Vous êtes sur la page 1sur 171

Conjugated Polymer Networks: Synthesis and Properties

by

Akshay Kokil

Submitted in partial fulfillment of the requirements For the degree of Doctor of Philosophy

Thesis Advisor: Dr. Christoph Weder

Department of Macromolecular Science and Engineering CASE WESTERN RESERVE UNIVERSITY

August 2005

CASE WESTERN RESERVE UNIVERSITY SCHOOL OF GRADUATE STUDIES

We hereby approve the dissertation of

Akshay Kokil ______________________________________________________


candidate for the Ph.D. degree *.

(signed)_______________________________________________ (chair of the committee)

Dr. Christoph Weder

________________________________________________

Dr. Kenneth D. Singer


Dr. David Schiraldi

________________________________________________

________________________________________________

Dr. Stuart Rowan

________________________________________________

________________________________________________

(date) _______________________

05/15/2005

*We also certify that written approval has been obtained for any proprietary material contained therein.

Table of Contents List of Tables List of Schemes List of Figure List of Acronyms and Abbreviations Abstract Chapter 1 Introduction Class of Organic Materials 1.2 Charge Transport in Conjugated Polymers 1.3 Conjugated Polymer Networks 1.4 Poly(p-phenylene ethynylene)s: An Important Class of Conjugated Polymers 1.5 Energetics of Conjugated Polymer Networks 1.6 References Chapter 2 Chapter 3 Scope and Objectives Charge Carrier Transport in Poly(p-phenylene ethynylene)s

1 4 5 7 13 16

1.1 (Semi)conducting Conjugated Polymers - An Emerging 19 20 30

39 42 46 53

3.1 Introduction 3.2 Experimental Section 3.3 Charge Carrier Transport in EHO-OPPE 3.4 Conclusions 3.5 References Chapter 4 Synthesis and Properties of Poly(2,5-dialkoxy-p-phenylene ethynylene)-PtII Networks

56 57 58 69 70

4.1 Introduction 1

71

4.2 Experimental Section 4.3 Charge Transport in PPE-PtII Networks 4.4 Conclusions 4.5 References 79 Chapter 5 Synthesis and Properties of Poly(2,5-dialkoxy-p-phenylene ethynylene)-Pt0 Networks

72 72 77

5.1 Introduction 5.2 Experimental Section 5.3 Two Alternative, Convenient Routes to Bis(diphenylacetylene)Pt0 5.4 Synthesis of Organometallic -Conjugated PPE-Pt0 Networks 5.5 Charge Carrier Mobility in Organometallic -Conjugated PPE-Pt0 Network 5.6 Conclusions 5.7 References Chapter 6

80 83 86 89

97 102 104

Synthesis and Properties of Organometallic Networks Based on 2,2Bipyridine-Containing Poly(p-phenylene ethynylene)s 106 107

6.1 Introduction 6.2 Experimental Section 6.3 Synthesis and Characterization of BipyPPEs and Their Organometallic Networks 6.4 Conclusions 6.5 References Chapter 7 Synthesis and Characterization of Cross-linked Poly(p-phenylene ethynylene)s

114 134 135

7.1 Introduction

138

7.2 Experimental Section 7.3 Bulk Synthesis and Characterization of Crosslinked Poly(p-phenylene ethynylenes)s 7.4 Synthesis of Crosslinked Conjugated Polymer Milli-, Micro- and Nanoparticles 7.5 Conclusions 7.6 References Chapter 8 Conclusions and Outlook

139

146

153 157 159 161 167 169

Acknowledgements List of Publications Based on this Thesis

List of Tables Table 3.1 Results of data fits to the Gaussian disorder model of charge transport. 68

Table 6.1

Optical absorption and PL emission data of BipyPPE1 (13) and BipyPPE2 (14) after complexation with different metal ions.

116

List of Schemes Scheme 1.1 Simplified schematic representation of cross-linked conjugated polymer networks with organometallic cross-links (top) and covalent cross-links (bottom). Scheme 1.2 Schematic representation of the cross-linking reaction reported to occur in zinc phorphyrin-linked poly(p-phenylene ethynylene)s. Scheme 1.3 Schematic representation of the cross-linking reaction proposed to occur in poly[(4-ethynyl)phenylacetylene] upon thermal treatment. Scheme 1.4 Schematic representation of the proposed cross-linking reaction occurring upon heat treatment of Cr(CO)3-benzene containing poly(arylene ethynylene)s. Scheme 1.5 Simplified representation of the ligand-exchange reaction between MEH-OPPE (1) and [Pt-(-Cl)Cl(PhCH=CH2)]2 (2), leading to cross-linked organometallic hybrid materials 3. Scheme 4.1 Simplified representation of the ligand-exchange reaction between EHO-OPPE (4) and [Pt-(-Cl)Cl(PhCH=CH2)]2 (2), leading to cross-linked organometallic conjugated polymer network. Scheme 5.1 Synthesis of Pt(2-Ph-CC-Ph)2 (8) through (i) ligand exchange exchange between Pt(PhCH=CH2)3 (7) and Ph-CC-Ph, and (ii) reduction of cis-PtCl2(PhCH=CH2)2 (6) with triphenylsilane in the presence of Ph-CC-Ph. Scheme 5.2 Ligand-exchange reaction between EHO-OPPE (4) and Pt(PhCH=CH2)3 (7) leading to the target EHO-OPPE-Pt0 (9) networks. Scheme 6.1 Reaction scheme for the synthesis of 5,5-bis((trimethylsilyl)ethynyl)-2,2-bipyridine. Scheme 6.2 Reaction scheme for the synthesis of 5,5-bis(ethynyl)-2,2bipyridine (9). Scheme 6.3 Synthesis and molecular structure of the 2,2-bipypridinecontaining poly(2,5-dialkyloxy-p-phenylene ethynylene)s BipyPPE1 (13) and BipyPPE2 (14). 30

33 34 35

37

73

87

89

109 110 115

Scheme 6.4 Schematic representation of the formation of metallosupramolecular networks via the formation of metal-bis-ligand complexes with 2,2-bipypridine-containing poly(2,5-dialkyloxy-pphenylene ethynylene)s BipyPPE1 (13) and BipyPPE2 (14). Scheme 7.1 Synthesis of cross-linked PPEs (18) by the palladium-catalyzed cross-coupling reaction of 2,5-diiodo-4-[(2-ethylhexyl)oxy] methoxybenzene (16) or 1,4-bis[(2-ethylhexyl)oxy]-2,5-diiodo benzene (10), 1,4-diethynyl-2,5-bis-(octyloxy)benzene (11) and the tri-functional cross-linker 1,2,4-tribromobenzene (17). R1=2ethylhexyl, R2=n-octyl, R3=methyl (16) or 2-ethylhexyl (10).

123

147

List of Figures Figure 1.1 Photographs of applications of conjugated polymers in (a) anticorrosion coatings, (b) field-effect transistors (FETs) and (c) light-emitting diodes (LEDs). Figure 1.2 Schematic representation of the doping of poly(p-phenylene ethynylene) (PPE) under formation of polarons (radical cations, radical anions) and bipolarons (dications, dianions). Note that multiple carriers can coexist on each macromolecule. Figure 1.3 Schematic representation of intrachain charge diffusion (left) and interchain charge diffusion (hopping, right) in polyacetylene. Figure 1.4 Chemical structures of (a) polyacetylene, (b) polyaniline, (c) poly (pphenylene vinylene), (d) poly(N-phenylimino-1,4-phenylene-1,2ethenylene-1,4-(2,5-dioctoxy)-phenylene-1,2-ethenylene-1,4phenylene), (e) poly-9,9dioctyl-fluorene-co-bithiophene, (f) poly (benzobisimidazobenzophenanthroline) and (g) poly(benzobisimidazole). Figure 1.5 Two different preferential orientations of the ordered domains of PAT with respect to the employed substrate. (a) in-plane orientation and (b) out-of-plane orientation. Figure 1.6 Schematic chemical structure of conjugated cross-links in electrochemically synthesized polypyrrole. Figure 1.7 Chemical structure of poly(o-toluidine) cross-linked with palladiumII. Figure 1.8 Bridged Pd terthiophene complex employed for electropolymerization. Figure 1.9 General schematic structures of poly(arylene ethynylene)s (PAEs), poly(arylene)s (PAs), poly(arylene vinylene)s (PAVs), and poly(diacetylene)s (PDAs). Figure 1.10 Chemical structure of poly[2,5-dioctyloxy-1,4-diethynyl-phenylenealt-2,5-bis(2'-ethylhexyloxy)-1,4-phenylene] (EHO-OPPE, 4). Figure 1.11 Simplified schematic of electron (top) and hole (bottom) transfer in conjugated polymers with matched energy levels (a and c) and unmatched energy levels (b and d). 19

22

23 26

27

32 35 38 40

42 43

Figure 1.12 Schematic structure of metallopolymer demonstrating charge transport by outer sphere electronic coupling. Figure 1.13 Schematic structure of metallopolymer demonstrating charge transport by inner sphere electronic coupling. Figure 3.1 Chemical structure of poly[2,5-dioctyloxy-1,4-diethynyl-phenylenealt-2,5-bis(2'-ethylhexyloxy)-1,4-phenylene] (EHO-OPPE, 4), the poly(p-phenylene ethynylene) (PPE) derivative investigated here. Figure 3.2 Schematic representation of sample preparation (left) and the architecture of the sample used for TOF measurements (right). Figure 3.3 Simplified representation of the conventional time-of-flight setup. The transients displayed are schematic representations of the signal observed for non-dispersive transport (top) and dispersive transport (bottom). Figure 3.4 Electron time-of-flight photocurrent transients of a solution-cast film of EHO- OPPE (4, L=8 m), measured at 295 K and an electric field of 2.5105 Vcm-1 in a (a) linear and (b) double logarithmic representation. Figure 3.5 Hole time-of-flight photocurrent transients of a solution-cast film of EHO-OPPE (4, L=8 m), measured at 295 K and an electric field of 2.5105 Vcm-1 in (a) linear and (b) double logarithmic representation. Figure 3.6 Electron (a) and hole (b) mobilities of solution cast films of EHOOPPE (4) as function of electric field at various film thicknesses (open triangles: L=6.5 m, filled circles: L=8 m, open squares: L=12 m). Figure 3.7 Temperature dependence of (a) electron and (b) hole mobilities of an EHO-OPPE (4) film (L=8 m) measured at E = 2105 (squares), 3105 (circles) and 4105 Vcm-1 (triangles).
Figure 3.8 Temperature dependence of the diagonal disorder parameter for holes (squares) and electrons (circles). Parameters have been obtained by fitting the experimental data (Figure 3.7) to Equations 3.2 and 3.6.
II (5), Figure 4.1 Hole TOF photocurrent transients of EHO-OPPE-Pt [PtII]/[PE]=0.17 with film thickness L=9 m (black solid line) and L=13 m (red solid line), linear (top) and logarithmic (bottom) plots, measured at 295 K and F=9104 Vcm-1.

44 45 56

58 59

61

62

63

65

69

74

Figure 4.2 Electron TOF photocurrent transients of EHO-OPPE-PtII (5), [PtII]/[PE]=0.17 with film thickness L=9 m (black solid line) and L=13 m (red solid line), linear (top) and logarithmic (bottom) plots, measured at 295 K and F=9104 Vcm-1. Figure 4.3 Electron (top) and hole (bottom) mobility of EHO-OPPE-PtII (5) as function of [PtII]/[PE] ([PtII]/[PE]: filled symbols =0.35, open symbols =0.17), electric field F and thickness (=18.4 m, =10.1 m, =11.3 m, =9 m, =13 m, =6.5 m). Mobilities for EHO-OPPE () are displayed for reference. Figure 5.1
195

75

76

Pt NMR spectra for Pt(PhCH=CH2)3 (7, a), 7 with 1.47 eq. of PhCC-Ph (b) and 7 with 3.42 eq. of Ph-CC-Ph (c).

88 91

Figure 5.2 Photoluminescence spectra of a solution of EHO-OPPE (4) in styrene (solid line, 5 mg / mL) and an EHO-OPPE-Pt0 (9) gel (broken line, [Pt0] / [PE] = 0.68, 5 mg PPE / mL styrene). The spectra were acquired under excitation at 380 nm. Figure 5.3 Left: Styrene solution of EHO-OPPE (4, 6.7 mg/mL). Right: EHOOPPE-Pt0 (9) / styrene gel ([Pt0] / [PE] = 0.68, 6.7 mg PPE / mL styrene), carrying a steel ball. The picture on the left was taken under illumination with UV light (365 nm). Figure 5.4 Pictures of an EHO-OPPE-Pt0 (9) film in styrene (left) and toluene (right) under illumination with UV-light (365 nm). The dissimilar dissolution characteristics of the photoluminescent polymer demonstrate the reversibility of the ligand-exchange reaction and the cross-linked nature of the material. Figure 5.5 Raman spectra of Ph-CC-Ph (a, top left), EHO-OPPE (4, b, top right), Pt(2-Ph-CC-Ph)2 (8, c, bottom left) EHO-OPPE-Pt0 (9, d, [Pt0] / [PE] ~ 0.17, bottom right). Figure 5.6 UV-Vis absorption spectra of films of EHO-OPPE (4) (solid line) and EHO-OPPE-Pt0 (9) with [Pt0] / [PE] = 0.086 (filled circles), 0.17 (open circles), 0.25 (filled triangles), and 0.34 (open triangles).
2 Figure 5.7 Crystal structure for Pt( -Ph-CC-Ph)2 (8).

92

93

94

95

96 96

Figure 5.8 Photoluminescence spectra of films of EHO-OPPE (4) (solid line) and EHO-OPPE-Pt0 (9) with [Pt0] / [PE] = 0.086 (filled circles), 0.17 (open circles), 0.25 (filled triangles), and 0.34 (open triangles).

Figure 5.9 Electron TOF photocurrent transients of EHO-OPPE (4, solid, L=8 m) and EHO-OPPE-Pt0 (9, dotted, L=30 m, [Pt0]/[PE]=0.17) films in linear (top) and logarithmic (bottom) plots, measured at 295 K and F = 1.5105 Vcm-1. Figure 5.10 Electron (top) and hole (bottom) mobility of EHO-OPPE-Pt0 (9) as function of [Pt0]/[PE] and electric field F ([Pt0]/[PE]: =0, =0.016, =0.086, =0.17, =0.25, =0.34). Figure 5.11 Electron (top) and hole (bottom) mobility at F = 1.1105 () and 1.5105 Vcm-1 () as function of [Pt0]/[PE]. Figure 5.12 Electron (a) and hole (b) mobility of films of EHO-OPPE-Pt0 (9) with [Pt0] / [PE] = 0.25 having thickness 25m (filled squares) and 20m (open squares). Figure 6.1 UV-Vis absorption (a) and PL emission (b) spectra acquired upon addition of tetrakis(acetonitrile)Cu(I)-hexafluorophosphate to BipyPPE1 (13) (concentration of polymer-bound Bipy = 1.9310-5 M) in CHCl3:CH3CN (15:1 v/v). Shown are spectra at selected [Cu+]:[Bipy] ratios of 0 (solid line), 0.09 (filled squares), 0.19 (filled circles), 0.28 (filled triangles), 0.38 (filled inverted triangles), 0.48 (filled rhombus), 0.57 (empty squares), 0.76 (empty circles), 0.96 (empty triangles) and 1.92 (empty inverted triangles). The insets show the absorption at 452 nm (a) and the emission at 459 nm (b) as a function of [Cu+]:[Bipy] ratio. Figure 6.2 UV-Vis absorption (a) and PL emission (b) spectra acquired upon addition of tetrakis(acetonitrile)Cu(I)-hexafluorophosphate to BipyPPE2 (14) (concentration of polymer-bound Bipy = 3.2310-5 M) in CHCl3:CH3CN (15:1 v/v). Shown are spectra at selected [Cu+]:[Bipy] ratios of 0 (solid line), 0.1 (filled squares), 0.2 (filled circles), 0.3 (filled triangles), 0.4 (filled inverted triangles), 0.5 (filled rhombus), 0.6 (empty squares), 0.8 (empty circles), 1.0 (empty triangles) and 2.0 (empty inverted triangles).The insets show the absorption at 440 nm (a) and the emission at 482 nm (b) as a function of [Cu+]:[Bipy] ratio. Figure 6.3 UV-Vis absorption spectra of model compound 15 in CHCl3:CH3CN (15:1 v/v). Solid line: preparation in-situ. Dashed line: isolated compound. Figure 6.4 Representation of the UV-Vis absorption data (a) and PL emission data (b) shown in the inset of Figures 6.1a and 6.1b respectively in a Scatchard plot.

98

99

100 101

117

118

120

122

10

Figure 6.5 Representation of the UV-Vis absorption data (a) and PL emission data (b) shown in the inset of Figures 6.2a and 6.2b respectively in a Scatchard plot. Figure 6.6 UV-Vis absorption (a) and PL emission (b) spectra acquired upon addition of free 2,2-bipyridine to a mixture of BipyPPE1 (13) (concentration of polymer-bound Bipy = 1.9310-5 M) and tetrakis(acetonitrile)Cu(I)-hexafluorophosphate (ratio of [Cu+]:[Bipy comprised in the polymer] = 1:0.96) in CHCl3:CH3CN (15:1 v/v). Shown are spectra for ratios of [free Bipy]:[polymer-bound Bipy] of 0:1 (dashed line), 1:1 (filled squares), 2:1 (filled circles), 5:1 (filled triangles) and 10:1 (filled rhombus). The spectra of neat BipyPPE1 (13, solid line) are included for reference. Figure 6.7 UV-Vis absorption (a) and PL emission (b) spectra acquired upon addition of cobalt(II)tetrafluoroborate hexahydrate (filled squares), nickel(II)perchlorate hexahydrate (empty circles), zinc(II)perchlorate hexahydrate (filled triangles), and cadmium(II)perchlorate hydrate (filled circles) to BipyPPE1 (13) (concentration of Bipy comprised in the polymer in solution = 1.9310-5 M) in CHCl3:CH3CN (15:1 v/v). Shown are spectra for [M2+]:[Bipy] ratios of 0.5:1. The spectra of neat BipyPPE1 (13. solid line) are included for reference. Figure 6.8 UV-Vis absorption (a) and PL emission (b) spectra acquired upon addition of cobalt(II)tetrafluoroborate hexahydrate (filled squares), nickel(II)perchlorate hexahydrate (empty circles), zinc(II)perchlorate hexahydrate (filled triangles), and cadmium(II)perchlorate hydrate (filled circles) to BipyPPE2 (14) (concentration of polymer-bound Bipy =3.2310-5 M) in CHCl3:CH3CN (15:1 v/v). Shown are spectra for [M2+]:[Bipy] ratios of 0.5:1. The spectra of neat BipyPPE2 (14, solid line) are included for reference. Figure 6.9 (a) UV-Vis absorption and (b) PL emission spectra of spin-coated films of complexes produced by ligand-exchange reactions between BipyPPE1 (13) and zinc perchlorate hexahydrate (filled rhombus), or cadmium perchlorate hydrate (filled squares). Shown are spectra for [metal]:[Bipy] ratios of 0.5:1. The spectra of neat BipyPPE1 (13, solid line) are included for reference. Figure 6.10 (a) UV-Vis absorption and (b) PL emission spectra of spin-coated films of complexes produced by ligand-exchange reactions between BipyPPE2 (14) and zinc perchlorate hexahydrate (filled rhombus), or cadmium perchlorate hydrate (filled squares). Shown are spectra for [metal]:[Bipy] ratios of 0.5:1. The spectra of neat BipyPPE2 (14, solid line) are included for reference.

125

127

128

130

131

132

11

Figure 7.1 Photograph of a gel of covalently cross-linked PPE network. Figure 7.2 Residual content of bromine detected by elemental analysis in the cross-linked PPEs, as function of the ratio of monomers 16:17 ( = 18a, = 18b, = 18c, = 18d, = 18e). Figure 7.3 Normalized equilibrium weight increase for polymers 18a 18e swollen in toluene as function of the ratio of monomers 16:17 ( = 18a, = 18b, = 18c, = 18d, = 18e). Figure 7.4 Photoluminescence spectra of the cross-linked PPEs investigated here (18a, open triangles, a; 18b, open circles, b; 18c, open squares, c; 18d, filled triangles, d; 18e, filled circles, e; all swollen in toluene) and a (dry) MEH-OPPE reference film (filled squares, f). Spectra are scaled to optimally fit the graph. Figure 7.5 Photograph of an image through a cross-linked film of MEHOOPPE-X3.4 18f. Figure 7.6 Photographs (a), optical micrographs (b) and scanning electron micrographs (c) of cross-linked conjugated milli- (a) micro- (b), and nanoparticles (prepared in reaction nanometer-sized particles a) (c). Photographs and optical micrographs were taken in fluorescence mode under excitation at 366 nm and transmission/reflection mode, with the polymer particles dispersed in toluene. Figure 7.7 Size-distribution of cross-linked PPE micro- (a) and nanoparticles (prepared in reaction nanometer-sized particles a) (b). The size of the individual particles was determined from optical transmission microscopy (a) and scanning electron microscopy (b) images. The particles evaluated in (a) were swollen with toluene. Figure 7.8 SEM image of cross-linked EHO-OPPE nanoparticles. Figure 7.9 Photoluminescence spectra of PPE milli (Method B, dashed line), micro- (Method C, dotted line) and nanoparticles (Method D, dashdotted line), suspended in toluene, and as a reference an MEH-OPPE solution in toluene (solid line). Figure 8.1 Electron (circles) and hole (squares) mobilities for covalently crosslinked nano-particles (9 % mol cross-linker, filled symbols) and EHO-OPPE (open symbols).

