junkers, M. H. Stenzel and

C. Barner-Kowollik, Macromolecules, 2008, 41, 7904.32 W. H. Carothers, Trans. Faraday Soc., 1936, 32, 39.33 C. Strazielle, H. Benoit and O. Vogl, Eur. Polym. J., 1978, 14, 331.34 M. Laus, R. Papa, K. Sparnacci, A. Alberti, M. Benaglia and

D. Macciantelli, Macromolecules, 2001, 34, 7269.35 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria,

M. A. Robb, J. R. Cheeseman, J. A. Montgomery, Jr., T. Vreven,K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi,V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega,G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota,R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda,O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian,J. B. Cross, C. Adamo, J. Jaramillo, R. Gomperts,R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli,J. W. Ochterski, P. Y. Ayala, K. Morokuma, G. A. Voth,P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich,A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick,A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz,Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov,G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin,D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng,A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson,W. Chen, M. W. Wong, C. Gonzalez and J. A. Pople, Gaussian 03,Revision B.03, Gaussian, Inc., Pittsburgh PA, 2003.

36 A. G. Baboul, L. A. Curtiss, P. C. Redfern and K. Raghavachari, J.Chem. Phys., 1999, 110, 7650.

37 See for example: (a) M. L. Coote,Macromol. Theory Simul., 2009, 18,388; (b) C. Y. Lin andM. L. Coote, Aust. J. Chem., 2009, 62, 1479; (c)C. Y. Lin, E. I. Izgorodina and M. L. Coote, Macromolecules, 2010,43, 553; (d) J. L. Hodgson, L. B. Roskop, M. S. Gordon, C. Y. Linand M. L. Coote, J. Phys. Chem. A, 2010, 114, 10458; (e)S. H. Kyne, C. Y. Lin, I. Ryu, M. L. Coote and C. H. Schiesser,Chem. Commun., 2010, 46, 6521.

38 See for example: M. L. Coote, C. Y. Lin, A. L. J. Beckwith andA. A. Zavitsas, Phys. Chem. Chem. Phys., 2010, 12, 9597.

39 See for example: H. Zare, M. Eslami, M. Namazian and M. L. Coote,J. Phys. Chem. B, 2009, 113, 8080.

40 See for example: J.HoandM.L.Coote,Theor. Chem.Acc., 2010, 125, 3.41 A. P. Scott and L. Radom, J. Phys. Chem., 1996, 100, 16502.42 E. I. Izgorodina, C. Y. Lin and M. L. Coote, Phys. Chem. Chem.

Phys., 2007, 9, 2507.43 E. I. Izgorodina,D.R.B.Brittain, J.L.Hodgson,E.H.Krenske,C.Y.Lin,

M. Namazian andM. L. Coote, J. Phys. Chem. A, 2007, 111, 10754.44 See for example: (a) D. R. Stull, E. F. Westrum, Jr and G. C. Sinke,

The Thermodynamics of Organic Compounds, John Wiley & Sons,New York, 1969; (b) P. J. Robinson, J. Chem. Educ., 1978, 55, 509;(c) J. I. Steinfeld, J. S. Francisco and W. L. Hase, Chemical Kineticsand Dynamics, Prentice Hall, Englewood Cliffs, New Jersey, 1989.

45 These formulae are outlined in full in theESI† of an earlier publication:M. L. Coote and L. Radom,Macromolecules, 2004, 37, 590.

46 See for example: B. A. Ellingson, V. A. Lynch, S. L. Mielke andD. G. Truhlar, J. Chem. Phys., 2006, 125, 084305.

47 This program is freely available from http://rsc.anu.edu.au/%mcoote/scripts.html.

48 M. D. Liptak, K. G. Gross, P. G. Seybold, S. Feldgus andG. C. Shields, J. Am. Chem. Soc., 2002, 124, 6421.

49 (a) A. Klamt, J. Phys. Chem., 1995, 99, 2224; (b) A. Klamt, COSMO-RS: From Quantum Chemistry to Fluid Phase Thermodynamics andDrug Design, Elsevier Science Ltd., Amsterdam, The Netherlands,2005; (c) A. Klamt, V. Jonas, T. Burger and J. C. W. Lohrenz, J.Phys. Chem. A, 1998, 102, 5074.

50 J. N. Louwen, C. Pye and E. V. Lenthe, ADF2008.01 COSMO-RS,SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam, TheNetherlands, 2008, http://www.scm.com.

51 C. C. Pye, T. Ziegler, E. van Lenthe and J. N. Louwen,Can. J. Chem.,2009, 87, 790.

52 (a) S. Kobayashi, M. Sugiura, H. Kitagawa and W. W.-L. Lam,Chem. Rev., 2002, 102, 2227; (b) R. Bastin, H. Albadri,A.-C. Gaumont and M. Gulea, Org. Lett., 2006, 8, 1033.

Polym. Chem. This journal is ª The Royal Society of Chemistry 2011

Dow

nloa

ded

on 0

5 Ja

nuar

y 20

12Pu

blish

ed o

n 26

Oct

ober

201

1 on

http

://pu

bs.rs

c.or

g | d

oi:1

0.10

39/C

1PY

0035

6A

View Online