148 149

150

151

152 154

155

156 157

165

12

List of Acronyms, Abbreviations and Chemical Compounds Bipy Eg k 2,2-bipyridine band gap boltzman constant carrier mobility carrier transit time charge on the charge carrier critical extent of reaction diagonal (energetic) disorder parameter electrical conductivity electric field ethylenediamine tetraacetate field-effect transistor highest occupied molecular orbital indium-tin-oxide light-emitting diode lowest unoccupied molecular orbital metal to ligand charge transfer nuclear magnetic resonance number-average molecular weight number average degree of polymerization number of charge carriers off-diagonal (positional) disorder parameter one dimensional phenylene ethynylene polyacetylene poly(3-alkyl thiophene) poly(arylene ethynylene) poly(arlene vinylene) poly(benzobisimidazobenzo phenanthroline)

ttr e c

c
E EDTA FET HOMO ITO LED LUMO MLCT NMR Mn Xn n

1D PE PA PAT PAE PAV BBL

13

PDA EHO-OPPE MEH-OPPE BipyPPE1 BipyPPE2

poly(diacetylene) poly[2,5-dioctyloxy-1,4-diethynyl-phenylene-alt-2,5,-bis(2'-ethylhexyloxy)-1,4-phenylene] poly[2,5-dioctyloxy-1,4-diethynyl-phenylene-alt-2-methoxy,5-2-ethylhexyloxy)-1,4-phenylene] poly[2,2-bipyridine-5,5diylethynylene[2,5-bis(2-ethylhexyl)oxy-1,4phenylene]ethynylene] poly[(2,2-bipyridine-5,5diylethynylene[2,5-bis(2-ethylhexyl)oxy-1,4phenylene]ethynylene)-co-(2,5-dioctyloxy-1,4-diethynyl-phenylene-alt2,5-bis(2'-ethylhexyloxy)-1,4-phenylene)]

PPE PPV LED PL

poly(p-phenylene ethynylene) poly(p-phenylene vinylene) polymer light-emitting diode photoluminescence prefactor mobility sample thickness sodium dodecyl sulphate temperature time-of-flight transient time ultraviolet-visible voltage wavelength of maximum absorption or emission

0
L SDS T TOF ttr UV-Vis V max

1 2 3 4 5 6

MEHO-OPPE [Pt-(-Cl)Cl(PhCH=CH2)]2 MEHO-OPPE-PtII networks EHO-OPPE EHO-OPPE-PtII networks PtCl2(PhCH=CH2)2 14

7 8 9 10 11 12 13 14 15 16 17 18a 18b 18c 18d 18e 18f

Pt(PhCH=CH2)3 Pt(2-Ph-CC-Ph)2 EHO-OPPE-Pt0 1,4-Bis[(2-ethylhexyl)oxy]-2,5-diiodobenzene 1,4-diethynyl-2,5-bis-(octyloxy) benzene 5,5-diethynyl-2,2-bipyridine BipyPPE1 BipyPPE2 Bis(2,2-bipyridine)Cu(I)-hexafluorophosphate 2,5-Diiodo-4-[(2-ethylhexyl)oxy]methoxybenzene 1,2,4-tribromobenzene O-OPPE-X MEHO-OPPE-X0.5 MEHO-OPPE-X1.3 MEHO-OPPE-X4.4 MEHO-OPPE-X10.5 MEHO-OPPE-X3.4

15

Abstract The experimental research program that forms the basis of this thesis has been directed towards the design, synthesis, processing and physical characterization of welldefined conjugated polymer networks. It attempts to provide answers to the questions how such materials can be synthesized and processed and how the introduction of crosslinks can be exploited for the creation of polymeric materials with optimized optic and electronic characteristics. Interestingly, this family of materials has received little attention in the past, at least as far as systematic studies of well-defined systems are concerned. This situation may be a direct consequence of the challenge to introduce conjugated cross-links into conjugated polymers and retain adequate processibility. We have shown that organometallic polymer networks based on linear conjugated polymers are readily accessible through ligand-exchange reactions. This approach was exemplified by exploiting the ethynyl moieties comprised in poly(p-phenylene ethynylene) (PPE) derivatives as ligand sites, which allow for complexation with selected metals and cross-linking via the resulting PPE-Metal complexes. Focusing on the dinuclear complex [Pt-(-Cl)Cl-PPE]2 and PPE-Pt0 as crosslinks, we have conducted an in-depth investigation on how the nature of the metal cross-links influences the materials characteristics, in particular the charge transport properties. We first investigated the charge carrier mobility of poly[2,5-dioctyloxy-1,4-diethynyl-phenylene-alt-2,5-bis(2'ethylhexyloxy)-1,4-phenylene] (EHO-OPPE), as a classic representative of poly(pphenylene ethynylene) (PPE) derivatives, which represent an important class of conjugated polymers. In what appears to be the first study ever conducted on the mobility of any PPE, we found that EHO-OPPE displays ambipolar charge transport

16

characteristics with very high electron (1.910-3 cm2V-1s-1) and hole (1.610-3 cm2V-1s-1) mobilities. Most importantly, the introduction of Pt0 cross-links was found to enhance the charge carrier mobility of the investigated systems by up to an order of magnitude. Electron and hole mobilities of the order of 1.510-2 cm2V-1s-1 were measured for these materials, representing the highest mobilities ever reported for disordered conjugated polymers. More importantly, the study unequivocally proves that the introduction of conjugated cross-links indeed leads to a significant improvement of the carrier mobility of conjugated polymers. Work involving [Pt-(-Cl)Cl-PPE]2 cross-links suggests that the nature of the metallic cross-links is exceedingly important. We have shown that the PtII centers act as traps for both electrons and holes and impede efficient charge transport. With the aim to expand the materials basis and exploit other chemical platforms for the cross-linking process, we integrated 2,2-bipyridine (bipy) moieties into the conjugated polymer backbone. Bipy is a most versatile ligand, which can bind to a broad variety of metals. We have synthesized a library of PPEs in which the bipyridine content is systematically varied. Using a variety of transition metal complexes, we have systematically investigated the ligand-exchange reactions of these polymers.

Complexation studies suggest that ligand exchange indeed leads to three-dimensional networks, which feature BipyPPE-metal-BipyPPE cross-links and display interesting optoelectronic properties. We also showed that the ligand exchange is a cooperative process, which in the presence of a competing ligand is fully reversible. Complexes with group 12 d10 ions (Zn2+ and Cd2+) are emissive, while other transition metals such as Cu+, Ni2+ form non-radiative metal-to-ligand charge-transfer complexes with the polymers.

17

In a second approach we have demonstrated that the processing issues associated with covalently cross-linked conjugated polymer networks can be readily overcome if these materials are synthesized in the form of cross-linked nanoparticles and processed as suspensions. In a series of ground-breaking experiments, we have shown that covalently cross-linked conjugated polymer particles can be produced by conducting cross-coupling reactions in aqueous emulsions instead of homogeneous solutions and using multifunctional cross-linkers. The size of the spherical polymer particles can easily be tuned over a wide range (mm to nm) by modification of the reaction conditions. The resulting materials can be processed in the form of suspension and promise interesting electronic properties.

18

Chapter 1: Introduction 1.1 (Semi)conducting Conjugated Polymers an Emerging Class of Organic Materials The role of polymers in the electronics industry has traditionally been associated with electrically insulating properties and macromolecular materials have found widespread use in packaging applications as passive dielectrics. However, the discovery of electrically (semi)conducting conjugated polymers has radically changed this situation.1 These materials are attracting significant interest not just academically but also

Figure 1.1 Photographs of applications of conjugated polymers in (a) anticorrosion coatings, (b) field-effect transistors (FETs) and (c) light-emitting diodes (LEDs).

industrially, since they may combine the processability and mechanical properties of polymers with the readily-tailored optoelectronic properties of organic molecules.2 Especially the potential use of these "synthetic metals" as electrical conductors,3 in lightemitting diodes (LEDs),4,5,6 field-effect transistors (FETs),7 photorefractive devices8, and

19

photovoltaic cells9 have motivated the development of synthesis and processing methods of conjugated polymer materials with unique field-responsive properties (Figure 1.1).3,10,11 Breathtaking progress has been made in the last two decades, and the field has matured to the onset of commercial exploitation of conjugated polymers as polymer conductions in corrosion control,12 in light-emitting diodes,4,5,6 sensors13 and a number of other applications.14 One key problem for the full commercial exploitation of the field is that the charge carrier mobility of state-of-the-art polymer semiconductors is much lower than required for many applications.14b This thesis is focused on the investigation of a new concept of improving this important property. It is based on the hypothesis that the charge transport in these polymers can be significantly improved through the introduction of conjugated cross-links between the individual polymer chains.

1.2 Charge Transport in Conjugated Polymers Conjugated polymers are characterized by a molecular structure that features alternating single and multiple bonds. The electronic and physico-chemical characteristics of conjugated macromolecules are not only governed by the nature of the polymer backbone, but also by the intermolecular interactions. Electrical conductivity is a direct consequence of delocalized chemical bonding.15 The absence of sp3 hybridized carbon atoms leads to a situation where overlap of p orbitals on successive carbon atoms enables the delocalization of -electrons along the polymer backbones. The large number of atomic orbitals in the macromolecules translates into a large number of molecular orbitals, which form a band of energies. In the case of a metal this energy band is a continuum; due to the high density of electronic states with electrons of relatively low

20

binding energy, "free electrons" can easily redistribute and under an applied electric field move easily from atom to atom, thus rendering the material electrically conducting. Hckel's theory predicts that in conjugated macromolecules -electrons are delocalized in similar fashion over the entire chain, so that one would expect that the electronic properties of a polymer material composed of sufficiently long conjugated chains are also described well by a continuous energy band. According to this model, an individual chain of the conjugated polymer would be a one-dimensional (1D) metallic conductor. However, as a result of the Peierls instability, the density of -electrons in conjugated organic molecules is not the same between all atoms;16 there is a distinct alternation between single and multiple bond character, as is for example evident from the chemical structure of poly(p-phenylene ethynylene) (PPE) shown in Figure 1.2. Thus, the electronic properties of conjugated polymers in their neutral oxidation state are usually better described by a filled valence band (-band, bonding) formed by the highest occupied molecular orbitals (HOMOs) and an empty conduction band (*-band, antibonding) formed by the lowest unoccupied molecular orbitals (LUMOs). Because the energy difference between the highest occupied and the lowest unoccupied band, referred to as band gap (Eg), is usually not near zero, and because there are no partially filled bands, conjugated polymers are typically semiconductors in their neutral, undoped state. Eg depends on the molecular structure of the polymer's repeat unit and can be controlled via modification of the latter. In their pioneering work on polyacetylene, MacDiarmid, Shirakawa and Heeger demonstrated that doping allows one to increase the electrical conductivity of conjugated polymers by many orders of magnitude so that metallic properties are achieved. Doping

21

refers to either removing (oxidation, p-doping) or adding electrons (reduction, n-doping) to the polymer, as shown schematically in Figure 1.2 at the example of PPE. This can be accomplished by either conventional chemical or electrochemical means. A number of, defect types can be formed upon doping, including radical cations or anions (referred to as polarons) and dications or dianions (referred to as bipolarons). As a result of the doping process, the electrochemical potential (Fermi level) of the polymer is moved into an energy regime with a high density of electronic states. Charge neutrality is maintained through counter-ions, and the doped polymer effectively becomes a salt. The extra electrons or vacancies (holes) introduced through doping act as charge carriers. The
+ - e(p-doping) PPE 2+ n + e(n-doping) + e2-

n e-

Figure 1.2 Schematic representation of the doping of poly(p-phenylene ethynylene) (PPE) under formation of polarons (radical cations, radical anions) and bipolarons (dications, dianions). Note that multiple carriers can coexist on each macromolecule.

delocalization of these carriers can be quite limited, partly because of Coulomb attraction to their counter-ions, and partly because of a local change in the equilibrium geometry of the doped relative to the neutral molecule. However, since every repeat unit is a potential redox site, the doping level of conjugated polymers can usually be rather well controlled.

22

For many materials platforms high levels of electrical conductivity can be achieved at high doping levels, which are concomitant with a high density of charge-carriers, especially in the case of trans-polyacetylene, conductivities upto 108 S/cm (similar to that of Cu) have been observed.17 The electrical conductivity (c) in conjugated polymer systems is expressed by:

c = n e

Equation 1.1

where n is the number of charge carriers (which can be adjusted by the level of doping), e is their charge and signifies the charge carrier mobility. The investigation of new concepts for the design of materials in which this important figure of merit is maximized is the focus of the present work. The charge transport characteristics of conjugated polymers are, of course,

.
+

.
+

.
+

.
+

+
+

intrachain charge diffusion

interchain charge diffusion

Figure 1.3 Schematic representation of intrachain charge diffusion (left) and interchain charge diffusion (hopping, right) in polyacetylene.

23

primarily governed by the nature of the polymer backbone itself, but intermolecular interactions also exert an important influence on the macroscopic materials properties. The charge carrier mobility of conjugated polymers is a function of intrachain charge diffusion and interchain interactions, i.e., hopping (Figure 1.3).3,7,9,10,11 Both factors depend on a number of variables; the former is mostly based on the polymer's chemical structure, the number and nature of defect sites, conformation of the polymer backbone, and the molecular weight, while the latter strongly depends on the supramolecular architecture, i.e., the degree of contact, order and orientation.18 The fact that conducting polymers usually show a different variation of conductivity with temperature than metals has been related to disorder effects, which may dramatically limit interchain interactions and, thus, charge transport in these systems.18b,d,h,k Consequently, much work has been focused on maximizing the supramolecular order in conjugated polymers, in order to optimize their charge transport characteristics and electrical conductivity, and also a variety of other properties.18 As a result, exciting progress has been documented for the charge carrier mobility (and also electrical conductivity) in semiconducting polymer systems with high molecular order and/or orientation. Amorphous, semiconducting polymers exhibit a field-effect mobility (for holes) of the order of 10-5 cm2/Vs (about five orders of magnitude lower than the mobility required for the potential application in plastic thin film transistors, 1 cm2/Vs). The original work on the development of polymer field-effect transistors based on conjugated polymers relied on polyacetylene (Figure 1.4a)19a and the carrier mobility of this material was reported to be in the range of 10-4 cm2/Vs.19b However, due to the lack of processablitiy of this polymer, the semiconducting film was prepared via direct polymerization on a substrate, which might

24

have lead to the presence of various defects which act as traps and limit the mobility. Field-effect mobilities of the order of 10-6 - 10-5 cm2/Vs for undoped polyaniline and 10-5 - 10-3 for polyaniline doped with 2,4,5-trichlorobenzene sulphonic acid have also been reported in the literature (Figure 1.4b).19f In an interesting recent study it has been reported that electrospun camphor-sulphonic-acid-doped polyaniline/polyethylene oxide nanofibers display FET behaviour with field-effect mobilities around 10-4 cm2/Vs.19g In the case of poly(p-phenylene vinylene)s (PPV), widely used as the electroluminescent layer in LEDs, field-effect mobilities (holes) of the order of 10-8 - 10-7 cm2/Vs were reported upon doping with iodine (Figure 1.4c).19c Also in this case, the PPV was polymerized directly on to the substrate using a precursor route, which might lead to various structural defects and impurities in the semiconducting layer. A time-of-flight (TOF) study carried out with a PPV synthesized using a different protocol exhibited a room-temperature charge carrier mobility of the order of 10-5 cm2/Vs at a field of 105 V/cm.19d A subsequent TOF study on phenyl amino substituted PPVs reported charge carrier mobilities in the range of 10-4 - 10-3 cm2/Vs (Figure 1.4d).19e More recent studies have shown that the charge carrier mobility is greatly enhanced in ordered structures due to improved interchain hopping. For example, charge carrier mobilities of the order of 0.1 cm2/Vs can be achieved at room temperature in selforganized semicrystalline films of poly(3-alkylthiophene)s (PATs).18e,20a In pioneering studies by Lovinger et al.18e and Sirringhaus et al. 20a the charge carrier mobility in PATs was related to the degree of self ordering observed in these materials. It was reported that in case of polymers with high regioregularities that the interdigitated side chains tend to orient perpendicular to a substrate and that -stacked conjugated polymer lamellae

25

deposit parallel to the substrate, leading to highly efficient in-plane charge transport
H N n n n

OC8H17 N n S

S n

H17C8O

H17C8

C8H17

d
O N N O N N n O N N

e
O N N

Figure 1.4 Chemical structures of (a) polyacetylene, (b) polyaniline, (c) poly(pphenylene vinylene), (d) poly(N-phenylimino-1,4-phenylene-1,2ethenylene-1,4-(2,5-dioctoxy)-phenylene-1,2-ethenylene-1,4-phenylene), (e) poly-9,9dioctyl-fluorene-co-bithiophene, (f) poly(benzobisimidazo benzo-phenanthroline) and (g) poly(benzobisimidazole).

(Figure 1.5).20a It has also been reported that the on-off ratio of field-effect transistors depends on the orientation of the polymer on the substrate and increases with better inplain orientation.18e Recently it was reported that the increase of the charge carrier mobility with increasing molecular weight coincided with the amount of overlap between the ordered regions of the film. It was demonstrated for low-molecular-weight PATs that the induced orientational order (thus the overlap) between aggregates is low when processing them

26

from a low-boiling-point solvent such as chloroform. However, the charge carrier

S S S S S S

b
S S

S S S

Figure 1.5 Two different preferential orientations of the ordered domains of PAT with respect to the employed substrate. (a) In-plain orientation and (b) out-ofplain orientation.

mobility was observed to increase if these films were annealed or processed from a highboiling-point solvent such as xylene or 1,2,4-trichlorobenzene, this was attributed to better ordering and overlap between the aggregates.20c,d Using a rigid-rod nematic conjugated polymer, poly9,9dioctyl-fluorene-cobithiophene (Figure 1.4e), it has also been demonstrated that the charge carrier mobility is enhanced upon macroscopic alignment of the conjugated polymer.20e While field-effect

27

transistors based on isotropic layer of this polymer displayed a charge carrier mobility of 0.003 - 0.005 cm2/Vs, the carrier mobility was increased to 0.009 - 0.02 cm2/Vs for devices in which the polymer chains had been aligned parallel to the source-drain gap by using an alignment layer and quenching the polymer (after annealing in a thermotropic liquid crystalline state) into an ordered glass. In the case of aligned samples a significant anisotropy for field-effect mobilities (/) was also noted.20e In an interesting study by Moses and coworkers it was reported that charge carrier mobilities of up to 0.2 cm2/Vs could be obtained in the case of field-effect transistors based on PATs, if an ultrathin film (20 40 ) of the semiconducting polymer was processesed using a dip-coating technique.20f This significant enhancement of the charge carrier mobility was explained with a significantly improved structural order of the conjugated macromolecules near the polymer - insulator interface.20f This charge carrier mobility is to our best knowledge, the highest that has been reported to date for conjugated polymers. All the above examples are p-type materials and the quoted mobilities are related to the transport of holes. Rather interestingly, none of these conjugated polymers displayed significant n-type transport, a behaviour that was attributed to the existence of strong traps for electrons but not for holes.21a Although a number of n-type organic molecules and oligomers with high electron mobility have been reported,21a,b,c the realization of high electron mobilities in conjugated polymers remains an important goal for the implementation of high performance plastic electronics. High electron mobilites up to 0.1 cm2V-1s-1 were recently reported by Babel et al. for a high-electron-affinity conjugated ladder polymer, poly(benzobisimidazobenzo

28

phenanthroline) (BBL Figure 1.4f).21d These values are significantly higher than the mobility reported for the non-ladder type analogue poly(benzobisimidazole) (Figure 1.4g) which was of the order of 10-6 cm2/Vs. This significant difference was explained with the ordered semicrystalline morphology observed in the ladder-type polymers which is in contrast with the amorphous nature of films obtained from the non-ladder poly(benzobisimidazole). In a very recent study, n-type behaviour was demonstrated in field-effect transistors based on a variety of conjugated polymers that earlier were reported to be good hole conductors only. The key for efficient electron transport was reportedly the use of a hydroxyl-free gate dielectric based on divinyltetramethysiloxane-(bis benzyl cyclobutene). The reported electron mobilities were in the range of 10-3 - 10-2 cm2/Vs for unaligned poly(fluorene) copolymers and dialkyl substituted PPVs.21e It was also reported that the lack of n-type transport in the FETs studied earlier was due to the electrochemical electron trapping by the hydroxyl group, occurring in the SiO2, poly(vinyl alcohol) or polyimide gate dielectrics. In summary, good (in-plain) charge transport is usually desired in FETs and other applications. While the above discussed approach of ordering/orienting conjugated polymers indeed leads to materials with significantly improved charge carrier mobilities, the required processing protocols are usually intricate and incompatible with preferred processes for plastic electronic manufacturing which include, inkjet printing,22,23 screen printing,24,25 roll to roll processes26 and others. Thus an alternative general approach for the improvement of the charge transport in these materials is desirable.

29

1.3 Conjugated Polymer Networks It can be expected that the charge transport characteristics (and other properties) of conjugated polymers can also be improved through the introduction of - conjugated cross-links between the conjugated macromolecules.18c One may surmise that in an ideal -conjugated macromolecular network, which features conjugated cross-links (Scheme 1.1), intrachain diffusion is the predominant mechanism for charge transport, while

Scheme 1.1 Simplified schematic representation of cross-linked conjugated polymer networks with organometallic cross-links (top) and covalent cross-links (bottom).

interchain processes, if at all, only play a subordinate role. Thus, this architecture might allow the creation of polymeric materials with most interesting characteristics, including

30

high charge carrier mobility, good electrical conductivity, and also large nonlinear optic response.18c,27 The approach may represent an attractive technological alternative to the aforementioned systems with high molecular order and/or orientation,18,20 which often require rather demanding processing conditions. As shown in Scheme 1.1 the networks can be designed to rely on either non-covalent or covalent interactions. In the first case a moiety is integrated in the conjugated polymer backbone that acts as a binding site for a metal and the network formation occurs through coordination bonds that provide adequate electronic interactions, i.e. conjugation. The second case takes inspiration from the framework used for the synthesis of known thermosetting polymers and is based on the introduction of a conjugated tri-functional monomer along with the conjugated bifunctional monomers. Obviously, this approach leads to a (non-processable) threedimensional cross-linked conjugated polymer network with covalent cross-links which has to be processed prior to or during network formation. Interestingly, despite the diverse research activities focused on the chemistry, material science and physics of conjugated polymers, the most interesting feature of conjugated cross-links has received little attention, at least as far as systematic studies and well-defined systems are concerned. This situation may be a direct consequence of the challenge to introduce such cross-links and retain adequate processability. While conjugated-polymer-based networks featuring non-conjugated cross-links based on covalent28 or non-covalent bonds29,30 have been deliberately prepared and studied by a number of research groups, examples of conjugated cross-links involving either covalent bonds31,32 or metal-complexes between chains33,34 are rare, and in many cases have been obtained serendipitously and lack unambiguous characterization. The knowledge base

31

generated through these studies however is important for the present thesis, as it represents the starting point for this research endeavor. In an important study, Joo et al. have compared the electronic characteristics of polypyrrole samples, which feature different degrees of conjugated side chains and/or
H N N H

H N H N N H NH HN NH NH N H H N H N N H

H N N H

H N H N H N N H HN H N H N N H

NH

Figure 1.6 Schematic representation of the chemical structure of conjugated cross-links in electrochemically synthesized polypyrrole.

cross-links (Figure 1.6). Unfortunately, the analytic techniques employed in this study do not allow an unambiguous discrimination between originally unintentionally introduced side chains and covalent cross-links. However, highest conductivities were found for the material for which the highest cross-link density was assumed.32b In another study by Krebs et al. it was found that the residual Pd0 catalyst trapped in the polymer after workup led to cross-linking of zinc phorphyrin-linked poly(p32

phenylene ethynylene).35 It was suggested that the cross-linking reaction caused the formation of substituted benzenes, which were formed from enynes and activated alkynes in a [4+2] cycloaddition reaction and from terminal alkynes in a [2+2+2] cycloaddition reaction in the presence of Pd0 without an inert atmosphere (Scheme 1.2). Unfortunately, the electronic properties of these materials have not been reported.

C8H17 2 I H17C8 Pd n H

C8H17

C8H17 H17C8 Pd

H17C8

C8H17

C8H17

H17C8 C8H17

H17C8

H17C8 H17C8

C8H17

Scheme 1.2 Schematic representationof the cross-linking reaction reported to occur in zinc phorphyrin-linked poly(p-phenylene ethynylene)s.

33

An interesting study by Lavastre and co-workers reported the formation of conjugated polymer networks through heat treatment of poly[(4-ethynyl)phenylacetylene] (Scheme 1.3).36 The cross-linking reactions were studied via thermo-gravimetric techniques, and the resulting product obtained was reported to be insoluble. Poor solubility of poly[(4-ethynyl)phenylacetylene] in common organic solvents was also mentioned. However, a thorough chemical characterization elucidating the role of intermolecular reactions as compared to the intramolecular reactions, determination of the cross-link density as well as the electronic and photophysical properties of the crosslinked products have not been reported.
n

300 oC

H n

Scheme 1.3 Schematic representation of the cross-linking reaction proposed to occur in poly[(4-ethynyl)phenylacetylene] upon thermal treatment.

The introduction of transition metals into conjugated polymers has also recently received considerable attention in particular due to the potential to manipulate the electronic properties of these materials.37,38 Conventional concepts for the design of conjugated organometallic polymers include the incorportation of metal centers into polymers either integrated directly in the polymer chain, or bound coordinatively to the 34

conjugated backbone,34,38,39,40 or alternatively, they can be attached via conjugated or nonconjugated spacer units in the form of side groups.40
- CO m Cr(CO)3 n Cr(CO)2 m n

Scheme 1.4 Schematic representation of the proposed cross-linking reaction occurring upon heat treatment of Cr(CO)3-benzene containing poly(arylene ethynylene)s.

A study by Wright claimed that upon thermal treatment or UV irradiation of poly(arylene ethynylene)s containing the Cr(CO)3benzene moiety, cross-linking occurred with the loss of CO and the formation of phenylene-Cr(CO)2-ethynyl cross-links (Scheme 1.4). However the linear polymers synthesized by Wright exhibited low solubility and prevented the in-depth characterization of the product afforded by this process. Thus exact chemical structure and the degree of cross-linking remains somewhat questionable and also in this case no investigation of the materials charge transport characteristics was undertaken. With the objective to exploit the redox properties after coordination of transition

N PdX2 N N

N H

N H

Figure 1.7 Chemical structure of poly(o-toluidine) cross-linked with palladiumII.

35

metal having a relevant redox function to a conjugated polymer, Hirao et al. synthesized a organometallic network of poly(o-toluidine) with PdII coordinating to the imine moieties in the polymer (Figure 1.7). The linear polymers synthesized were reported not to be completely soluble and only dissolved fractions were used for the study. These species having number average molecular weight of ~3000 were further observed to polymerize during experiments conducted to study the electrochemical behaviour of the systems. In another study conducted in the Weder group, which nucleated the present thesis, Huber et al. demonstrated that the unsaturated (acyclic) carbon-carbon bonds in the poly(p-phenylene ethynylene) (PPE) backbone can be utilized as binding motif.41 The conjugated polymer employed in the reported work was poly[2,5-dioctyloxy-1,4diethynyl-phenylene-alt-2-methoxy,5-2'-ethylhexyloxy-1,4-phenylene] (MEH-OPPE, 1), a highly soluble PPE and offers two ethynylene moieties per repeat unit as potential ligand sites (Scheme 1.5). The dinuclear [Pt-(-Cl)Cl(PhCH=CH2)]2 (2)42 was employed as the cross-linker in the initial experiments. It was demonstrated that the ethynylene moieties comprised in the PPE can readily coordinate to PtII, in exchange with weaklybound styrene ligands. The bifunctional [Pt-(-Cl)Cl(PhCH=CH2)]2 employed under appropriate conditions indeed allowed the formation of PPE-Pt networks (Scheme1.5, 3, z>0), as demonstrated in an extensive NMR study. In dilute solutions, the equilibrium of the investigated PPE-Pt systems dictated non-cross-linked structures (Scheme 1.5, 3, z=0), and the system remained homogeneous and therewith processable. Spin-coating resulted in films of good optical quality, which were unequivocally cross-linked. While the latter were completely insoluble in solvents that would readily dissolve the uncomplexed polymer, they readily dissolved upon addition of an excess of an olefinic

36

ligand (e.g. styrene), demonstrating the reversibility of the complexation. As expected, the coordination of PtII markedly influences the photophysical characteristics of the PPE; its PL is efficiently quenched, and at high Pt-contents, the absorption maximum

+
n
O

Cl Pt Cl Cl Pt

Cl

OR

OR

OR

x
RO RO

y Cl Cl Pt Pt Cl Cl RO Cl Cl Pt Pt Cl

z n Cl OR

RO

Scheme 1.5 Simplified representation of the ligand-exchange reaction between MEHOPPE (1) and [Pt-(-Cl)Cl(PhCH=CH2)]2 (2), leading to cross-linked organometallic hybrid materials 3.

experienced a hypsochromic shift. Thus, it was successfully demonstrated that welldefined organometallic polymer networks can be easily synthesized and processed by employing carefully designed ligand-exchange reactions, but again no charge transport studies were conducted for the materials employed. It should be noted that while this study has an important character for the present thesis,41 the Cl-bridged dinuclear joints

37

employed in the above studies, may not provide significant -conjugation between tween chains. In parallel to the experiments summarized above, Wolf et al.30b recently synthesized polythiophenes, which are cross-linked via a dinuclear Pd complex (Figure 1.8). Their approach relies on the electropolymerization of a dimeric metal-terthiophene complex (which, remarkably, was rather similar to the Pt-complex employed) and, conceptually, provides an alternative access path to these target-structures; however, it lacks the freedom which the ligand-exchange approach imparts to the processing of these materials.

S S Pd P Ph S Ph Ph Cl P Pd Cl S S

Ph

Figure 1.8 Bridged Pd terthiophene complex employed for electropolymerization.

Thus in summary, the knowledge base regarding the possibility to synthesize and process -conjugated materials which feature conjugated cross-links is very limited. Concomitantly, the question of how this structural motif can be employed to design polymeric materials with optimized optic and electronic characteristics remains essentially unanswered.

38

1.4 Poly(p-phenylene ethynylene)s: An Important Class of Conjugated Polymers Among a plethora of materials platforms, poly(arylene ethynylene) (PAE) derivatives have attracted the attention of numerous research groups and we elected to exemplarily employ this family of conjugated polymers as the basis for the present study. As the name implies, PAEs feature aromatic rings and ethynylene groups in the polymer backbone. The connection of these moieties
Ar n PA PAV n PDA n

results

in

an

alternating

Ar n PAE

Ar

Figure 1.9 General schematic structures of poly(arylene ethynylene)s (PAEs), poly(arylene)s (PAs), poly(arylene vinylene)s (PAVs), and poly(diacetylene)s (PDAs).

sequence of single- and multiple bonds and gives rise to -conjugation along the macromolecules. PAEs are closely related to poly(arylene)s (PAs),43 poly(arylene vinylene)s (PAVs) and poly(diacetylene)s,44 which all represent important classes of conjugated polymers (Figure 1.9). The chemical structure of PAEs can readily and significantly be manipulated, for example via the choice of the aromatic moiety, the connectivity of the latter (e.g. meta vs. para substitution), the introduction of heteroatoms or metals, the nature of solubilizing side chains, and non-covalent interactions with metals. The possibility to integrate conjugated moieties other than arylenes and ethynylenes (for example vinylene groups, non-conjugated aliphatic spacers, etc.) represents another synthetic tool, which leads to PAE copolymers. These structural

39

changes allow one to tailor the property matrix of these polymers over a wide range. Hundreds of different PAEs and PAE copolymers have been reported to date, and during the last ten years this family of materials has established itself as an important class of conjugated polymers with interesting optical and electronic properties. A number of excellent texts have reviewed the synthesis, physico-chemical characteristics, and optical properties of these materials. The early work on PAEs has been summarized in 1996 by Giesa.45 A review covering various aspects regarding PAEs was published by Bunz in 200046 and several complements and updates to this compilation have appeared since.47,48,49 Yamamoto has reviewed the important subject of heteroaromatic PAEs.50,51 PAE electrolytes are part of an outstanding article by Schanze.52 Swager has published a number of excellent texts that discuss the application of PAEs and other polymers in sensors.13,53,54,55 The intriguing nanoarchitectures that can be created with PAEs have been addressed by Moore.56,57 The most comprehensive review on the synthesis, properties, structures, and application of PAEs was edited by Weder et al. in 2005.58The electrically (semi)conducting nature of PAEs has traditionally received comparably little attention. However, during the last decade more and more research efforts have been devoted to this subject, and PAEs have eventually been recognized as a potentially very powerful class of polymeric semiconductors.59 Previous claims that PPEs do not function well as the emitter in light emitting devices were negated by Weder et al. who clearly demonstrated the usefulness of PPEs in LEDs.60 Ofer et al. have reported electrical conductivities for doped PPEs in the range of ~0.18 and ~4.5 Scm-1 at a potential of ~1.6 V vs. SCE. The doping was performed under rather severe conditions at -70C in liquid SO2 comprising [(n-Bu)4N]AsF6 as electrolyte, due to the high oxidation and reduction

40

potentials observed for the PPEs.61 The conductivities reported in this study are the highest reported to date for poly(p-phenylene ethynylene)s.

Figure 1.10 Chemical structure of poly[2,5-dioctyloxy-1,4-diethynyl-phenylene-alt-2,5bis(2'-ethylhexyloxy)-1,4-phenylene] (EHO-OPPE, 4).

Owing to the linear rigid rod backbone and - stacking, unsubstituted PPEs are intractable i.e. they are insoluble and as well cannot be thermally processed.62 To increase the solubility of PPEs one of the strategies adopted is to introduce solubilizing side chains on the aromatic rings in the polymer backbone. Poly[2,5-dioctyloxy-1,4-diethynylphenylene-alt-2,5-bis(2'-ethylhexyloxy)-1,4-phenylene] (EHO-OPPE, 4), a PPE

derivative consisting of alternating units substituted with sterically hindered and linear alkoxy side chains (Figure 1.10) displays an interesting array of characteristics. Employing facile synthetic protocols EHO-OPPE can be directly synthesized in high purity without the need for precursor polymers.63 It has been reported by Weder et al. that this derivative of PPE has a high PL quantum efficiency of around 85% in solution, around 35% in the solid state and can be used in LEDs as the electroluminescent layer.

41

Owing to the rigidity and the linearity of the conjugated backbone, EHO-OPPE displays outstanding orientability of the molecules. In a clever study by Weder et al this concept was exploited for fabrication of devices in which polarized absorption and / or emission of light are desired.64. Interesting NLO properties have also been reported with extremely efficient two photon absorption for EHO-OPPE.65 Owing to this interesting property matrix EHO-OPPE was elected to be employed as the central materials platform for the present experimental study that forms the basis of this thesis.

1.5 Energetics of Conjugated Polymer Networks As discussed already in Chapter 1.3 the introduction of conjugated cross-links may play a significant role in the improvement of inter-chain charge transfer processes and thus effectively improve charge transport in conjugated macromolecules. Matched energy levels (viz. HOMO and LUMO or ionization potential or electron affinity) of both the conjugated polymer and the employed cross-linker are required for efficient transport to occur. In the case of conjugated polymer with no chemical defects due to the presence of isoenergetic electronic states the charge transport is not hindered, however the presence of a chemical defect in the conjugated backbone may introduce energy levels with energies differing from the polymer main chain that can act as traps and hinder the charge transport through the backbone (Figure 1.11a,c). Low lying energy levels compared to the HOMO can act as electron traps (Figure 1.11b) and those compared to LUMO can act as hole traps (Figure 1.11d). In the case of covalently cross-linked conjugated polymers by careful selection of cross-linking moieties with similar chemical structure as the repeating unit this possible energy mismatch can be avoided.

42

Figure 1.11 Simplified schematic of electron (top) and hole (bottom) transfer in conjugated polymers with matched energy levels (a and c) and un-matched energy levels (b and d).

Recently, Swager and Holliday discussed the different electron transport mechanisms in conducting metallopolymers by outer and inner sphere electronic coupling between the metal center and conjugated polymer.66 Charge transport by outer sphere electronic coupling is predominantly observed in the case of redox active metal centers tethered to conjugated polymer backbone via a conjugated or non-conjugated linker (Figure 1.12). Thus in such materials there is a lack of electronic interaction between the

43

metals d-orbitals and the conjugated polymers -system. In the case of organometallic complexes an isoenergic series of electronic states exits, which facilitates the charge transport through the system. When these organometallic complexes are tethered to the conjugated polymer (possessing its own band of electronic states) these isoenergic electronic states remain un-affected and distinct, thus as a result these metal complexes

Figure 1.12 Schematic structure of metallopolymer demonstrating charge transport by outer sphere electronic coupling.

may not be significantly involved in the macroscopic charge transport through the material. The redox and the charge transport properties of these hybrid materials depend on the length and the nature of the tether employed for appending the metal center to the conjugated polymer. It has been reported before that as the length of the tether is decreased the conductivity observed in these systems switches from the self-exchange mechanism between the attached organometallic moieties to one involving the conjugated polymer backbone.67 Metallopolymers with the transition metal centers incorporated in the polymer backbone (Figure 1.13) predominantly display charge transport via inner sphere electronic coupling. Intimate mixing of the metals d orbitals and the polymers system 44

Figure 1.13 Schematic structure of metallopolymer demonstrating charge transport by inner sphere electronic coupling.

is observed in this case and the metal centers can from the part of the intrachain charge transport. In such metallopolymers a variety of ligands can be employed for the coordination of the metal center to the polymer backbone. Importantly, while there may be a direct overlap between the orbitals of the polymer backbone and the transition metal in these materials, the electronic properties of the system are still highly dependent on the electronic levels of the involved polymer and metal. To obtain optimum electronic properties in these systems matched energy levels of the conjugated polymer and the metal center are an important pre-requisite.

45

1.6 References 1 Shirakawa, H.; Louis, E. J.; MacDiarmid, A. G.; Chiang, C. K.; Heeger, A. J. J. Chem. Soc. Chem. Comm. 1977, 578. 2 (a) Heeger, A. J. J. Phys. Chem. 2001, 105, 8475. (b) MacDiarmid, A. G. Rev. Mod. Phys. 2001, 73, 701. (c) Shirakawa, H. Rev. Mod. Phys. 2001, 73, 713. 3 Skotheim, T. J.; Elsenbaumer, R. L.; Reynolds, J. R. Eds. Handbook of conducting polymers 2nd edn. Dekker, New York, 1998. 4 5 6 Kraft, A.; Grimsdale, A. C.; Holmes, A. B. Angew. Chem. Int. Ed. 1998, 37, 403. Mitschke, U.; Buerle, P. J. Mater. Chem. 2000, 10, 1471. Greiner, A.; Weder, C. Light-emitting diodes, In: Kroschwitz J. I. Ed. Encyclopedia of polymer science and technology 3rd edn. Wiley-Interscience, New York, 2003, Vol. 3 p 87. 7 8 9 10 Horowitz, G. Adv. Mater. 1998, 10, 365. Moerner, W. E.; Silence, S. M. Chem. Rev. 1994, 94, 127. Brabec, C. J.; Sariciftci, N. S.; Hummelen, J. C. Adv. Funct, Mater. 2001, 11, 15. Nalwa, H. S. Ed. Handbook of organic conductive molecules and polymers, John Wiley & Sons, New York, 1996. 11 Mark, J. A. Ed. Physical properties of polymers handbook, American Institute of Physics, New York, 1996. 12 Tallman, D. E.; Spinks, G.; Dominis, A.; Wallace, G. G. J. Solid State Electrochem. 2002, 6, 73. 13 Swager, T. M. Chem. Res. Toxicol. 2002, 15, 125.

46

14

(a) Karg, S.; Riess, W.; Meier, M.; Schwoerer, M. Synth. Met. 1993, 55-57, 4186. (b) Hoeben, F. J. M.; Jonkheijm, P.; Meijer, E. W.; Schenning, A. P. H. J. Chem. Rev. 2005, 105, 1491.

15

Roth, S; Caroll, D. Eds. One-dimensional metals, 2nd edn. Wiley-VCH, Weinheim, 2004.

16 17 18

Peierls, R. E. Quantum Theory of Solids, Clarendon Press, Oxford, 1955. MacDiarmid, A. G. Angew. Chem. Int. Ed. 2001, 40, 2581. (a) Cao, Y.; Andreatta, A.; Heeger, A. J.; Smith, P. Polymer 1989, 30, 2305. (b) Wang, Z. H.; Li, C.; Scherr, E. M.; MacDiarmid, A. G.; Epstein, A. G. Phys. Rev. Lett. 1991, 66, 1745. (c) MacDiarmid, A. G.; Min, Y.; Wiesinger, J. M.; Oh, E. J.; Scherr, E. M.; Epstein, A. J. Synth. Met. 1993, 55, 753. (d) Joo, J.; Oblakowski, Z.; Du, G.; Pouget, J. P.; Oh, E. J.; Weisinger, J. M.; Min, Y.; MacDiarmid, A. G.; Epstein, A. J. Phys. Rev. B 1994, 49, 2977. (e) Bao, Z.; Dodabalapur, A.; Lovinger, A. J. Appl. Phys. Lett. 1996, 69, 4108. (f) Aleshin, A.; Kiebooms, R.; Menon, R.; Wudl, F.; Heeger, A. J. Phys. Rev. B 1997, 56, 3659. (g) Aleshin, A.; Kiebooms, R.; Reghu, M.; Heeger, A. J. Synth. Met. 1997, 90, 61. (h) Kim, J. H.; Sung, H. K.; Kim, H. J.; Yoon, C. O.; Lee, H. Synth. Met. 1997, 84, 71.

19

(a) Ebisawa, F.; Kurokawa, T.; Nara, S. J. App. Phys. 1983, 54, 3255. (b) Burroughes, J. H.; Jones, C. A.; Friend, R. H. Nature 1988, 335, 137. (c) Pichler, K.; Jarett, C. P.; Friend, R. H.; Ratier, B.; Moliton, A. J. Appl. Phys. 1995, 77, 3523. (d) Lebedev, E.; Dittrich, T.; Petrova-Koch, V.; Karg, S.; Brtting, W. Appl. Phys. Lett. 1997, 71, 2686. (e) Hertel, D.; Bssler, H.; Scherf, U.; Hrhold, H. H. J. Chem. Phys. 1999, 110, 9214. (f) Kho, C. T.; Chiou, W. H. Synth. Met. 1997, 88,

47

23. (g) Pinto, N. J.; Johnson, A. T. Jr.; MacDiarmid, A. G.; Mueller, C. H.; Theofylaktos, N.; Robinson, D. C.; Miranda, F. A. Appl. Phys. Lett. 2003, 83, 4244. 20 (a) Sirringhaus, H.; Brown, P. J.; Friend, R. H.; Nielsen, M. M.; Bechgaard, K.; Langeveld-Voss, B. M. W.; Spiering, A. J. H.; Janssen, R. A. J.; Meijer, E. W; Herwig, P.; de Leeuw, D. M. Nature 1999, 401, 685. (b) sterbacka, R.; An, C. P.; Jiang, X. M.; Vardey, Z. V. Science 2000, 287, 839. (c) Kline, R. J.; McGehee, M. D.; Kadnikova E. N.; Liu, J.; Frchet, J. M. J.; Toney, M. F. Macromolecules 2005, 38, 3312. (d) Chang, J. F.; Sun, B.; Breiby, D. W.; Nielsen, M. M.; Slling, T. I.; Giles, M.; McCulloch, I.; Sirringhaus, H. Chem. Mater. 2004, 16, 4772. (e) Sirringhaus, H.; Wilson, R. J.; Friend, R. H.; Inbasekaran, M.; Wu, W.; Woo, E. P.; Grell, M.; Bradley, D. D. C. Appl. Phys. Lett. 2000, 77, 406. (f) Wang, G.; Swensen, J.; Moses, D.; Heeger, A. J. J. Appl. Phys. 2003, 93, 6137. 21 (a) Dimitrakopoulos, C. D.; Malenfant, P. R. L. Adv. Mater. 2002, 14, 99. (b) Katz, H. E.; Lovinger, A. J.; Johnson, J.; Kloc, C.; Siegrist, T.; Li, W.; Lin, Y. Y.; Dodabalapur, A. Nature 2000, 404, 478. (c) Pappenfus, T. M.; Chesterfield, R. J.; Frisbie, C. D.; Mann, K. R.; Casado, J.; Raff, J. D.; Miller, L. L. J. Am. Chem. Soc. 2002, 124, 4184. (d) Babel, A.; Jenekhe S. A. J. Am. Chem. Soc. 2003, 125, 13656. (e) Chua, L. L.; Zaumsell, J.; Chang, J. F.; Ou, E. C. W.; Ho, P. K. H.; Sirringhaus, H.; Friend, R. H. Nature, 2005, 434, 194. 22 Kawase, T.; Shimoda, T.; Newsome, C.; Sirringhaus, H.; Friend, R. H. Thin Solid Films, 2003, 438 - 439, 279. 23 Wang, J. Z.; Zheng, Z. H.; Li, H. W.; Huck, W. T. S.; Sirringhaus, H. Nature Mat. 2004, 3, 171.

48

24

Bao, Z.; Feng, Y.; Dodabalpur, A; Raju, V. R.; Lovinger, A. J. Chem. Mater. 1997, 9, 1299.

25 26

Bao, Z.; Rogers, J. A.; Katz, H. E. J. Mater. Chem. 1999, 9, 1895. Bcklund, T. G.; Sandbert, H. G. O.; sterbacka, R.; Stubb, H.; Mkel, T.; Jussila, S. Synth. Met. 2005, 148, 87.

27

Brdas, J. L.; Meyers, F.; Heeger, A. J. Conjugated Polymers: On the Parallel Between the Electrical Conduction Mechanism and the Nonlinear Optical Response. In Organic Molecules for Nonlinear Optics and Photonics Messier, J.; Prasad, P.N.; Kajzar, F. Eds. Kluwer Academic Publishers: Dordrecht, 1991, 25.

28 (a) Li, X. C.; Yong, T. M.; Gruner, J.; Holmes, A. B.; Moratti, S. C.; Cacialli, F.; Friend; R. H. Synth. Met. 1997, 84, 437. (b) Liu, G.; Freund, M. S. Macromolecules 1997, 30, 5660. (c) Roitman, D. B.; Antoniadis, H,; Helbing, R.; Pourmizaie, F.; Sheats, J. R. Proc. SPIE 1998, 3476, 232. (d) Muller, D.; Gross, M.; Meerholz, K.; Braig, T.; Bayerl, M. S.; Bielefeldt, F.; Nuyken, O. Synth. Met. 2000, 111, 34. Schanze, K. S.; Bergstedt, T. S.; Hauser, B. T.; Cavalaheiro, C. S. P. Langmuir 2000, 16, 795. (e) Inaoka, S.; Advincula, R. Macromolecules 2002, 35, 2426. (f) Taranekar, P.; Baba, A.; Fulghum, T. M.; Advincula, R. Macromolecules 2005, 38, 3679. 29 Angelopoulos, M.; Dipietro, R.; Zheng, W. G.; MacDiarmid, A. G.; Epstein, A. J. Synth. Met. 1997, 84, 35. 30 (a) Deronzier, A.; Moutet, J. C. Coord. Chem. Rev. 1996, 147, 339. (b) Clot, O.; Wolf, M. O.; Patrick, B. O. J. Am. Chem. Soc. 2000, 122, 10456. 31 Kumar, A.; Reynolds, J. R. Macromolecules 1996, 29, 7629.

49

32

(a) Joo, J.; Lee, J. K.; Lee, S. Y.; Jang, K. S.; Oh, E. J.; Epstein, A. J. Macromolecules 2000, 33, 5151. (b) Joo, J.; Lee, J. K.; Baeck, J. S.; Kim, K. H.; Oh, E. J.; Epstein, A. J. Synth. Met. 2001, 117, 45.

33 34 35 36

Wright, M. E. Macromolecules, 1989, 22, 3256. Hirao, T.; Yamaguchi, S. Fukuhara, S. Tetetrahedron Lett. 1999, 40, 3009. Nielsen, K. T.; Spanggaard, H.; Krebs, F. C. Macromolecules 2005, 38, 1180. Lavastre, O.; Cabioch, S.; Dixneuf, P. H.; Sedlacek, J.; Vohlidal, J. Macromolecules 1999, 32, 4477.

37 38 39

Manners I. Angew. Chem. Int. Ed. Engl. 1996, 35, 1062. Rehahn M. Acta Polymer 1998, 49, 201. Ley, K. D.; Whittle, C. E.; Bartberger, M. D.; Schanze, K. S. J. Am. Chem. Soc. 1997, 119, 3423.

40 41 42 43 44

Wong C. T.; Chan W. K. Adv. Mater. 1999, 11, 455. Huber, C.; Bangerter, F.; Caseri, W.; Weder, C. J. Am. Chem. Soc. 2001, 123, 3857. Albinati, A.; Caseri, W. R.; Pregosin, P. S. Organometallics 1987, 6, 788. Scherf, U. Top. Curr. Chem. 1999, 201, 163. Cantow, H. J. Ed. Advances in Polymer Science, Vol 63, Springer-Verlag, Berlin, 1998.

45 46 47

Giesa, R. J. M. S. Rev. Macromol. Chem. Phys. 1996, C36, 631. Bunz, U. H. F. Chem. Rev. 2000, 100, 1605. Bunz, U. H. F. Acc. Chem. Res. 2001, 34, 998.

50

48

Bunz, U. H. F. 2002, The ADIMET reaction: synthesis and properties of poly(dialkylparaphenyleneethynylene)s. In: chemistry. VCH-Wiley, 217. Astruc, D. Ed., Modern arene

49

Bunz, U. H. F. 2003, Acyclic diyne metathesis utilizing in situ transition metal catalysts: an efficient access to alkyne-bridged polymers. In: Grubbs, R. H. Ed., Handbook of metathesis. VCH-Wiley, Vol. 3: 345.

50 51 52 53 54 55 56 57

Yamamoto, T. Bull. Chem. Soc. Jpn. 1999, 72, 621. Yamamoto, T. Synlett. 2003, 425. Pinto, M. R.; Schanze, K. S. Synthesis 2002, 1293. Swager, T. M. Acc. Chem. Res. 1998, 31, 201. McQuade, D. T.; Pullen, A. E.; Swager, T. M. Chem. Rev. 2000, 100, 2537. Wosnick, J. H.; Swager, T. M. Curr. Opin. Chem. Biol. 2002, 4, 715. Moore, J. S. Acc. Chem. Res. 1997, 30, 402. Hill D. J.; Mio M. J.; Prince R. B.; Hughes T. S.; Moore J. S. Chem. Rev. 2001, 101, 3893.

58

Weder, C. Ed. Poly(arylene ethynylene)s From Synthesis to Applications; Advances in Polymer Science Series Vol. 177; Springer, Heidelberg, 2005.

59

Voskerician, G.; Weder, C. Electronic Properties of PAEs. In: Weder, C. Ed. Poly(arylene ethynylene)s From Synthesis to Applications; Advances in Polymer Science Series Springer, Heidelberg, 2005, Vol. 177: 209.

60 61 62

Montali, A.; Smith, P.; Weder, C. Synth. Met. 1998, 97, 123. Ofer, D.; Swager, T. M.; Wrighton, M. S. Chem. Mater. 1995, 7, 418. Trumbo, T. R.; Marvel, C. S. J. Polym. Sci. Part A, 1987, 25, 1027.

51

63 64

Weder, C.; Wrighton, M. S. Macromolecules 1996, 29, 5157. Weder, C.; Sarwa, C.; Montali, A.; Bastiaansen, C.; Smith, P. Science 1998, 279, 835.

65

Weder, C.; Wrighton, M. S.; Spreiter, R.; Bosshard, C.; Gnter, P. J. Phys. Chem. 1996, 100, 18931.

66 67

Holliday, B. J.; Swager, T. M. Chem. Commun. 2005, 23. Wolf, M. O. Adv. Mater. 2001, 13, 545.

52

Chapter 2: Scope and Objectives Motivated by the significant interest in conjugated polymers with high charge carrier mobility, this thesis is focused on the design, synthesis, processing and characterization of cross-linked conjugated polymer systems or conjugated polymer networks. As summarized in Chapter 1.2, the charge carrier mobility in conjugated polymers is typically governed by the disorder effects. While exciting progress has been documented in the case of semiconducting polymer systems with high degree of molecular order and/or orientation, the high carrier mobility in such systems comes at the expense of elaborate processing protocols that are rather incompatible with low-cost manufacturing processes such as inkjet printing, screen printing or roll-to-roll processing. In an orthogonal approach we postulated that the charge carrier mobility (and other properties) in conjugated polymers can also be improved by the incorporation of conjugated crosslinks. It was the objective of the present thesis to prove this hypothesis. With this overall goal in mind, an experimental research program was designed that sought to (a) develop synthetic routes for the prepartation of well-defined conjugated polymer networks with covalent as well as non-covalent conjugated cross-links; (b) model adequate processing protocols for these materials; and (c) conduct in-depth studies of the charge transport in these new materials. Among the various families of conjugated polymers, poly(p-phenylene ethynylene)s (PPE) encompass an interesting array of materials properties as summarized earlier in Chapter 1.4, and therefore PPEs were elected as the central materials platform for the present study. Rather surprisingly, the charge transport properties of

representatives of this family have hitherto not been studied. Thus, in order to develop a

53

thorough understanding of this most important aspect, Chapter 3 presents an in-depth study on the charge transport of poly[2,5-dioctyloxy-1,4-diethynyl-phenylene-alt-2,5bis(2'-ethylhexyloxy)-1,4-phenylene] (EHO-OPPE), on the basis of time-of-flight (TOF) experiments. In order to investigate the influence of presumably non-conjugated cross-links on charge transport characteristics of EHO-OPPE, Chapter 4 presents a study of the charge transport in EHO-OPPE-PtII networks that were synthesized according to previously reported protocols via ligand-exchange reactions between the polymer and a low molecular PtII complex. The synthesis of organometallic PPE networks that incorporate Pt0, as a presumably conjugated cross-linker is reported in Chapter 5, together with a model study on low-molecular-weight Pt0 model compounds and an in-depth characterization of the optic and electronic properties of these polymer systems. Particular attention was, of course, given to the investigation of the charge transport characterization, of the EHOOPPE-Pt0-EHO-OPPE networks produced. Incorporation of auxiliary ligands in the conjugated polymer backbone may allow an access to a plethora of metals centers for obtaining the polymer-metal networks. With this approach in mind 2,2-bipyridine, was incorporated in the PPE backbone. Chapter 6 discusses the synthesis of PPEs having varying amounts of incorporated 2,2-bipyridine as well as their organometallic networks with transition metals. Systematic in-depth characterization of optic properties of these polymers and their organometallic networks was performed and is reported in Chapter 6.

54

In an alternate approach, PPE networks containing covalent cross-links were also synthesized. Importantly to facilitate the processing, an approach was developed to synthesize millimeter to nanometer sized particles of these networks structures. The detailed synthetic protocols and the optical characterization of these materials in bulk and in the form of particles is further discussed in Chapter 7. The results obtained during the present research initiative and the insights gained in the respective Chapters are summarized in Chapter 8. In addition, possible follow-up research initiatives are also presented in the outlook.

55

Chapter 3: Charge Carrier Transport in Poly(p-phenylene ethynylene)s*

3.1 Introduction As summarized in Chapter 1.4, poly(p-phenylene ethynylene) (PPE) derivatives have attracted the attention of a number of research groups.1 Rather surprisingly, however, the charge carrier mobilities of PAEs have hitherto been completely unexplored. In this chapter, we present an in-depth study of the charge transport properties of poly[2,5-dioctyloxy-1,4-diethynyl-phenylene-alt-2,5,-bis(2'-ethylhexyloxy)-

Figure 3.1 Chemical structure of poly[2,5-dioctyloxy-1,4-diethynyl-phenylene-alt-2,5bis(2'-ethylhexyloxy)-1,4-phenylene] (EHO-OPPE, 4), the poly(p-phenylene ethynylene) (PPE) derivative investigated here.

1,4-phenylene] (EHO-OPPE, 4, Figure 3.1),2 a soluble poly(p-phenylene ethynylene) derivative, which is representative of this family of conjugated polymers. To our best knowledge this is the first charge transport study of a PPE ever conducted.

Parts of this Chapter have been published: Kokil, A.; Shiyanovskaya, I.; Singer, K. D.; Weder, C. Synth. Met. 2003, 138, 513.

56

3.2 Experimental Section Materials and Methods All reagents and solvents were purchased from Aldrich Chemical Co. or FLUKA (analytical grade quality) and were used without further purification. EHO-OPPE (4) of a number-average molecular weight of about 10,000 gmol-1 was prepared as described before.

Preparation of TOF samples ITO / EHO-OPPE / gold sandwich structures were produced by casting a toluene solution of the polymer (10 mgmL-1) onto ITO-coated glass-slides (pre-patterned as sketched in Figure 3.2, E.H.C. Co., Japan) that were preheated to 40 C, drying the resulting films in vacuo at 40 C for at least 12 h (film thickness, L = 6.5 - 12 m as determined with a Dektak profilometer) and depositing a 3 x 3 mm2 gold electrode of a thickness of 60 nm by sputtering through a shadow-mask. To produce EHO-OPPE films with a thickness of 10 m, the polymer solution was drop-cast on the ITO coated glass slide (that was lined with scotch tape and preheated to 40 C in a vacuo) using the above mentioned processing conditions (Figure 3.2). The thickness of the scotch tape used was around 64 m. The metallic leads for the electrodes (ITO and gold) were connected to the sample using commercially available conductive epoxy (Circuit Wokrs CW2400), as per the configuration as shown in Figure 3.2. This configuration resulted in samples that did not short as compared to the ones where the conductive epoxy was applied directly on the polymer film.

57

TOF measurements The TOF measurements were conducted on a conventional setup, using a 3.6-ns Q-switched Nd-YAG (yttrium aluminum garnet) laser, which was frequency doubled and shifted using a high-pressure H2 Raman cell. The third anti-Stokes shift (320 nm) of the 532 nm pump with an intensity of < 10 J/pulse was used for the carrier generation.

Scotch Tape EHO-OPPE

ITO EHO-OPPE Gold Conductive Epoxy Leads

Figure 3.2 Schematic representation of sample preparation (left) and the architecture of the samples used for TOF measurements (right).

3.3 Charge Carrier Transport in EHO-OPPE The charge transport in EHO-OPPE (and all other materials investigated in this research project) was investigated using time-of-flight (TOF) measurements on indiumtin-oxide (ITO) / polymer / gold sandwich structures (Figure 3.3).3 In this technique, a short light pulse incident on the polymer through a semitransparent electrode creates a

58

+ + + + -

Signal for non-dispersive transport

Laser Pulse

Signal for dispersive transport

Figure 3.3 Simplified representation of the conventional time-of-flight setup employed. The transients displayed are schematic representations of the signal observed for non-dispersive transport (top) and dispersive transport (bottom).

thin sheet of charge carriers and, depending on the polarity of the electric field E applied between the electrodes, electrons or holes are driven across the sample. The absorption depth of the optical excitation is small compared to the sample thickness L (for the present polymer ca. 1 m at the wavelength of interest, cf. Ref. 2) and the duration of the optical pulse is short compared to the transit time ttr of the charge carriers. Thus, using:

= L / ttr E

Equation 3.1

the carrier mobility can be obtained from the displacement photocurrent transients. The polymer sample used in the present study was of a number-average molecular weight,

59

Mn,

of

about

10,000

gmol-1

and

was

synthesized

according

to

standard

procedures. / PPE / gold sandwich structures with a polymer film thickness between 6.5 and 12 m were produced by casting a toluene solution of the polymer onto ITO-coated glass-slides. To obtain polymer films with the desired thickness, the ITOcoated glass-slide was lined with scotch tape, before casting the polymer solution. As it was reported earlier, the rapid evaporation of the solvent lead to amorphous films, in which the macromolecules are highly disordered. Typical photocurrent transients are shown in Figure 3.4 for electrons and Figure 3.5 for holes. The shape of these curves is representative for all transients observed here and is characteristic of dispersive transport.4,5,6,7,8 This mechanism is usually observed in materials with a high degree of spatial and/or energetic disorder and is consistent with a wide variation of the local transport rates.9 Thus, in this case the charge carriers starting at the semitransparent metal electrode do not reach the ITO electrode as a sheet of charges. Large dark currents were observed for samples that were not adequately dried; this was attributed to the fact that traces of solvent significantly increases the ionic transport in the material. By contrast, dark currents of appropriately dried samples were of the order of 710-6 A. The carrier mobility was determined from the inflection point in the double logarithmic plots (Figures 3.4b and 3.5b). TOF measurements were performed as a function of carrier type, applied field and film thickness (Figure 3.6). As can be seen from Figure 3.6, the drift mobility independent of the film thickness L, demonstrating that the photocurrents are not range-limited but indeed reflect the drift of the carrier sheet across the entire sample.

60

a
Photocurrent (10 A)
-7

-2

-4

-6 0 3 6 9 -5 Time (10 s) 12 15

Photocurrent (10 A)

slope -0.03

-7

slope -0.87

Time (10 s)

-5

10

Figure 3.4 Electron time-of-flight photocurrent transients of a solution-cast film of EHOOPPE (4, L=8 m), measured at 295 K and an electric field of 2.5105 Vcm-1 in a (a) linear and (b) double logarithmic representation.

61

a
Photocurrent (10 A)
-7

6 9 -5 Time (10 s)

12

15

b
Photocurrent (10 A)
-7

slope -0.1

slope -1.2

1 -5 Time (10 s)

10

Figure 3.5 Hole time-of-flight photocurrent transients of a solution-cast film of EHOOPPE (4, L=8 m), measured at 295 K and an electric field of 2.5105 Vcm-1 in (a) linear and (b) double logarithmic representation.

62

a
ln ( cm V s )
-1 -1 2

-6.0 -6.5 -7.0 -7.5 -8.0 -8.5 -9.0 200 400 Field
1/2

600 (Vcm )
-1 1/2

800

b
ln (10 cm V s )
-1 -1 2 -3

-6.0 -6.5 -7.0 -7.5 -8.0 -8.5 -9.0 200 400 Field
1/2

600 ( V cm )
-1 1/2

800

Figure 3.6 Electron (a) and hole (b) mobilities of solution cast films of EHO-OPPE (4) as function of electric field at various film thicknesses (open triangles: L=6.5 m, filled circles: L=8 m, open squares: L=12 m).

63

High electron (2.210-3 cm2V-1s -1) and hole (1.810-3 cm2V-1s-1) mobilities were found at low field (3.1104 Vcm-1). This finding is most intriguing, since these values compare favorably with, to our knowledge, the highest values yet observed for an ambipolar conjugated polymer.10 The data plots and fits in Figure 3.7 indicate that the temperature dependence is well-described by the Gaussian disorder model presented below. As is evident from Figure 3.6, the mobility of both, electrons and holes, strongly depends on the field strength, and decreases with increasing bias. This somewhat exceptional behavior has been observed before,11,12 particularly at low fields, and is consistent with a hopping transport model that accounts for off-diagonal (positional) disorder caused by variations of intersite distances in addition to diagonal (energetic) disorder in the transport manifold.13,14 The disorder transport model considers charge transport in organic solids as time-independent random walks within a distribution of localized hopping states broadened by disorder effects. The large off-diagonal disorder may result in a negative field dependence of the mobility at low fields, because a stronger applied electric field favors forward hopping and inhibits more facile routes for carriers involving hops transverse to the applied electric field. The negative field dependence was also predicted for quasi one-dimensional transport in the presence of defects and barriers.15 To analyze the negative field dependence of the mobility in EHO-OPPE within the Gaussian disorder transport formalism and to determine the diagonal (energetic)
disorder parameter and the off-diagonal (positional) disorder parameter , we

employed the relationship between the charge mobility and the disorder parameters:13

64

a
ln (cm /Vs)

-7.2 -7.4 -7.6 -7.8 -8.0 -8.2 -8.4 -8.6 8 9 10 2 -2 10 /T (K )


6

11

-7.4 ln (cm /Vs) -7.6 -7.8 -8.0 -8.2 -8.4 8 9 10 6 2 -2 10 /T (K ) 11

Figure 3.7 Temperature dependence of (a) electron and (b) hole mobilities of an EHOOPPE (4) film (L=8 m) measured at E = 2105 (squares), 3105 (circles) and 4105 Vcm-1 (triangles).

65

( , , E ) = 0 exp exp C ( 2 2 ) E1/ 2 3

2 2

Equation 3.2

where 0 is the prefactor mobility, E is the electric field, and C is an empirical constant.
The diagonal disorder parameter is related to the width of the Gaussian density of

states , as = / kT , where k is the Boltzman constant and T is the temperature.

Equation (3.2) can be represented in the form:

ln = ( E )

1 + g(E) T2

Equation 3.3

where

(E) =

2
2

4 C E k 9

Equation 3.4

and

g ( E ) = ln 0 C E 2

Equation 3.5

The experimental temperature dependence of the mobility parametric in the applied electric field (Figure 3.6) allows for calculating and g for different values of the electric field. Because ( E ) scales linearly with
by plotting vs.
E , the value E =0 can be determined

E and then determining the intersection of the linear fit of the data

with E = 0 . Knowing E =0 allows us to determine the width of density of states from the relation:

66

= k 2 E =0

3 2

Equation 3.6

Finally, the implicit function g ( ) should be linear, because the derivative constant:

g is a

g g E C 2 2 k 2 = = = 2 = 2 C 2 E k

Equation 3.7

Therefore, calculating the slope =

g of the plot g ( ) allows for the determination of

the value of the off-diagonal disorder parameter :

2 =

2
k2

Equation 3.8

Table 3.1 shows the calculated values for the density of states at zero field E =0 ,
the disorder parameters and , and the constant C for holes and electrons in EHO-

OPPE. Figure 3.8 displays the temperature dependence of the diagonal disorder
parameter . It is apparent from the data presented in Table 3.1 and in Figure 3.8 that the

off-diagonal disorder parameter is larger than diagonal disorder parameter over the entire experimental temperature range. The temperature-independent off-diagonal

67

disorder parameter was 3.4 and 3.6 for electrons and holes, respectively, while the
range of the temperature-dependent diagonal disorder parameter was between 2.8 and

2.5 and 3.0 and 2.7 in the temperature range between 20 and 50 C for electrons and holes, respectively. Monte Carlo simulations of carrier hopping in the transport manifold

Charge carrier

E =0 (eV )

20 C
o

C (cm / V )1/ 2

Holes

0.072

3.6

3.0

5.410-4

Electrons

0.068

3.4

2.8

5.710-4

Table 3.1. Results of data fits to the Gaussian disorder model of charge transport.

with energy levels with a Gaussian distribution11,14 predicted that for large

, ln / E becomes negative at low electric fields ( 5105 Vcm-1) and does not
obey conventional Poole-Frenkel law ln (E)1/2. At higher fields ln / E becomes positive and gives rise to Poole-Frenkel-like behavior. While we were not able to measure the mobility at higher fields because of the dielectric breakdown of the material, the observed negative field dependence of the mobility in conjugated polymer EHOOPPE with high off-diagonal disorder is consistent with the disorder transport formalism applied to a system with both positional and energetic disorders.

68

3.4 3.2
e =3.4 h =3.6

3.0 2.8 2.6 290 300 310 320 330 Temperature (K) 340

Figure 3.8 Temperature dependence of the diagonal disorder parameter for holes

(squares) and electrons (circles). Parameters have been obtained by fitting the experimental data (Figure 3.7) to Equations 3.2 and 3.6.

3.4 Conclusions

In summary, we have shown that the -conjugated semiconducting polymer EHO-OPPE exhibits high ambipolar mobility (up to ~ 210-3 cm2V-1s-1). The positional (off-diagonal) and energetic (diagonal) disorder parameters have been calculated from the experimental temperature and field dependences of the hole and electron mobilities. The observed negative field dependence of the mobility was explained within the Gaussian disorder formalism to originate from high off-diagonal disorder. The experimental temperature and field dependent mobilities are consistent with a Gaussian disorder transport formalism applied to systems exhibiting both positional and energetic disorders.

69

3. 5 References

Weder, C. Ed. Poly(arylene ethynylene)s From Synthesis to Applications; Advances in Polymer Science Series Vol. 177; Springer, Heidelberg, 2005..

2 3

Weder, C.; Wrighton, M. S. Macromolecules 1996, 29, 5157. Shiyanovskaya, I.; Singer, K. D.; Twieg, R. L.; Sukhomlinova, L.; Gettwert, V.
Phys. Rev. E 2002, 65, 41715.

4 5

Scher, H.; Montroll, E. W. Phys. Rev. B 1975, 12, 2455. Lebedev, E.; Dittrich, T.; Petrova-Koch, V.; Karg, S.; Brtting, W. Appl. Phys. Lett.
1997, 71, 2686.

6 7

Hertel, D.; Bssler, H.; Scherf, U.; Hrhold, H. H. J. Chem. Phys. 1999, 110, 9214. Campbell, I. H.; Smith, D. L.; Neef, C. J.; Ferraris, J. P. Appl. Phys. Lett. 1999, 74, 2809.

Inigo, A. R.; Tan, C. H.; Fann, W.; Huang, Y. S.; Perng, G. Y.; Chen, S. A. Adv.
Mater. 2001, 13, 504.

9 10 11 12 13 14 15

Scott, J. C.; Pautmeier, L. T.; Schein, L. B. Phys. Rev. B 1992, 46, 8603. Babel, A. M.; Jenekhe, S. A. Adv. Mater. 2002, 14, 371. Borsenberger, P. M.; Peatier, L.; Bssler, H. J. Chem. Phys. 1991, 94, 5447. Hertel, D.; Bssler, H.; Scherf, U.; Hrhold, H. H. J. Chem. Phys. 1999, 110, 9214. Bssler, H. Phys. Stat. Sol. (b) 1993, 175, 15. Peatier, L.; Richert, R.; Bssler, H. Synth. Met. 1990, 37, 271. Movaghar, B.; Murray, D. W.; Donovan, K. J.; Wilson, E. G. J. Phys. C: Solid
State Phys. 1994, 17, 1247.

70

Chapter 4: Synthesis and Properties of Poly(2,5-dialkoxy-p-phenylene ethynylene)PtII Networks 4.1 Introduction As described in Chapter 3, poly[2,5-dioctyloxy-1,4-diethynyl-phenylene-alt-2,5bis(2'-ethylhexyloxy)-1,4-phenylene] (EHO-OPPE, 4) exhibits high ambipolar charge carrier mobilities (~210-3 cm2V-1s-1). Based on the hypothesis that the introduction of cross-links may improve the intermolecular charge transport and provide a commercially viable means for producing high charge carrier mobility materials, we embarked on the development and investigation of organometallic cross-linked conjugated polymers. The possibility to synthesize organometallic neworks based on PPE and a (presumably) nonconjugated dinuclear PtII complex by exploitation of ligand-exchange reactions was reported in earlier work by the Weder group (Chapter 1.4).1 These networks were synthesized employing poly[2,5-dioctyloxy-1,4-diethynyl-phenylene-alt-2-methoxy,5-2'ethylhexyloxy-1,4-phenylene] (MEH-OPPE, 1), a highly soluble PPE derivative,2 which displays essentially similar electro-optic characteristics as that of EHO-OPPE.3 PtII is well known to form stable bis-ethynyl complexes,4 its square-planar ligand sphere offers the advantage of a restricted number of isomers, as compared to octahedral coordinated metal centers. The dinuclear [Pt-(-Cl)Cl(PhCH=CH2)]2 (2) organometallic PtII complex, which possesses a particularly attractive bifuntional nature, was elected for the study. In order to investigate the influence of these presumably non-conjugated cross-links on the charge transport characteristics, we here describe the investigation of the charge transport in the EHO-OPPE-PtII networks (Scheme 4.1).

71

4.2 Experimental Section Materials and Methods All reagents and solvents were purchased from Aldrich Chemical Co. or FLUKA (analytical grade quality), and were used without further purification except styrene, which was destabilized with CaH2 and distilled prior to use. EHO-OPPE (4) of a numberaverage molecular weight of about 10,000 gmol-1 and [Pt-(-Cl)Cl(PhCH=CH2)]2 (2) were prepared as described before.

Synthesis and Processing of EHO-OPPE-PtII (5) Films The ITO coated glass slides were lined with scotch tape and preheated to 40 C in vacuo. A solution of [Pt-(-Cl)Cl(PhCH=CH2)]2 (2) in CHCl3 (3-5 mg in 0.2 mL) was added to a solution of EHO-OPPE (4) in CHCl3 (3-8 mg in 0.3 mL, [PtII]/[PE] ~0.17 and 0.35). The clear, yellow solution was rapidly shaken, drop cast on the preheated substrates with simultaneous application of vacuum and the films thus obtained were dried as described in Chapter 3.2. Their thickness, as determined with a Dektak profilometer was between 6 - 18 m.

Preparation of TOF samples and TOF measurements The TOF samples were prepared and the measurements conducted on a conventional setup, as mentioned in Chapter 3.2.

72

4.3 Charge Transport in EHO-OPPE-PtII Networks EHO-OPPE-PtII (5) networks were synthesized according to literature employing EHO-OPPE (4) (a highly soluble PPE derivative) and a bridged PtII organometallic complex [Pt-(-Cl)Cl(PhCH=CH2)]2 (2, Scheme 4.1). It should be pointed out that the previous study on organometallic cross-linked polymer networks conducted in the Weder group was based on MEH-OPPE. In order to allow an apples to apples comparison

+
n
O

Cl Pt

Cl
Pt

Cl

Cl

OR

OR

OR

x
RO
RO

y
RO

z n
Cl
Pt

Cl

Cl

Pt

Cl

OR

RO

Scheme 4.1 Simplified representation of the ligand-exchange reaction between EHOOPPE (4) and [Pt-(-Cl)Cl(PhCH=CH2)]2 (2), leading to cross-linked organometallic conjugated polymer network 5.

with the results presented in Chapter 3.3 we elected to employ EHO-OPPE for the work presented. Due to the various equilibria among the reacting species in the ligandexchange reaction previously reported, the 4/2 reaction mixture remains soluble

73

a
Photocurrent (10 A)
-7

0 0 2 -5 Time (10 s) 4

b
Photocurrent (10 A)
-7

10

0.1

0.01

0.1 1 -5 Time (10 s)

Figure 4.1 Hole TOF photocurrent transients of EHO-OPPE-PtII (5), [PtII]/[PE]=0.17 with film thickness L=9 m (black solid line) and L=13 m (red solid line), linear (a) and logarithmic (b) plots, measured at 295 K and F=9104 Vcm-1.

rendering the system processable. Once the solvent is evaporated, the processed films were insoluble in common organic solvents, but dissolved in styrene suggesting that indeed the organometallic cross-linked network 5 is obtained. 74

The carrier mobilities of 5 and the reference 4 were determined by employing the time-of-flight (TOF) technique as discussed in detail in Chapter 3.3.5 The TOF measurements were conducted with varying field strengths, PtII concentration and film a
Photocurrent (10 A)
-7

-9

-18

2 4 -5 Time (10 s)

b
Photocurrent (10 A) 10
-7

1 0.1 -5 Time (10 s) 1

Figure 4.2 Electron TOF photocurrent transients of EHO-OPPE-PtII (5), [PtII]/[PE]=0.17 with film thickness L=9 m (black solid line) and L=13 m (red solid line), linear (a) and logarithmic (b) plots, measured at 295 K and F=9104 Vcm-1.

75

a
Mobility (x10 cm /Vs)
2 -2

0.8

0.4

0.0 0

20 40 4 Field (x10 V/cm)

60

Mobility (x10 cm /Vs)

0.8

-2

0.4

0.0 0

20 40 4 Field (x10 V/cm)

60

Figure 4.3 Electron (top) and hole (bottom) mobility of EHO-OPPE-PtII (5) as function of [PtII]/[PE] ([PtII]/[PE]: filled symbols =0.35, open symbols =0.17), electric field F and thickness (=18.4 m, =10.1 m, =11.3 m, =9 m, =13 m, =6.5 m). Mobilities for EHO-OPPE () are displayed for reference.

76

thicknesses. The shape of the photocurrent transients of EHO-OPPE-PtII ([PtII]/[PE]= 0.17), shown in Figure 4.1 and Figure 4.2, are representative for all transients observed here and represent a dispersive mode of transport.6,7,8,9 However, as can be seen from Figure 4.1b and Figure 4.2b the transient time (ttr) did not scale according to the film thickness and hence the photocurrents observed were range limited. In the reference experiment conducted using 4, eventhough a dispersive mode of transport was observed, the transient time obtained indeed scaled with film thickness. The charge carrier mobilities calculated using the extracted transient times (which were obtained from the double logarithmic plot of the photocurrent transients) are plotted in Figure 4.3. The ambipolar characteristic of the parent polymer 4 were observed to be distorted upon complexation with PtII. The materials employed in this investigation, were synthesized and processed by similar techniques used in the TOF experiments conducted using 4. Thus the introduction of PtII as cross-links in the system, lead to range limited photocurrents as well as trapping of both holes and electrons.

4.4 Conclusions In summary we have demonstrated that organometallic conjugated polymer networks based on EHO-OPPE and PtII can be readily synthesized via ligand-exchange reactions. These systems are highly processable with the network formation occurring during the processing step. The nature of the metal center used as the cross-link is exceeding important for the resulting charge transport characteristics of the polymer network structures. The obtained photocurrent transients and the calculated carrier

77

mobility suggests that the introduction of PtII cross-links leads to range limited photocurrents and significant charge trapping of both electrons and holes.

78

4.5 References 1 2 Huber, C.; Bangerter, F.; Caseri, W.; Weder, C. J. Am. Chem. Soc. 2001, 123, 3857. Dellsperger, S.; Dtz, D.; Smith, P.; Weder, C. Macromol. Chem. Phys. 2000, 201, 192. 3 4 5 Weder, C.; Wrighton, M. S. Macromolecules 1996, 29, 5157. Albinati, A.; Caseri, W. R.; Pregosin, P. S. Organometallics 1987, 6, 788. Shiyanovskaya, I.; Singer, K. D.; Twieg, R. J.; Sukhomlinova, L.; Gettwert, V. Phys. Rev. E 2002, 65, 41715. 6 Lebedev, E.; Dittrich, T.; Petrova-Koch, V.; Karg, S.; Brtting, W. Appl. Phys. Lett. 1997, 71, 2686. 7 8 Hertel, D.; Bssler, H.; Scherf, U.; Hrhold, H. H. J. Chem. Phys. 1999, 110, 9214. Campbell, I. H.; Smith, D. L.; Neef, C. J.; Ferraris, J. P. Appl. Phys. Lett. 1999, 74, 2809. 9 Inigo, A. R.; Tan, C. H.; Fann, W.; Huang, Y.-S.; Perng, G.-Y.; Chen, S.-A. Adv. Mater. 2001, 13, 504.

79

Chapter 5: Synthesis and Properties of Poly(2,5-dialkoxy-p-phenylene ethynylene) Pt0 Networks*

5.1 Introduction As described in detail by the Weder group1 (see Chapters 1.2 and 4), the ethynyl moieties composed in the backbone of poly(p-phenylene ethynylene)s such as poly[2,5dioctyloxy-1,4-diethynyl-phenylene-alt-2,5-bis(2'-ethylhexyloxy)-1,4-phenylene] (EHOOPPE, 4,)2 form 2-alkyne complexes with a number of metals. Importantly it was also demonstrated that ligand-exchange reactions between the polymer and low-molecular metal complexes with weakly bound ligands represent an adequate framework for the simultaneous synthesis and processing of the targeted organometallic polymer networks. However, as evidenced by the experimental studies described in Chapter 4, the dinuclear PtII cross-links act as traps, and actually deteriorate, rather than improve the charge transport upon binding to EHO-OPPE. This Chapter focuses on the use of Pt0 as a crosslinking agent. Pt0 is well known to form stable bis(alkyne) complexes3 and appeared uniquely suited for the task at hand. In platinum(0)-alkyne complexes, usually two ligands coordinate to the platinum atoms. The resulting linear coordination geometry offers the advantage of a very limited number of isomers (when compared to squareplanar, tetrahedrically, or octahedrically coordinated metal centers). A -conjugation across the platinum centers is considered to be possible via interaction of the bond of the ligands with the dx2-y2 orbital of the Pt atoms or via -backbonding from the dxz

Parts of this Chapter have been published: (a) Huber, C.; Kokil, A.; Caseri, W.; Weder, C. Organometallics 2002, 21, 3817. (b) Kokil, A.; Shiyanovskaya, I.; Singer, K.D.; Weder, C. J. Am. Chem. Soc. 2002, 124, 9978. (c) Kokil, A.; Huber, C.; Caseri, W.; Weder, C. Macromol. Chem. Phys. 2003, 204, 40.

80

orbital of Pt to * orbitals of the ligands.4 Further, about 1/3 of the platinum atoms possess a nuclear spin of 1/2 (195Pt). Therefore, ligand-exchange reactions can in many cases be followed in-situ by nuclear magnetic resonance (NMR) spectroscopy, which probes the metal centers themselves. Concomitantly, we have elected to employ tris(styrene)platinum(0), Pt(PhCH=CH2)3 (7),5 as the source of Pt0 (Scheme 5.2). Stable organometallic complexes almost always contain 18 valence electrons - this statement is known as the 18 - Electron Rule or the Effective Atomic Number Rule.6 This statement stems from the fact that transition metals have nine valence orbitals: five nd orbitals, three (n+1)p orbitals and one (n+1)s orbital. The maximum capacity of the bonding and the non-bonding orbitals in the complex will always be the same, i.e. 18 electrons, because each of the nine metal valence orbitals must give rise to a bonding or a nonbonding molecular orbital. It is well known that coordinated alkenes act as a two-electron donor and also serve as a acceptor due to the low lying anti-bonding orbital. Thus, in the case of complexes such as 7, the organometallic complex has 16 valence electrons and has a trigonal-planar structure which maximizes the metal-ligand back-bonding. Therefore these organometallic complexes can then undergo ligand-exchage reactions to form stable 18 electron species. Particularly attractive is the fact that the liberated ligand (styrene) is volatile and can be readily removed from the product. In order to explore the behavior of 7 in ligand-exchange reactions with phenylene ethynylenes, we first conducted a number of model reactions with diphenylacetylene, Ph-CC-Ph, which are summarized in Chapter 5.3. Since zerovalent platinum complexes are important for a wide variety of homogeneous catalytic processes,7 these model reactions may be important in their own

81

right. Bis-(2-alkyne)platinum(0) complexes represent a particularly interesting subgroup that is viewed as a family of (potentially) useful synthetic intermediates, since the alkyne ligands are readily replaced by other donor ligands such as PMe3, PEt3, PPh3, and tBuCN. These alkyne complexes are frequently synthesized by ligand-exchange reactions from bis(1,5-cyclooctadiene)platinum(0)3a,b or tris(2-ethylene)platinum(0).3c However, the synthesis of the latter compounds is laborious and requires rigorous exclusion of oxygen and water.3a,b The preparation of bis(2-alkyne)platinum(0) complexes via decomposition of bis(4-penten-1-yl)platinum(II) in the presence of alkynes has also been reported,3c but the related procedures require not only a protective atmosphere but also high pressure, which renders the preparation intricate. Another method starts from the hydrosilylation catalyst hexachloroplatinic(IV) acid, which is heated in the presence of an excess of sym-tetramethyldivinylsiloxane and exposed in-situ to alkynes.8 Surprisingly, H2[PtCl6].xH2O is reduced without explicit addition of a reducing agent. Thus, the reduction appears to be induced by impurities, subjecting this reaction to some imponderables. Indeed, it is well established that the reduction of hexachloroplatinic(IV) acid in hydrosilylation catalysis is accompanied by an induction period of variable length.9 In Chapter 5.3, we also describe two simple, well-defined and reproducible methods for the synthesis of Pt(2-Ph-CC-Ph)23b,8 (8), which can be performed under ambient conditions. One of these methods - the ligand-exchange between Pt(PhCH=CH2)3 with phenylene ethynylenes was subsequently employed for the synthesis of EHO-OPPE-Pt0 networks, as described in detail in Chapter 5.4. In Chapter 5.5, finally, an in-depth characterization of the charge transport properties of these organometallic network structures is presented.

82

5.2 Experimental Section Materials and Methods All reagents and solvents were purchased from Aldrich Chemical Co. or FLUKA (analytical grade quality), and were used without further purification except styrene, which was destabilized with CaH2 and distilled prior to use. EHO-OPPE (4) of a numberaverage molecular weight of about 10,000 gmol-1 and Pt(PhCH=CH2)3 (7) were prepared as described before; ion-coupled plasma-optical emission spectroscopy (ICP-OES), conducted by Solvias AG, Basel, revealed a Pt content of ca. 0.4 % w/w (1.910-2 M) for the solution of 7 in styrene. A more dilute styrene solution of 7 (~610-3 M) was also prepared by the same method.
195

Pt NMR spectral data were obtained in toluene-d8 on a

Bruker 300 MHz NMR spectrometer. Chemical shifts () are expressed in ppm relative to tetramethyl silane (1H) or sodium hexachloroplatinate(IV) (195Pt), respectively. IR

spectra of films were recorded on KBr substrates on a Bruker IFS 66v. UV-Vis absorption spectra were obtained on a Perkin Elmer Lambda 800. PL spectra were measured under excitation at 380 nm on a SPEX Fluorolog 3 (Model FL3-12). Corrections for the spectral dispersion of the Xe-lamp, the instrument throughput, and the detector response were applied.

Synthesis of Pt(2-Ph-CC-Ph)2 (8) by ligand-exchange reaction Ph-CC-Ph (3.00 g, 16.9 mmol) was added to a solution of 7 (0.37 mmol) in styrene (20 mL), and styrene was distilled off (10 mbar, 50 C) to reduce the volume of the mixture to 5 mL. Pentane (15 mL) was added, and the mixture was cooled to 0 C. The resulting solid was filtered off, washed with pentane, dried in vacuo, and

83

reprecipitated from CH2Cl2/pentane (5/30 mL) at -78 C to afford a pale yellow solid (117 mg, 57 %).
195

Pt NMR (64 MHz, CD2Cl2, 263 K): -3553 ppm. Anal. Calcd for

C28H20Pt: C, 60.98; H, 3.65. Found: C, 60.81; H, 4.06.

Synthesis of Pt(2-Ph-CC-Ph)2 (8) via reduction of PtCl2(PhCH=CH2)2 (6) by triphenylsilane in the presence of Ph-CC-Ph Triphenylsilane (439 mg, 1.68 mmol) was added to a mixture of 6 (400 mg, 0.844 mmol) and Ph-CC-Ph (1.50 g, 8.46 mmol) in CH2Cl2 (16 mL), and the reaction mixture was stirred for 30 min. Aqueous ammonia (37 mL, 3%) was added, and the mixture was vigorously stirred for 90 min. The organic phase was separated off, extracted with water (33 mL), and filtered through a 3 cm plug of Alox (basic, activity I), which was subsequently rinsed with CH2Cl2 (16 mL), and the organic layers were combined. CH2Cl2 was distilled off to reduce the volume of the mixture to 5 mL, pentane (20 mL) was added, and the mixture was cooled to 0 C. The resulting solid was filtered off, washed with pentane, and dried in vacuo to afford a pale yellow solid (121 mg, 26 %). 195Pt NMR (64 MHz, CD2Cl2, 263 K): -3553 ppm. Anal. Calcd for C28H20Pt: C, 60.98; H, 3.65. Found: C, 61.14; H, 4.13.

Investigation of the in-situ ligand-exchange reaction between Pt(PhCH=CH2)3 (7) and Ph-CC-Ph In an NMR tube, Ph-CC-Ph (5 mg, 0.028 mmol) was added to a mixture of the styrene solution of 7 (1 mL, 1.910-2 M, 0.019 mmol) and toluene d8 (0.1 mL). After dissolution of the Ph-CC-Ph, the solution was cooled to 263 K, and a
195

Pt NMR

spectrum was acquired (64 MHz, 263 K): -3547.5, -5874.8, -5891.6. The experiment 84

was repeated, but 11.6 mg (0.065 mmol) of Ph-CC-Ph were added:

195

Pt NMR -

3548.0. The experiment was repeated, but 59.6 mg (0.335 mmol) of Ph-CC-Ph were added:
195

Pt NMR -3548.1.

Investigation of the in situ ligand-exchange reaction between EHO-OPPE (4) and Pt(PhCH=CH2)3 (7) In an NMR tube, a solution of 4 (1 mg, 0.0028 mmol of ethynylene moieties) in distilled styrene (0.5 mL) was added to a mixture of a styrene solution of 7 (1 mL, 1.9102

M, 0.019 mmol) and toluene d8 (0.1 mL). The mixture gelled ca. 30 seconds after
195

mixing.

Pt NMR (64 MHz, 263K): -5876.9, -5891.5. The experiment was repeated,
195

but 7.5 mg (0.021 mmol) of EHO-OPPE were used: Synthesis of EHO-OPPE-Pt0 (9) gels

Pt NMR -5877.1, -5891.8.

A dilute solution of 7 in styrene (1 mL, 1.910-2 M, 0.019 mmol), was added to a solution of 4 (7.5 - 10 mg, 0.021 - 0.028 mmol of ethynylene moieties) in distilled styrene (0.5 mL), causing an immediate gelation.

Synthesis and processing of EHO-OPPE-Pt0 (9) films A dilute (~610-3 M) solution of 7 in styrene (0.1-0.47 mL) was added to a solution of 4 in styrene (3-12 mg in 0.3 mL, [Pt0]/[PE] ~0.016-0.34). The clear, yellow solution was rapidly shaken and immediately (typically within 5 seconds after mixing, and, of course, before the onset of gelation) processed into thin films by spin coating (optical absorption and emission spectra) or solution-casting (TOF measurements). The films were dried in vacuum at 40 C for at least 12 h. Their thickness, as determined with

85

a Dektak profilometer, was between 0.5 - 1 m (optical absorption and emission spectra) and 16 - 30 m (TOF measurements), respectively. The organometallic polymer network 9 was processed as mentioned in chapter 4.2 to obtain polymer films for TOF measurements. Raman spectrum ([Pt0] / [PE] ~0.17): 3057, 3005, 2905, 2854, 2199, 1842, 1595, 1547, 1448, 1290, 1198, 1157, 1059, 1033, 1000, 793, 670, 620, 398, 168 cm-1.

Preparation of TOF samples and TOF measurements Samples for the TOF technique were prepared and the measurements were conducted on a conventional setup, as outlined in Chapter 4.2.

5.3 Two Alternative, Convenient Routes to Bis(diphenylacetylene)Pt0 The first procedure investigated here for the preparation of the model compound Pt(2-Ph-CC-Ph)2 (8) is based on a ligand-exchange reaction between Ph-CC-Ph and Pt(PhCH=CH2)3 (7). The latter complex is readily obtained by reduction of cisPtCl2(PhCH=CH2)210 (6) with triphenylsilane in a styrene solution. The styrene solution of 7 is stable for months even in the absence of a protective gas. Extensive NMR investigations revealed that 7 readily converts in-situ into a variety of Pt0 derivatives by ligand-exchange, in particular with 1,2-bis(diphenylphosphino)ethane (diphos), P(OEt)3, and 1,5-cyclooctadiene (cod). Regrettably, however, neither 8 nor derivatives prepared via the latter could be isolated in the past from the styrene solution. We now have found that treating a styrene solution of 7 with Ph-CC-Ph leads to the formation of 8 (Scheme 5.1). In situ 195Pt NMR experiments revealed that the signals related to 7 at -5874.8 and -

86

5893.2 ppm are fully suppressed upon addition of 2 equiv of Ph-CC-Ph per Pt atom, and the only detectable resonance is that of 8 at -3548 ppm (Figure 5.1). Thus, even in the

Pt(II)Cl2

styrene

Cl (II) Cl Pt

triphenylsilane diphenylacetylene 26 %

triphenylsilane styrene

Pt

(0)

diphenylacetylene 57 %

Pt

(0)

Scheme 5.1 Synthesis of Pt(2-Ph-CC-Ph)2 (9) through (i) ligand exchange exchange between Pt(PhCH=CH2)3 (7) and Ph-CC-Ph, and (ii) reduction of cisPtCl2(PhCH=CH2)2 (6) with triphenylsilane in the presence of Ph-CC-Ph.

presence of a large excess of styrene, the styrene ligands coordinated to Pt0 are very efficiently replaced by Ph-CC-Ph. In contrast to the complexes mentioned above that were exclusively prepared in situ and could not be isolated, we were able to isolate 8 by precipitation upon addition of pentane. We subsequently explored the possibility of preparing 8 directly through the reduction of 6 with triphenylsilane in the presence of PhCC-Ph (Scheme 5.1) and under ambient conditions. Gratifyingly, 8

could indeed be isolated with this method in high purity. Preliminary experiments have

87

Figure 5.1

195

Pt NMR spectra for Pt(PhCH=CH2)3 (7, a), 7 with 1.47 eq. of Ph-CC-Ph

(b) and 7 with 3.42 eq. of Ph-CC-Ph (c).

shown that also Pt(2-HOMe2C-CC-CMe2OH)2 can be synthesized by this framework, while attempts to prepare bis(1,5-cyclooctadiene)platinum(0) using the same procedure have hitherto been unsuccessful.

88

5.4 Synthesis of Organometallic -Conjugated PPE - Pt0 Networks The conjugated polymer employed here was EHO-OPPE (4), a common poly(pphenylene ethynylene) (PPE) derivative, which comprises ethynylene groups that may act as ligand sites (Scheme 5.2). Pt0 was chosen as cross-linker, since it forms stable bis(ethynylene) complexes, which due to -back-bonding allow for electronic conjugation. A styrene solution of Pt(PhCH=CH2)3 (7) served as the Pt0 source. Since styrene is volatile, the ligands can be removed from the product after exchange with the PPE.

O + Pt
(0)

OR

OR

OR

x
RO RO

y
RO

z n Pt
(0)

OR

RO

Scheme 5.2 Ligand-exchange reaction between EHO-OPPE (4) and Pt(PhCH=CH2)3 (7) leading to the target EHO-OPPE-Pt0 (9) networks.

89

As shown in Scheme 5.1 and mentioned in Chapter 5.2, the organometallic complex 7 was synthesized according to a previously published procedure. Importantly, styrene is a good solvent for 4; thus, the ligand-exchange reactions can be conducted in the presence of styrene as a (co-)solvent, which allows one to employ the concentration of styrene as a variable to direct the reaction equilibrium during processing. In analogy to the model reactions (Chapter 5.2), we have prepared EHO-OPPE-Pt0 (9) networks through the reaction of 4 and 7 by combining styrene solutions of these reactants (Scheme 5.2). The ratio of the molar concentrations of Pt0 and phenylene ethynylene moieties, [Pt0] / [PE], was varied over a broad range (vide infra). Gratifyingly, even when [Pt0] / [PE] was as low as ~0.016, the reaction mixtures formed gels within ca. 10 to 30 seconds, consistent with the formation of cross-linked network structures (Scheme 1, z > 0, Figure 5.3). As for the low-molecular model reaction, we studied the ligandexchange between 7 and 4 by means of in-situ 195Pt NMR experiments. The acquisition of proper solution NMR spectra, however, was stifled by the rapid gelation of the reaction mixtures, which apparently increased the relaxation time of the
195

Pt spins beyond the

detection limit. The spectra of these gels (recorded under solution conditions) displayed signals at ca. -5877 and -5892 ppm, which coincide with those of the neat 7 and indicate that at least some fraction of the latter remains unreacted and dissolved under the conditions applied. The NMR spectra were void of other signals, including the expected resonances of ethynylene-Pt0 complexes around -3548 ppm. This finding is consistent with the immobilization of a vast majority of the PPE molecules in the network structure. Figure 5.2 shows that the emission spectrum of a gel of 9 exhibits essentially similar features as a styrene solution of the neat PPE, but the intensity is substantially

90

reduced. This behavior is consistent with the formation of a conjugated network in which exciton migration to low-energy defect sites is substantially enhanced (vide infra). As is obvious from Figure 5.3, gels of the organometallic network 9 can exhibit an appreciable

60 50 PL intensity (cps) 40 30 20 10 0 450 525 600 Wavelength (nm) 675 750

Figure 5.2 Photoluminescence spectra of a solution of EHO-OPPE (4) in styrene (solid line, 5 mg / mL) and an EHO-OPPE-Pt0 (9) gel (broken line, [Pt0] / [PE] = 0.68, 5 mg PPE / mL styrene). The spectra were acquired under excitation at 380 nm.

mechanical strength, even when the solvent-content is as high as 99 % w/w. Even at a very high Pt-content, the gels remain photoluminescent (Figures 5.2 and 5.3). Importantly, the reaction mixtures of 9 could be spin-cast and solution-cast into homogeneous films of good optical quality, if processed before the onset of gelation. The films thus produced were insoluble in common solvents, consistent with the formation of cross-linked EHO-OPPE-Pt0 networks and the elimination of not coordinated styrene from the films. On the other hand, styrene was found to readily re-dissolve the EHOOPPE-Pt0 network structures, demonstrating that the ligand-exchange reaction is a

91

Figure 5.3 Left: Styrene solution of EHO-OPPE (4, 6.7 mg/mL). Right: EHO-OPPE-Pt0 (9) / styrene gel ([Pt0] / [PE] = 0.68, 6.7 mg PPE / mL styrene), carrying a steel ball. The picture on the left was taken under illumination with UV light (365 nm).

reversible process (Figure 5.4). While other spectroscopic techniques were impeded by the insolubility of the materials, Raman spectra of a selected sample of 9 ([Pt0] / [PE] ~ 0.17), in contrast to the undecorated 4, showed a distinct, broad signal at 1842 cm-1, 92

Figure 5.4 Pictures of an EHO-OPPE-Pt0 (9) film in styrene (left) and toluene (right) under illumination with UV-light (365 nm). The dissimilar dissolution characteristics of the photoluminescent polymer demonstrate the reversibility of the ligand-exchange reaction and the cross-linked nature of the material.

which was also observed for the 8 model compound (split band at 1882 and 1864 cm-1) and is indicative of the formation of the Pt0-ethynylene complex (Figures 5.5). In the case of Ph-CC-Ph the CC stretch observed at 2221 cm-1 was completely suppressed upon complexation. Upon complexation of EHO-OPPE with Pt0, the CC stretch was observed at 2198 cm-1 however with a reduced intensity which corresponded to the chosen Pt0

93

Figure 5.5 Raman spectra of Ph-CC-Ph (a), EHO-OPPE (4, b), Pt(2-Ph-CC-Ph)2 (8, c) EHO-OPPE-Pt0 (9, d, [Pt0] / [PE] ~ 0.17).

94

concentration ([Pt0] / [PE] ~ 0.17). In view of the sometimes limited stability of Pt0 complexes, it is an important notion that films of the organometallic network 9 could be prepared and handled under ambient conditions and appeared to be stable for weeks. The absorption spectra of films of 9 with [Pt0] / [PE] ~ 0.016 - 0.34 and neat 4 as a reference are shown in Figure 5.6 (note that the spectra have been scaled to optimally fit the graphs). The complexation of 4 with Pt0 leads to a slight hypsochromic shift of the

0.28

Absorbance (a.u.)

0.21

0.14

0.07

0.00 300

400

500 600 Wavelength (nm)

700

800

Figure 5.6 UV-Vis absorption spectra of films of EHO-OPPE (4) (solid line) and EHOOPPE-Pt0 (9) with [Pt0] / [PE] = 0.086 (filled circles), 0.17 (open circles), 0.25 (filled triangles), and 0.34 (open triangles).

absorption band and a reduction of the extinction coefficient. These effects become more pronounced with increasing Pt-content, at least for [Pt0] / [PE] ~ 0.016 - 0.17, suggesting that the conjugation in the PPE backbone is affected by the complexation. The crystal structure of the 8 model compound3a reveals that the phenyl rings bend slightly away from the platinum (angle between phenyl and C=C group 153, Figure 5.7).3a Thus, these

95

Figure 5.7 Crystal structure for Pt(2-Ph-CC-Ph)2 (8).3a

1.00 PL Intensity (a.u.) 0.75 0.50 0.25 0.00 450

500

550

600

650

700

Wavelength (nm)

Figure 5.8 Photoluminescence spectra of films of EHO-OPPE (4) (solid line) and EHOOPPE-Pt0 (9) with [Pt0] / [PE] = 0.086 (filled circles), 0.17 (open circles), 0.25 (filled triangles), and 0.34 (open triangles).

96

data indicate that the complexation may affect the effective conjugation length of the PPE. Concomitantly, a weak, broad absorption band centered around 625 - 650 nm is observed for films with [Pt0] / [PE] ~0.17-0.34, which may be indicative of the formation of colloidal Pt. The complexation of 4 with Pt0 also leads to a substantial reduction of the photoluminescence intensity (Figure 5.8). The shape of the emission spectrum remains essentially unchanged, and the films investigated display essentially identical emission characteristics. This behavior is again fully consistent with efficient exciton migration to the complexation sites, which either provide pathways for non-emissive relaxation processes or allow for an improved energy transfer to low-band-gap quenching sites.

5.5 Charge Carrier Mobility in Organometallic -Conjugated PPE Pt0 Networks The carrier mobilities of EHO-OPPE-Pt0 (9) and an EHO-OPPE (4) reference were determined by time-of-flight (TOF) measurements on indium-tin

oxide/polymer/gold samples as described in more detail in Chapter 3.3.11 The shape of the photocurrent transients of 4 and 9 ([Pt0]/[PE] ~ 0.17), shown in Figure 5.9, is representative for all transients observed here and is characteristic of dispersive transport.12 This mechanism is typical for materials with a high degree of spatial and/or energetic disorder and is concomitant with a wide variation of local transport rates.13 TOF measurements were performed as functions of carrier type (by changing the bias), electric field, and Pt0 content. It was confirmed for 4 and 9 with [Pt0]/[PE] ~ 0.25 that the transient time scales with L, which demonstrates that the photocurrents are not rangelimited. Further, no photocurrent was observed in a reference film (L = 2 m) of PMMA comprising 50 % w/w 9, confirming that the current was not caused by a Pt0-ethynylene

97

a
Photocurrent (10 A)
-7

0 0 5 -5 Time (10 s) 10

b
Photocurrent (10 A)
-7

10

0.1

1 -5 Time (10 s)

10

Figure 5.9 Electron TOF photocurrent transients of EHO-OPPE (4, solid, L=8 m) and EHO-OPPE-Pt0 (9, dotted, L=30 m, [Pt0]/[PE]=0.17) films in linear (a) and logarithmic (b) plots, measured at 295 K and E = 1.5105 Vcm-1.

complex per se. High electron (1.910-3 cm2V-1s-1) and hole (1.610-3 cm2V-2s-1) mobilities were found at low E (3.8104 Vcm-1) for 4. These data compare favorably with the highest values yet observed for an ambipolar conjugated polymer.14 The mobility depends on E

98

a
Mobility (10 cm V s )
-1 -2 2 -1

1.6

1.2

0.8

0.4

0.0 0.0

0.5

1.0 1.5 5 -1 Field (10 V cm )

2.0

2.5

b
Mobility (10 cm V s )
-1 2 -2 -1

1.6

1.2

0.8

0.4

0.0 0.0

0.5

1.0 1.5 5 -1 Field (10 V cm )

2.0

2.5

Figure 5.10 Electron (a) and hole (b) mobility of EHO-OPPE-Pt0 (9) as function of [Pt0]/[PE] and electric field E ([Pt0]/[PE]: =0, =0.016, =0.086, =0.17, =0.25, =0.34).

and decreases with increasing bias (Figure 5.10). This behavior was discussed in detail in Chapter 3.3 and is explained by a hopping transport model that accounts for off-diagonal (positional) disorder caused by variations of intersite distances in addition to diagonal (energetic) disorder in the transport manifold.12b,15 The large off-diagonal disorder results

99

a
Mobility (10 cm V s )
-1 -3 2 -1

0.0

0.1

0.2 0 [Pt ] / [PE]

0.3

0.4

b
Mobility (10 cm V s )
-1 -3 2 -1

0.0

0.1

0.2 0 [Pt ] / [PE]

0.3

0.4

Figure 5.11 Electron (a) and hole (b) mobility at E = 1.1105 () and 1.5105 Vcm-1 () as function of [Pt0]/[PE].

in a negative field dependence of the mobility at low fields, since higher fields favor forward hopping and inhibit faster routes for carriers involving hops transverse to the electric field. The negative field dependence was also predicted for quasi onedimensional transport in the presence of defects and barriers.16 The data shown in Figure

100

a
Mobility (10 cm /Vs)
2

1.6 1.2 0.8 0.4 0.0 0.0

-2

0.5

1.0 1.5 2.0 5 Field (10 V/cm)

2.5

b
Mobility (10-2 cm2/Vs)

1.6 1.2 0.8 0.4 0.0 0.0

0.5

1.0

1.5
5

2.0

2.5

Field (10 V/cm)

Figure 5.12 Electron (a) and hole (b) mobility of films of EHO-OPPE-Pt0 (9) with [Pt0] / [PE] = 0.25 having thickness 25m (filled squares) and 20m (open squares).

5.10 and 5.11 also show that the carrier mobility strongly increases upon introduction of Pt0. TOF measurements conducted on films of 9 [Pt0]/[PE] ~ 0.25 displayed that the charge carrier mobility does not scale with the film thickness as can be seen from the data shown in Figure 5.12. While a distinct enhancement of the mobility is observed for 9

101

with small [Pt0]/[PE], the effect levels off at a [Pt0]/[PE] of 0.17, where mobilities of 1.610-2 cm2V-1s-1 (electrons) and 1.410-2 cm2V-1s-1 (holes) are reached (at E=3.8104 Vcm-1), cf. Figure 5.10. These values are an order of magnitude higher than those of the neat PPE. Higher Pt0 contents did not further increase . Thus, as indicated by the optical data, the cross-link density may have reached a saturation limit (Scheme 5.2), and no further cross-links but colloidal Pt0 is formed. The absorption spectra of samples containing high concentration of Pt0 ([Pt0]/[PE] ~ 0.086-0.34) display a broad band centered around 642 nm and confirm that colloidal Pt0 is indeed formed when higher concentrations of 7 are employed. Further, an excessive complexation may limit the effective conjugation length of the PPE. The networks 9 display the same field dependence as the parent PPE, which suggests that introduction of Pt0 preserves the large off-diagonal disorder in the network.

5.5 Conclusions Thus in conclusion, we have demonstrated two new and convenient routes for the synthesis of bis(2-alkyne)platinum(0). Both methods depart from readily accessible, chemically stable intermediates, are based on straight forward and undemanding reaction conditions, and afford the desired product in high purity. We have also demonstrated that organometallic conjugated polymer networks can be synthesized and processed by ligand-exchange between a linear conjugated polymer and a labile metal complex. The introduction of conjugated cross-links between the conjugated macromolecules leads to a substantial increase of the charge carrier mobility. The charge carrier mobilities of the present EHO-OPPE-Pt0 networks (1.610-2 cm2V-1s-1 for electrons and 1.410-2 cm2V-1s-1

102

for holes) represent the highest mobilities yet observed in disordered conjugated polymers and also compare well to the hole mobilities of ordered materials (10-1 cm2 V-1 s-1).17 The ease of processing and the ambipolar characteristics of the new materials are particularly intriguing and may lead to a new generation of higher-performance semiconducting devices.

103

5.6 References 1 2 3 Huber, C.; Bangerter, F.; Caseri, W.; Weder, C. J. Am. Chem. Soc. 2001, 123, 3857. Weder, C.; Wrighton, M. S. Macromolecules 1996, 29, 5157. (a) Green, M.; Grove, D. M.; Howard, J. A. K.; Spencer, J. L.; Stone, F. G. A. J. Chem. Soc, Chem. Comm. 1976, 759. (b) Boag, N. M.; Green, M.; Grove, D. M.; Howard, J. A. K.; Spencer, J. L.; Stone, F. G. A. J. Chem. Soc., Dalton Trans. 1980, 2170. (c) Young, B. in Comprehensive Organometallic Chemistry II, Abel, E. W.; Stone, F. G. A.; Wilkinson, G. Eds., Pergamon Press, Oxford, 1995, p. VIIII / 533. 4 (a) Malatesta, L.; Cenini, S Eds., Zerovalent Compounds of Metals, Academic Press, London, 1974. (b) Collman, J. P.; Hegedus, L. S. Eds. Principles and Applications of Organotransition Metal Chemistry, University Science Books, Mill Valley (California), 1980. 5 6 Caseri, W.; Pregosin, P. S. Organometallics 1988, 7, 1373. Collman, J. P.; Hegedus, L. S.; Norton, J. R.; Finke, R. G. Principles and Applications of Organotransition Metal Chemistry; University Science Books, Sausalito California, 1987. 7 Hartley, F. R. in Comprehensive Organometallic Chemistry, Wilkinson, G., Ed., Pergamon Press, Oxford, 1982; Vol. 6, p. 471. 8 Chandra, G.; Lo, P. Y.; Hitchcock, P. B.; Lappert, M. F. Organometallics 1987, 6, 191. 9 10 Speier, J. L. Adv. Organomet. Chem. 1979, 17, 407. Albinati, A.; Caseri, W. R.; Pregosin, P. S. Organometallics 1987, 6, 788.

104

11

Shiyanovskaya, I.; Singer, K. D.; Twieg, R. J.; Sukhomlinova, L.; Gettwert, V. Phys. Rev. E 2002, 65, 41715.

12

(a) Lebedev, E.; Dittrich, T.; Petrova-Koch, V.; Karg, S.; Brtting, W. Appl. Phys. Lett. 1997, 71, 2686. (b) Hertel, D.; Bssler, H.; Scherf, U.; Hrhold, H. H. J. Chem. Phys. 1999, 110, 9214. (c) Campbell, I. H.; Smith, D. L.; Neef, C. J.; Ferraris, J. P. Appl. Phys. Lett. 1999, 74, 2809. (d) Inigo, A. R.; Tan, C. H.; Fann, W.; Huang, Y. S.; Peng, G. Y.; Chen, S. A. Adv. Mater. 2001, 13, 504.

13 14 15 16

Scott, J. C.; Pautmeier, L. T.; Schein, L. B. Phys. Rev. B 1992, 46, 8603. Babel, A. M.; Jenekhe, S. A. Adv. Mater. 2002, 14, 371. Pautmeier, L.; Richert, R.; Bssler, H. Synth. Met. 1990, 37, 271. Movaghar, B.; Murray, D. W.; Donovan, K. J.; Wilson, E. G. J. Phys. C: Solid State Phys. 1984, 17, 1247.

17

Sirringhaus, H.; Brown, P. J.; Friend, R. H.; Nielsen, M. M.; Bechgaard, K.; Langeveld-Voss, B. M. W.; Spiering, A. J. H.; Janssen, R. A. J.; Meijer, E. W.; Herwig, P.; de Leeuw, D. M. Nature 1999, 401, 685.

105

Chapter 6: Synthesis and Properties of Organometallic Networks Based on 2,2Bipyridine-Containing Poly(p-phenylene ethynylene)s* 6.1 Introduction As discussed in Chapter 5, we have demonstrated that organometallic conjugated polymer networks may exhibit substantially better charge transport characteristics than the linear parent polymers and can, at least in part, overcome the problems associated with inter-chain charge transfer between individual macromolecules.1 As is evident from the studies summarized in Chapter 4 and Chapter 5, our previous work in this arena was based on the exploitation of the ethynyl moieties comprised in the PPE backbone as ligand sites, which can form 2-alkyne complexes with a number of metals. While bis(2-diphenyl ethynylene)Pt0 complexes2 have been tremendously useful for our initial work, other examples of suitable bis-(2-diphenyl ethynylene)metal complexes, which could provide electronic conjugation between chains, are rare.3 In order to broaden the scope of our work to other metal complexes that could be used as cross-links, we considered the introduction of auxiliary ligands into the polymer backbone and opted to incorporate 2,2-bipyridine (Bipy) moieties. This ligand has extensively been used as a chelating ligand for a broad variety of metals and was recently heralded as the most widely used ligand.4 Heterocyclic chelating ligands such as bipyridines tend to bind metals through the donation of the lone pair of electrons present on the nitrogen atoms rather than the C-N bond. Such chelated complexes are entropically more favoured as compared to complexes with non-chelating ligands.5 Indeed, the Bipy moiety has already

Parts of this Chapter have been published: Kokil, A.; Yao, P.; Weder, C. Macromolecules 2005, 38, 3800.

106

been introduced into a plethora of macromolecules that form the basis of metallosupramolecular systems.6 Pioneering work on PPEs with Bipy groups in the polymer backbone and linear metal complexes of these polymers has been carried out by the groups of Schanze7 and Klemm.8 Interestingly, the metal-complexed PPEs investigated in these studies were virtually exclusively prepared by polymerizing metal-complexed monomers, rather than by complexation of the Bipy-containing polymer with metals.9 We show here that the latter framework, which mainly with sensor applications in mind and not under consideration of potential network formation has been applied by a number of groups for a variety of other conjugated polymer platforms,10,11,12,13,14,15 is formidably suited to prepare metallo-supramolecular PPE networks.

6.2 Experimental Section Materials and Methods 1,4-Bis[(2-ethylhexyl)oxy]-2,5-diiodobenzene (10),16 1,4-diethynyl-2,5-bis-

(octyloxy) benzene (11) and 5,5-dibromo-2,2-bipyridine17 were prepared as described before. All other solvents, reagents, metal complexes and catalysts were purchased from Aldrich Chemical Co., Fisher Scientific, Strem Chemicals or FLUKA (analytical grade quality), and were used without further purification. Unless otherwise stated, spectroscopic grade CHCl3 (Aldrich) was employed for all optical absorption and emission experiments and was passed through a plug of neutral Alox before use. Anhydrous toluene and anhydrous diisopropyl amine used for the polymerization reactions were purchased from Aldrich Chemical Co. and were deoxygenated by sparging with Ar for at least 1 h prior to use. All cross-coupling reactions were carried out under a

107

dry Ar atmosphere using standard vacuum-line and glove box techniques. Unless otherwise stated, 1H NMR spectra were recorded in CDCl3 at 22 C on a Varian XL (300 MHz) or a Varian Inova (600 MHz) spectrometer and chemical shifts are expressed in ppm relative to an internal TMS standard. Elemental analyses were obtained from Prevalere Life Sciences Inc. and Galbraith Laboratories Inc. UV-Vis absorption spectra were recorded on a Perkin Elmer Lambda 800 spectrometer. Photoluminescence (PL) spectra were measured under excitation at 410 nm on a SPEX Fluorolog 3 (Model FL312) spectrometer; corrections for the spectral dispersion of the Xe-lamp, the instrument throughput and the detector response were applied.

Synthesis of 5,5-Bis((trimethylsilyl)-ethynyl)-2,2-bipyridine18,19 (Scheme 6.1) 5,5-Dibromo-2,2-bipyridine (1.00 g, 3.18 mmol), Pd(P(Ph3)3)2Cl2 (178 mg, 0.25 mmol) and CuI (48 mg, 0.25 mmol) were dissolved in a mixture of triethylamine (20 mL) and DMF (16 mL) at 55 C, and trimethylsilyl acetylene (625 mg, 6.36 mmol) was added over the period of 1 h under vigorous stirring. After the addition of trimethylsilyl acetylene was completed, the reaction mixture was stirred at reflux for 1 h. The resulting suspension was cooled to RT, CHCl3 (100 mL) was added and the now clear solution was washed with a saturated aqueous EDTA solution (100 mL) for 1 h. The organic phase was separated off and the aqueous layer, which had turned turquoise, was re-extracted with CHCl3 (100 mL). The combined organic layers were washed with deionized water (200 mL), dried over MgSO4, and the solvent was evaporated in vacuum. The solid residue was purified by column chromatography (silica gel; hexanes/ethyl acetate 4:1) to

108

yield white crystals (800 mg, 72.1%). 1H NMR (300 MHz) 8.72 (d, 2H, ar), 8.34 (d, 2H, ar), 7.84 (dd, 2H, ar), 0.28 (s, 18H, SiCH3).

Synthesis of 5,5-Diethynyl-2,2-bipyridine (12) (Scheme 6.2)


CH3 H3C Si CH3

Br N N

Br

CuI triethylamine DMF

(P(PH3)3)2Pd2Cl2

CH3 H3C Si CH3 N N

CH3 Si CH3 CH3

Scheme 6.1 Reaction scheme for the synthesis of 5,5-bis((trimethylsilyl)-ethynyl)-2,2bipyridine.

Methanol (19 mL) and aqueous KOH (1.3 mL, 20 %) were added to a stirred solution of 5,5-bis((trimethylsilyl)-ethynyl)-2,2-bipyridine (750 mg, 2.15 mmol) in THF (38 mL) and the mixture was stirred at RT for 2 h. The reaction mixture was filtered, washed twice with deionized water (50 mL) and dried over MgSO4. The organic phase was separated off and the solvents were evaporated in vacuum to obtain an off-white crystalline solid. The product was purified by column chromatography (silica gel, hexanes/ethyl acetate 10:1) to afford 12 in the form of off-white crystals (324 mg, 73.7%), which were stored in the dark under Ar. 1H NMR (300 MHz) 8.77 (d, 2H, ar), 8.38 (d, 2H, ar), 7.88 (dd, 2H, d), 3.30 (s, 2H, CH). Anal. Calcd. for C14N2H8: C, 82.33% N, 13.71%; H, 3.94%. Found: C, 80.23%; N, 13.17%; H, 3.75%.

109

CH3 H3C Si CH3 N N

CH3 Si CH3 CH3

MeOH THF

KOH

12

Scheme 6.2 Reaction scheme for the synthesis of 5,5-Bis(ethynyl)-2,2-bipyridine (12).

Synthesis

of

Poly{2,2-bipyridine-5,5diylethynylene[2,5-bis(2-ethylhexyl)oxy-1,4-

phenylene]ethynylene} (BipyPPE1, 13) 1,4-Bis[(2-ethylhexyl)oxy]-2,5-diiodobenzene (10) (362 mg, 0.617 mmol), 5,5diethynyl-2,2-bipyridine (12, 150 mg, 0.734 mmol), p-iodotoluene (51.2 mg, 0.234 mmol), Pd(PPh3)2 (34.1 mg, 0.0295 mmol) and CuI (6.0 mg, 0.032 mmol) were combined in a mixture of toluene (23 mL) and diisopropyl amine (10 mL) and the reaction mixture was stirred under Ar at 70 C. The formation of what appeared to be ammonium iodide salts was observed immediately after the start of the reaction and at the same time the mixture became highly fluorescent. After 24 h the reaction was stopped by adding the mixture to a saturated aqueous EDTA solution (50 mL). After stirring for 2 h the aqueous phase had turned turquoise. The organic phase was collected, washed with deionized water (2 x 50 mL) and added drop-wise to vigorously stirred methanol. The reddish polymer thus precipitated was filtered off and dried in vacuum. The polymer was re-dissolved in CHCl3 (50 mL) and washed with saturated aqueous EDTA solution (3 x 50 mL). The organic layer was separated off and the aqueous phase was re-extracted with CHCl3 (50 mL). The combined organic layers were washed with deionized water (50

110

mL), concentrated in vacuum and introduced drop-wise to vigorously stirred methanol. The solid thus precipitated was filtered off, washed with hot hexanes (50 mL) and dried in vacuum to yield BipyPPE1 (13) as a red solid (227 mg, 62.4 %). 1H NMR (600 MHz) 8.82 (s, Bipy, 2H), 8.43 (s, Bipy, 2H), 7.92 (s, Bipy, 2H), 7.46, 7.17 (dd, 4H, ar/end groups), 7.34, 6.94 (2 x s, 2H, ar/end groups), 7.07 (s, 2H, ar), 3.96 (m, 4H, OCH2), 2.39 (s, 6H CH3/end groups), 1.81 (s, 6H, 2 x CH2 + 2 x CH), 1.58 (m, 16H, CH2), 1.24 (s, 20H, CH2), 0.89 (d, 6H, CH3); Xn=11.

Synthesis

of

Poly{(2,2-bipyridine-5,5diylethynylene[2,5-bis(2-ethylhexyl)oxy-1,4-

phenylene]ethynylene)-co-(2,5-dioctyloxy-1,4-diethynyl-phenylene-alt-2,5-bis(2'-ethylhexyloxy)-1,4-phenylene)} (BipyPPE2, 14) 1,4-Bis[(2-ethylhexyl)oxy]-2,5-diiodobenzene (10, 548 mg, 0.934 mmol), 1,4diethynyl-2,5-bis-octyloxy benzene (11, 285 mg, 0.743 mmol), 5,5-diethynyl-2,2bipyridine (12, 50.0 mg, 0.244 mmol), p-iodotoluene (41.3 mg, 0.189 mmol), Pd(PPh3)2 (47.6 mg, 0.0412 mmol) and CuI (8.2mg, 0.043 mmol) were combined in a mixture of toluene (30 mL) and diisopropyl amine (12 mL) and the reaction mixture was stirred under Ar at 70 C. The formation of what appeared to be ammonium iodide salts was observed immediately after the start of the reaction and at the same time the mixture became highly fluorescent. After 24 h the reaction was stopped by adding the mixture to a saturated aqueous EDTA solution (50 mL). After stirring for 2 h the aqueous phase had turned turquoise. The organic phase was collected, washed with deionized water (2 x 50 mL) and added drop-wise to vigorously stirred methanol. The reddish polymer thus precipitated was filtered off and dried in vacuum. The polymer was re-dissolved in CHCl3 (50 mL) and washed with saturated aqueous EDTA solution (3 x 50 mL). The

111

organic layer was separated off and the aqueous phase was re-extracted with CHCl3 (50 mL). The combined organic layers were washed with deionized water (50 mL), concentrated in vacuum and introduced drop-wise to vigorously stirred methanol. The solid thus precipitated was filtered off, washed with hot hexanes (50 mL) and dried in vacuum to yield BipyPPE2 (14) as a red solid (486 mg, 72.7 %). 1H NMR (600 MHz, 42 C) 8.82 (s, Bipy, 2H), 8.43 (s, Bipy, 2H), 7.92 (s, Bipy, 2H), 7.42, 7.12 (dd, 4H, ar/end groups), 7.29, 6.91 (s, 2H, ar/end groups), 7.07 (s, 2H, ar), 3.96 (m, 4H, OCH2), 2.39 (s, 6H CH3/End Groups), 1.81 (m, 6H, 2 x CH2 + 2 x CH), 1.58 (m, 16H, CH2), 1.24 (m, 20H, CH2), 0.89 (dd, 6H, CH3); Xn=16.

Preparative Synthesis of Bis(2,2-bipyridine)Cu(I)-hexafluorophosphate (15) 2,2-Bipyridine (502.7 mg, 3.218 mmol) and tetrakis(acetonitrile)Cu(I)hexafluorophosphate (599.7 mg, 1.609 mmol) were separately dissolved in anhydrous CHCl3 (5 mL) and CH3CN (6 mL), respectively. Over a period of 20 min the colorless tetrakis(acetonitrile)Cu(I)-hexafluorophosphate solution was added dropwise to the stirred colorless 2,2-bipyridine solution and the mixture immediately turned deep red. The solvents were evaporated in vacuum to yield 15 as a red solid (829 mg, 98.7%). UVVis (CHCl3): max = 433 nm and 523 nm (shoulder).

In-Situ Synthesis of Bis(2,2-bipyridine)Cu(I)-hexafluorophosphate (15) An aliquot of a solution of tetrakis(acetonitrile)Cu(I)-hexafluorophosphate in CH3CN (0.2 mL, 2.4210-4 M) was added to a solution of 2,2-bipyridine in CHCl3 (3 mL, 3.2310-5 M). The mixture was allowed to stand for 15 min before an optical

112

absorption spectrum was recorded. UV-Vis (CHCl3:CH3CN = 15:1 v/v): max = 348 nm, 433 nm and 523 nm (shoulder).

Optical Measurements of Polymer-Metal Complexes The polymers were dissolved in CHCl3 at a concentration of 1.9310-5 M (BipyPPE1, 13) and 3.2310-5 M (BipyPPE2, 14) (all concentrations are quoted with respect to the Bipy ligand comprised in the polymers). The metal complexes (tetrakis(acetonitrile)Cu(I)-hexafluorophosphate, cobalt tetrafluoroborate hexahydrate, nickel perchlorate hexahydrate, zinc perchlorate hexahydrate and cadmium perchlorate hydrate were dissolved in spectroscopic grade CH3CN (concentration range 2.1610-5 5.5510-4 M). Polymer-metal complexes were produced by adding aliquots of a solution of the selected metal complex in CH3CN (0.2 mL) to a solution of 13 or 14 in CHCl3 (3 mL); the concentration of the metal complex solution was varied to produce solutions of different [metal]:[Bipy] ratios but identical solvent composition of CHCl3:CH3CN = 15:1 v/v. The mixtures were allowed to stand for 15 min to 4 h after addition of the metal salt to the polymer solutions, before conducting optical studies on these systems.

Film Preparation The polymers (13 or 14) were dissolved in CHCl3 at a concentration of ~ 2.5 % w/v. Metal complexes (see above) were dissolved in CH3CN at appropriate concentrations and aliquots of metal solutions (0.1 mL) and polymer solution (0.2 mL) were mixed and allowed to stand for 15 20 min before thin films were produced by spin-coating onto glass substrates.

113

6.3 Synthesis and Characterization of BipyPPEs and Their Organometallic Networks The conjugated polymers employed in the present study, poly{2,2-bipyridine5,5diylethynylene[2,5-bis(2-ethylhexyl)oxy-1,4-phenylene]ethynylene} (BipyPPE1, 13) and a statistical copolymer comprising 5,5-diethynyl-2,2-bipyridine (12) and 1,4diethynyl-2,5-bis-(alkyloxy) benzene moieties (BipyPPE2, 14), were designed to

comprise different fractions of the Bipy moiety (13: x:y = 1:0; 14: x:y = 1:2.5, see Scheme 6.3) in order to explore the influence of the latter on the electronic properties of the polymer (vide infra). These polymers were synthesized by the Pd0-catalyzed crosscoupling reaction of 1,4-bis[(2-ethylhexyl)oxy]-2,5-diiodobenzene (10), 1,4-diethynyl2,5-bis-(octyloxy) benzene (11)16,20 and 5,5-diethynyl-2,2-bipyridine (12)17,18,19 (Scheme 6.3). Monomers 10-12 were synthesized according to established routes.16-20 The polymerizations were conducted in analogy to Klemms protocol8a with monomer concentrations that were slightly lower than routinely employed.16,20,21 If the reactions were conducted at higher monomer concentrations, the mixtures gelled; in view of the facts that the molecular weight was deliberately limited by offsetting the monomer stoichiometry and using p-iodotoluene as an end-capper and that the isolated polymers were soluble, this finding is indicative of the formation of network structures due to complexes between the Bipy moieties and the copper co-catalyst employed for the polymerization reaction. In order to remove the catalyst from the final product, the workup of the polymers included rigorous extraction with aqueous ethylenediamine tetraacetate. The isolated polymers were soluble in concentrations of at least 1 % w/v in solvents such as CHCl3 and toluene. 1H NMR spectra confirm the chemical structure of

114

OR2
x+y I

OR1

+
R1O 11

R2O 10

N
12

Pd(0), CuI Toluene, Diisopropylamine Iodobenzene (endcapper)

OR2 E N N
x

OR2

OR1
y

R2O

R2O

R1O OR2

N n-1

R1 =

R =

E= R2O

or

CH3

BipyPPE1 (13): x=1; y=0 and BipyPPE2 (14): x=1; y=2.5

Scheme 6.3 Synthesis and molecular structure of the 2,2-bipypridine-containing poly(2,5-dialkyloxy-p-phenylene BipyPPE2 (14). ethynylene)s BipyPPE1 (13) and

the polymers, reveal 4-toluene and 2,5-bis[(2-ethylhexyl)oxy]-4-iodobenzene end groups, and put the number-average molecular weights to ca. 3100 for 13 (number-average degree of polymerization, Xn, = 11) and 6100 for 14 (Xn, = 16), respectively. The photophysical characteristics of the un-complexed polymers were investigated in dilute solutions. A solvent mixture of CHCl3:CH3CN (15:1 v/v) was found suitable for the ligand-exchange experiments conducted (vide infra) and was therefore employed for all optical measurements conducted in this study. The results are

115

summarized in Table 6.1; UV-Vis absorption and photoluminescence (PL) emission spectra are shown in Figures 6.1 and 6.2 for 13 and 14, respectively. Polymer 13 exhibits

Type of Ion Ion free Cu+ Co2+ Ni2+ Zn2+ Cd2+

BipyPPE1 (13) Absorption max (nm) 423 452 458 458 464 468 Emission (nm) 459 -c -c -c 619 591 maxb

BipyPPE2 (14) Absorption max (nm) 440 447 450 450 450 448 Emission maxb (nm) 482 482 482 482 482, 646 482, 646

Table 6.1 Optical absorption and PL emission data of BipyPPE1 (13) and BipyPPE2 (14) after complexation with different metal ions.a
a

All spectra were recorded in CHCl3:CH3CN (15:1 v/v) at a concentration of

polymer-bound Bipy of 1.9310-5 M (13) or 3.2310-5 M (14) and an ion concentration of 2.1610-5 - 5.5510-4 M. b Excitation at 410 nm. c Emission is fully quenched.

116

Absorbance at 452 nm

0.8 Absorbance (-) 0.6 0.4 0.2 0.0 300

0.75

0.50

0.25

0.00 0.0

0.5

1.0 1.5 [Cu(I)] / [Bipy]

2.0

400

500 600 700 Wavelength (nm)

800

b
PL Intensity (10 cps)
8

10
9

8 6 4 2 0 400

Emission at 459nm

6 3 0 0.0

0.5 1.0 1.5 [Cu(I)] / [Bipy]

2.0

500 600 700 Wavelength (nm)

800

Figure 6.1 UV-Vis absorption (a) and PL emission (b) spectra acquired upon addition of tetrakis(acetonitrile)Cu(I)-hexafluorophosphate to BipyPPE1 (13) (concentration of polymer-bound Bipy = 1.9310-5 M) in CHCl3:CH3CN (15:1 v/v). Shown are spectra at selected [Cu+]:[Bipy] ratios of 0 (solid line), 0.09 (filled squares), 0.19 (filled circles), 0.28 (filled triangles), 0.38 (filled inverted triangles), 0.48 (filled rhombus), 0.57 (empty squares), 0.76 (empty circles), 0.96 (empty triangles) and 1.92 (empty inverted triangles). The insets show the absorption at 452 nm (a) and the emission at 459 nm (b) as a function of [Cu+]:[Bipy] ratio.

117

3
Absorbance at 440 nm 2.8 2.6 2.4 2.2 2.0 0.0 0.5 1.0 1.5 [Cu(I)] / [Bipy] 2.0

Absorbance (-)

0 300

400

500 600 700 Wavelength (nm)

800

Emission at 482 nm

b
PL Intensity (10 cps)

4 3 2 1 0 400

5 4 3 2 1 0 0.0 0.5 1.0 1.5 [Cu(I)] / [Bipy] 2.0

500

600

700

800

Wavelength (nm)

Figure 6.2 UV-Vis absorption (a) and PL emission (b) spectra acquired upon addition of tetrakis(acetonitrile)Cu(I)-hexafluorophosphate to BipyPPE2 (14) (concentration of polymer-bound Bipy = 3.2310-5 M) in CHCl3:CH3CN (15:1 v/v). Shown are spectra at selected [Cu+]:[Bipy] ratios of 0 (solid line), 0.1 (filled squares), 0.2 (filled circles), 0.3 (filled triangles), 0.4 (filled inverted triangles), 0.5 (filled rhombus), 0.6 (empty squares), 0.8 (empty circles), 1.0 (empty triangles) and 2.0 (empty inverted triangles).The insets show the absorption at 440 nm (a) and the emission at 482 nm (b) as a function of [Cu+]:[Bipy] ratio.

118

essentially identical absorption and emission characteristics as a polymer of the same structure investigated in detail before.8d The absorption spectrum features a strong band with max = 423 nm (Figure 6.1a), which in line with previous studies,7,8,16,20 is assigned to the * transition of the polymer backbone. Optical excitation of solutions of 13 at wavelengths that match this transition cause intense PL with max = 459 nm and features that are consistent with * fluorescence of the polymer (Figure 6.1b). The absorption and emission spectra of 13 are significantly blue-shifted compared to poly(2,5-dialkoxy1,4-phenylene ethynylene)s such as poly[2,5-dioctyloxy-1,4-diethynyl-phenylene-alt-2,5bis(2'-ethylhexyloxy)-1,4-phenylene] (EHO-OPPE, 4, max 436 and 474 nm for

absorption and emission, respectively).16,20 It was suggested earlier that this blue-shift is related to low torsional barriers for motions of the aromatic chain segments8d and the non-planar nature of the bipyridine moiety,22 which significantly limits the conjugation.11 14, which comprises fewer Bipy units than 13, displays essentially identical absorption (max = 440 nm) and emission (max = 482 nm) characteristics (Figure 6.2) as poly(2,5dialkoxy-1,4-phenylene ethynylene)s that comprise no Bipy.16,20 This is consistent with the formation of sufficiently long oligo(p-phenylene ethynylene) segments, which exhibit optical absorption and PL emission properties that have converged to those of macromolecular chains. These segments (presumably of statistical length) are linked by Bipy moieties, which (in their uncomplexed state) appear to exert little influence on the optical properties of the PPE segments, but appear to efficiently isolate individual chromophores (vide infra).8 [CuI(Bipy)2]X complexes represent a class of carefully studied, kinetically stable bis(bipyridine) complexes and we therefore elected to conduct a systematic complexation

119

study involving 13 or 14 on one hand and Cu+ on the other. Adopting the general protocol employed for the preparation of similar complexes by Mller et al.,23 we first conducted a series of reference experiments and prepared the known [CuI(Bipy)2]PF6 (15).24 This complex was synthesized by combining a solution of free (i.e. not polymerbound) 2,2-bipyridine in anhydrous CHCl3 with a solution of [CuI(CH3CN)4]PF6 in CH3CN, using two equivalents of ligand with respect to the metal ion. The rather

0.08 Absorbance (-)

0.04

0.00 300

400 500 600 Wavelength (nm)

Figure 6.3 UV-Vis absorption spectra of model compound 15 in CHCl3:CH3CN (15:1 v/v). Solid line: preparation in-situ. Dashed line: isolated compound.

concentrated mixture (~0.29 M Bipy) immediately adopted the characteristic deep red color of [CuI(Bipy)2]+. The complex could be quantitatively isolated as the hexafluorophosphate by evaporation of the solvents. We also conducted this ligandexchange reaction under more dilute conditions (~3.210-5 M Bipy) in a 15:1 v/v CHCl3:CH3CN mixture and characterized 15 in-situ by means of UV-Vis spectroscopy (Figure 6.3). The absorption spectrum of the mixture displays a maximum at 433 nm, a shoulder at 523 nm and a feature at higher energy (348 nm) and matches the one of the

120

isolated compound and the one published for [CuI(Bipy)2]ClO4 elsewhere, suggesting that under the selected conditions the ligand exchange proceeds smoothly and predominantly leads to the formation of 15. The main absorption band centered around 433 nm is characteristic of a Cu+-diimine system and is attributed to a metal to ligand charge transfer (MLCT) from the filled Cu 3d orbitals (3d10) to the lowest empty * orbitals (LUMO) of the 2,2-bipyridyl ligand.23,25 As it is the case for many other Cu+diimine systems, [CuI(Bipy)2]PF6 does not display any appreciable fluorescence. We observed that a similar exchange reaction takes place upon addition of CH3CN solutions of [CuI(CH3CN)4]PF6 to CHCl3 solutions of 13 or 14. If the polymer concentration was of the order of 1 % w/v, the mixtures gelled (14) or formed a gellous precipitate (13), indicating the formation of metallo-supramolecular networks with Bipymetal-Bipy cross-links. Gelation and/or formation of visible precipitation could be circumvented by reducing the absolute concentrations of the polymers (and concomitantly [CuI(CH3CN)4]PF6) to a level of Bipy (as comprised in the polymer) of between 1.9310-5 and 3.2310-5 M. In this case, the color of the polymer solutions changed instantaneously from a pale to a deep yellow upon addition of Cu+. With the goal of attaining insight into the formation of the polymer-metal complexes and to elucidate their optical properties, we titrated [CuI(CH3CN)4]PF6 to 13.26 UV-Vis absorption and PL emission spectra acquired in CHCl3:CH3CN are shown in Figure 6.1 for selected [Cu+]:[Bipy] ratios. As expected, the absorption spectrum of 13 changed significantly upon addition of Cu+. The intensity of the characteristic * transition associated with the conjugated polymer backbone around 423 nm systematically weakened when [CuI(CH3CN)4]PF6 was added and a new band centered around 452 nm

121

a
-10 PL Intensity/[Cu] 30

20

10 2 4 6 -PL Intensity 8

2.0 10 A/[Cu] 1.6 1.2 0.8 0.4 0.1 0.2 0.3


A

0.4

Figure 6.4 Representation of the UV-Vis absorption data (a) and PL emission data (b) shown in the inset of Figures 6.1a and 6.1b respectively in a Scatchard plot.

developed (Figure 6.1a). As can be seen from the inset in Figure 6.1a, the intensity of the new absorption band at 452 nm steadily intensified with increasing [Cu+]:[Bipy] ratio, before leveling off at a [Cu+]:[Bipy] ratio of about 0.5. Similarly, the polymers PL was gradually quenched upon addition of [CuI(CH3CN)4]PF6 (Figure 6.1b). Also this effect scaled with the [Cu+]:[Bipy] ratio and the PL was completely suppressed beyond a

122

OR N N
+

RO

RO N
M

OR

N N RO

OR

Scheme 6.4 Schematic representation of the formation of metallo-supramolecular networks via the formation of ligand-metal-ligand complexes with 2,2bipypridine-containing poly(2,5-dialkyloxy-p-phenylene ethynylene)s BipyPPE1 (13) and BipyPPE2 (14).

[Cu+]:[Bipy] ratio of about 0.5. Scatchart plots27 (Figure 6.4) of the data presented in Figure 6.1 are curved and characteristic of positive cooperative binding.28 Control experiments in which [CuI(CH3CN)4]PF6) was added to poly[2,5-dioctyloxy-1,4diethynyl-phenylene-alt-2,5-bis(2'-ethylhexyloxy)-1,4-phenylene] (EHO-OPPE, a PPE derivative lacking the Bipy recognition site) under similar conditions showed no optical changes. These results are consistent with the formation of BipyPPE1-Cu+-BipyPPE1 relatively large binding constants. In view of the similarities to the optical characteristics of the low-molecular model compound 15, we tentatively interpret the optical properties of the Cu+-BipyPPE1 networks with the formation of a non-radiative metal to ligand

123

charge transfer complex between the metals 3d and the * orbitals of polymer 13. The absorption maximum of the polymeric metal complex is red-shifted by 19 nm when compared to 15, presumably as a result of the extended conjugation of the polymeric ligand. The fact that the partially metalated 13 retained a significant extent of PL emission (Figure 6.1b) is indicative of limited exciton migration along the polymer to the non-radiative low-band-gap sites. This unusual feature was reported before7,8d and appears to be related to the de-conjugated nature of un-complexed, twisted Bipy moieties. These groups cause cross-links between the conjugated macromolecules (Scheme 6.4) and point to efficient optical insulation, which allows the co-existence of multiple chromophores on the same macromolecule. The fact that they display weak electronic coupling is in marked contrast to the PPE-based polymer systems reported by Swager and co-workers, which act as molecular wires and display energy migration over up to ~50 repeat units.20,29 The titration of [CuI(CH3CN)4]PF6 to 14 in CHCl3:CH3CN caused optical changes (Figure 6.2) that were qualitatively similar to those observed for the Cu+-BipyPPE1 complexes. The shift of the main absorption band from 440 to 447 nm (Figure 6.2a) was less pronounced than in the latter case ( max = 29 nm), due to the smaller band-gap of the un-complexed 14. In contrast to the Cu+BipyPPE1 complexes, the luminescence of polymer 14 was not fully extinguished upon addition of more than stoichiometric amounts of Cu+ (Figure 6.2b). This feature reflects the absence of dominant intermolecular energy transfer processes and significant internal absorption and is consistent with the statistical product distribution dictated by the nature of the polymerization reaction; the number-average degree of polymerization and the Bipy concentration in the polymer are both low, and due to the statistical incorporation of

124

a
-10 A / [Cu]

2.0 1.6 1.2 0.8 0.4 0.2 0.3 0.4 -A 0.5 0.6

-10 PL Intensity / [Cu]

1.60 1.20 0.80 0.40 2 3 8 -10 PL Intensity 4

Figure 6.5 Representation of the UV-Vis absorption data (a) and PL emission data (b) shown in the inset of Figures 6.2a and 6.2b respectively in a Scatchard plot.

the Bipy monomer a small fraction of the macromolecules does not comprise any Bipy moieties and therefore remains emissive upon addition of [CuI(CH3CN)4]PF6. Similar to 13 the optical changes caused by titration of 14 with Cu+ leveled off at a [Cu+]:[Bipy] ratio of about 0.5, but minor spectral changes can be discerned upon further addition of Cu+ (Figure 6.2). Scatchart plots (Figure 6.5) of the data presented in Figure 6.2 are almost linear and indicate the absence of cooperative binding. Thus, while these results 125

13

also point to the formation of BipyPPE2-Cu+-BipyPPE2 cross-links, the driving force for their formation appears to be less pronounced than in case of 13. This may reflect the fact that in contrast to 13, which features a regular structure of alternating 1,4-diethynyl-Bipy and 2,5-bis(2-ethylhexyl)oxy-1,4-phenylene moieties, the Bipy moieties comprised in 14 are statistically distributed, stifling the formation of BipyPPE2-Cu+-BipyPPE2 moieties at higher Cu+ concentrations due to geometric constraints and limitations with respect to chain mobility. In order to investigate the reversibility of the complexation of the Bipy-containing PPEs with Cu+, we have titrated a CHCl3:CH3CN solution comprising equimolar amounts of 13 and [CuI(CH3CN)4]PF6 with free (i.e. not polymer-bound) 2,2-bipyridine. Gratifyingly, upon addition of Bipy, the UV-Vis absorption (Figure 6.6a) and PL emission (Figure 6.6b) spectra converged back to those of the un-complexed polymer. At a [free Bipy]:[polymer-bound Bipy] ratio of 10:1 the optical properties of 13 were fully restored. Thus, the experiment unequivocally demonstrates that de-complexation of the polymer-metal system occurs upon addition of a competing ligand and that the ligand exchange reaction is fully reversible. With the objective to explore the influence of the metallic cross-linker on the optical properties of the polymers at hand, we have conducted complexation experiments involving 13 or 14 and perchlorates of Co2+, Ni2+, Zn2+ or Cd2+. With reference to the above-summarized complexation study with Cu+ and with the objective to form [M(II)(Bipy)2]2+ complexes that function as cross-links between the macromolecules, the [M2+]:[Bipy] ratio was kept constant at a level of 0.5:1. UV-Vis absorption and PL emission spectra acquired in CHCl3:CH3CN are shown for 13 and 14 in Figures 6.7 and

126

a
Absorbance (-)

1.0 0.8 0.6 0.4 0.2 0.0 300

400

500 600 700 Wavelength (nm)

800

b
PL Intensity (10 cps)
8

12 9 6 3 0 400

500

600

700

800

Wavelength (nm)

Figure 6.6 UV-Vis absorption (a) and PL emission (b) spectra acquired upon addition of free 2,2-bipyridine to a mixture of BipyPPE1 (13) (concentration of polymer-bound Bipy = 1.9310-5 M) and tetrakis(acetonitrile)Cu(I)hexafluorophosphate (ratio of [Cu+]:[Bipy comprised in the polymer] = 1:0.96) in CHCl3:CH3CN (15:1 v/v). Shown are spectra for ratios of [free Bipy]:[polymer-bound Bipy] of 0:1 (dashed line), 1:1 (filled squares), 2:1 (filled circles), 5:1 (filled triangles) and 10:1 (filled rhombus). The spectra of neat BipyPPE1 (13, solid line) are included for reference.

127

a
Absorbance (-)

0.8 0.6 0.4 0.2 0.0 300

400

500 600 700 Wavelength (nm)

800

b
PL Intensity (10 cps)
8

8 6 4 2 0 400

500

600 700 Wavelength (nm)

800

Figure 6.7 UV-Vis absorption (a) and PL emission (b) spectra acquired upon addition of cobalt(II)tetrafluoroborate hexahydrate (filled squares), nickel(II)perchlorate hexahydrate (empty circles), zinc(II)perchlorate hexahydrate (filled triangles), and cadmium(II)perchlorate hydrate (filled circles) to BipyPPE1 (13) (concentration of Bipy comprised in the polymer in solution = 1.9310-5 M) in CHCl3:CH3CN (15:1 v/v). Shown are spectra for [M2+]:[Bipy] ratios of 0.5:1. The spectra of neat BipyPPE1 (13, solid line) are included for reference.

128

6.8 and the results are summarized in Table 6.1. The addition of Co(ClO4)2 or Ni(ClO4)2 led to very similar optical changes as were observed before for [CuI(CH3CN)4]PF6 at the same [M2+]:[Bipy] ratio. Complexation with Co2+ or Ni2+ shifted the maximum of the absorption band of 13 (Figure 6.7a) from 423 to 458 nm (Cu+: 452 nm) and that of 14 (Figure 6.8a) from 440 to 450 nm (Cu+: 447 nm). The PL of 13 was completely quenched upon addition of Co2+ or Ni2+ (Figure 6.7a), while 14 displayed the weak residual emission (Figure 6.8b) that appears to be associated with the Bipy-free fraction of macromolecules (vide supra). These results are consistent with the formation of BipyM2+-Bipy moieties and reflect the fact that Ni2+ exhibits a strong tendency for the formation of metal to ligand charge transfer complexes with imine ligands.30 Interestingly, the addition of Zn(ClO4)2 or Cd(ClO4)2 caused a somewhat different response. Complexation with these metals shifted the maximum of the absorption band of 13 from 423 to 464 (Zn2+) and 468 nm (Cd2+) nm (Figure 6.7a) and that of 14 from 440 to 450 (Zn2+) and 448 nm (Cd2+) (Figure 6.8a). Thus, the changes were somewhat more pronounced than in case of the other metals investigated here. The characteristic PL band of 13 was completely quenched upon addition of Zn2+ or Cd2+, while 14 displayed the weak residual emission associated with the Bipy-free macromolecules (vide infra). In case of both metals, however, new, broad, structure-less emission bands centered at 619 (BipyPPE1-Zn2+), 591 (BipyPPE1-Cd2+), and 646 nm (BipyPPE2-Zn2+ and BipyPPE2Cd2+) developed (Figures 6.7b and 6.8b). These results reflect the fact that Zn2+ and Cd2+ both exhibit a fully occupied d-orbital (Zn2+: 3d10, Cd2+: 4d10), which frequently displays a weak tendency for the formation of metal to ligand charge transfer.31 Hence, the complexation of these metals with Bipy-containing

129

a
Absorbance (-) 2

0 300

400

500 600 700 Wavelength (nm)

800

b
PL Intensity (10 cps)
8

4 3 2 1 0 400

500 600 700 Wavelength (nm)

800

Figure 6.8 UV-Vis absorption (a) and PL emission (b) spectra acquired upon addition of cobalt(II)tetrafluoroborate hexahydrate (filled squares), nickel(II)perchlorate hexahydrate (empty circles), zinc(II)perchlorate hexahydrate (filled triangles), and cadmium(II)perchlorate hydrate (filled circles) to BipyPPE2 (14) (concentration of polymer-bound Bipy =3.2310-5 M) in CHCl3:CH3CN (15:1 v/v). Shown are spectra for [M2+]:[Bipy] ratios of 0.5:1. The spectra of neat BipyPPE2 (14, solid line) are included for reference.

130

a
Absorbance (a.u.)

1.0 0.8 0.6 0.4 0.2 0.0 300

400

500 600 700 Wavelength (nm)

800

b
Pl Intensity (10 cps)
6

4.5

3.0

1.5

0.0 400

500 600 Wavelength (nm)

700

Figure 6.9 (a) UV-Vis absorption and (b) PL emission spectra of spin-coated films of complexes produced by ligand-exchange reactions between BipyPPE1 (13) and zinc perchlorate hexahydrate (filled rhombus), or cadmium perchlorate hydrate (filled squares). Shown are spectra for [metal]:[Bipy] ratios of 0.5:1. The spectra of neat BipyPPE1 (13, solid line) are included for reference.

131

a
Absorbance (a.u.)

1.2 0.9 0.6 0.3 0.0 300

400

500 600 700 Wavelength (nm)

800

b
PL Intensity (10 cps)
6

0 400

500 600 Wavelength (nm)

700

Figure 6.10 (a) UV-Vis absorption and (b) PL emission spectra of spin-coated films of complexes produced by ligand-exchange reactions between BipyPPE2 (14) and zinc perchlorate hexahydrate (filled rhombus), or cadmium perchlorate hydrate (filled squares). Shown are spectra for [metal]:[Bipy] ratios of 0.5:1. The spectra of neat BipyPPE2 (14, solid line) are included for reference.

132

polymers does not usually lead to MLCT complexes. Rather, the optical changes appear to be related to a significant reduction of the polymers * transition on account of a planarization of the Bipy moiety11,12 as well as electron density variation upon complexation with the electron-poor metals.14 In view of the fully occupied d-orbital of the metal, the observed emission cannot be related to a d-d transition, but appears to be caused by intra-ligand * transitions. Homogeneous thin films of good optical quality could be produced by spincoating solutions of the polymer-metal complexes. The films thus produced were insoluble in CHCl3, which is a good solvent for the un-complexed polymers, consistent with the formation of cross-linked BipyPPE1-metal and BipyPPE2-metal networks. On the other hand, if free (i.e. not polymer-bound) 2,2-bipyridine was added to the solvent, the films readily re-dissolved, demonstrating again that the complexation reactions are reversible. The metal-complexed films show similar absorption spectra as in solution, with absorption maxima that range from 420 nm (uncomplexed film) to 454 nm (Zn2+) in case of 13 (Figure 6.9a) and from 427 nm (uncomplexed film) to 462 nm (Zn2+) in case of 14 (Figure 6.10a). The emissive characteristics of the Zn(II) and Cd(II)-complexes of 13 and 14 were retained after processing these materials into thin films (Figures 6.9b and 6.107b), and these films displayed orange-red emission with maxima at 612 nm and 609 nm, respectively. While the development of red solid-state emitters was clearly beyond the original focus of our work, the relatively intense emission - in particular of the Zncoordinated PPEs appears to represent an interesting by-product of our work.

133

6.4 Conclusions In summary, we have presented a new synthetic framework that allows for the synthesis and processing of conjugated organometallic polymer networks. We have shown at the example of poly(p-phenylene ethynylene) derivatives that the introduction of 2,2-bipyridine moieties leads to conjugated polymers, which are formidable precursors for metallo-supramolecular PPE networks that are accessible via ligandexchange reactions. The resulting three-dimensionally cross-linked, conjugated BipyPPE1-metal and BipyPPE2-metal networks display interesting optoelectronic properties. As expected, the coordination of the metal markedly influences the photophysical characteristics of the polymer. It is obvious that the approach is broadly applicable to other conjugated polymer platforms and also to other ligands. We are aware of the fact that the (pseudo)-tetrahedral coordination of four-coordinated metals (as in case of some of the complexes explored here) might ultimately not represent the most desirable geometry from a charge-transfer point of view, and that the ionic nature of the ions used here might lead to charge trapping. After having employed these complexes in the present study to demonstrate the possibility to prepare a new family of -conjugated organometallic polymer hybrid systems, we envision to extend the present concepts to the use of square planar complexes and conduct experiments related to the charge-carrier transport in such systems.

134

6.5 References 1 See for example: (a) Wang, Z. H.; Li, C.; Scherr, E. M.; MacDiarmid, A. G.; Epstein, A. J. Phys. Rev. Lett. 1991, 66, 1745. (b) Aleshin, A.; Kiebooms, R.; Menon, R.; Wudl, F.; Heeger, A. J. Phys. Rev. B 1997, 56, 3659. 2 See for example: (a) Green, N.; Grove, D. M.; Howard, J. A.; Spencer, J. L.; Stone, F. G. A.; J. Chem. Soc., Chem. Comm. 1976, 759. (b) Boag, N. M.; Green, M.; Grove, D. M.; Howard, J. A. K.; Spencer, J. L.; Stone, F. G. A. J. Chem. Soc., Dalton Trans. 1980, 2170. 3 4 5 Howard, J. A.; Sutcliffe, R.; Tse, J. S.; Mile, B. Organometallics 1984, 3, 859. Kaes, C.; Katz, A.; Hosseini, M. W. Chem. Rev. 2000, 100, 3553. Collman, J. P.; Hegedus, L. S.; Norton, J. R.; Finke, R. G. Principles and Applications of Organotransition Metal Chemistry; University Science Books, Sausalito California, 1987. 6 For a recent review see: Schubert, U. S.; Eschbaumer, C. Angew. Chem. Int. Ed. 2002, 41, 2892. 7 (a) Ley, K. D.; Whittle, C. E.; Bartberger, M. D.; Schanze, K. S. J. Am. Chem. Soc. 1997, 119, 3423. (b) Ley, K. D.; Schanze, K. S. Coord. Chem. Rev. 1998, 171, 287. 8 (a) Egbe, D. A. M.; Klemm, E. Macromol. Chem. Phys. 1998, 199, 2683. (b) Pautzsch, T.; Klemm, E. Macromolecules 2002, 35, 1569. (c) Pautzsch, T.; Blankenburg, L.; Klemm, E. J. Polym. Sci.: Part A: Polym. Chem. 2004, 42, 722. (d) Grummt, U. W.; Pautzsch, T.; Birckner, E.; Sauerbrey, H.; Utterodt, A.;

135

Neugebauer, U.; Klemm, E. J. Phys. Org. Chem. 2004, 17, 199. (e) Pautzsch, T.; Klemm, E. J. Polym. Sci.: Part A: Polym. Chem. 2004, 42, 2911. 9 An example for complexation for ligand exchange can be found in the

supplementary material to Ref. 7d. 10 (a) Yamamoto, T.; Maruyama, T.; Zhou, Z.; Ito, T.; Fukuda, T.; Yoneda, Y.; Begum, F.; Ikeda, T.; Sasaki, S.; Takezoe, H.; Fukuda, A.; Kubota, K. J. Am. Chem. Soc. 1994, 116, 4832. (b) Saitoh, Y.; Yamamoto, T. Chem. Lett. 1995, 785. 11 12 13 Wang, B.; Wasielewski, M. R. J. Am. Chem. Soc. 1997, 119, 12. Peng, Z.; Gharavi, A. R.; Yu, L. J. Am. Chem. Soc. 1997, 119, 4622. Liu, B.; Yu, W. L.; Liu, S. Y.; Lai, Y. L.; Huang, W. Macromolecules 2001, 34, 7932. 14 15 Zhang, M.; Lu, P.; Ma, Y.; Shen, J. J. Phys. Chem. B 2003, 107, 6535. Bangcuyo, C. G.; Rampey-Vaughn, M. E.; Quan, L. T.; Angel, S. M.; Smith, M. D.; Bunz, U. H. F. Macromolecules 2002, 35, 1563. 16 17 18 19 20 21 Weder, C.; Wrighton, M. S. Macromolecules 1996, 29, 5157. Schwab, P. F. H.; Fleischer, F.; Michl, J. J. Org. Chem. 2002, 67, 443. Iyer, P. K.; Beck, J. B.; Weder, C.; Rowan, S. R. Chem. Comm. 2005, 319. Grosshenny, V; Romero, F. M.; Ziessel, R. J. Org. Chem. 1997, 62, 1491. Swager, T. M.; Gil, C. J.; Wrighton, M. S. J. Phys. Chem. 1995, 99, 4886. Sanechika, K.; Yamamoto, T.; Yamamoto, A. Bull. Chem. Soc. Jpn. 1984, 57, 752. 22 Cumper, C. W. N.; Ginman, R. F. A.; Vogel, A. I. J. Chem. Soc. 1962, 1188.

136

23

Mller, E.; Piguet, C.; Bernardinelli, G.; Williams, A. F. Inorg. Chem. 1984, 27, 849.

24

Munakata, M.; Kitagawa, S.; Asahara, A.; Masuda, H. Bull. Chem. Soc. Jpn. 1987, 60, 1927.

25

(a) Daul, C.; Schlpfer, C. W.; Goursot, A.; Pnigault, E.; Weber, J. Chem. Phys. Lett. 1981, 78, 304. (b) Phifer, C. C.; McMillin, D. R. Inorg. Chem. 1986, 25, 1329.

26

The term titration is employed here to convey the spirit of the experiment. For experimental reasons, the actual data were collected from independently prepared samples with selected [Cu+]:[Bipy] ratios, rather than by gradually adding a metal to a polymer solution.

27 28

Scatchard, G. Ann. N.Y. Acad. Sci. 1949, 51, 660. Conners, K.A. Binding Constants - The Measurement of Molecular Complex Stability; Wiley-Interscience, New York, 1987.

29

(a) Zhou, Q.; Swager, T. M. J. Am. Chem. Soc. 1995, 117, 12593. (b) Swager, T. M. Acc. Chem. Res. 1998, 31, 201.

30

(a) Bedel, S. A.; Martell, A. E. Inorg. Chem.1983, 22, 364. (b) Harris, C. M.; McKenzie, E. D. J. Inorg. Nucl. Chem. 1967, 29, 1047.

31

Jaeger, F. M.; van Dijk, J. A. Z. Anorg. Ch. 1938, 227, 273.

137

Chapter 7: Synthesis and Characterization of Cross-linked Poly(p-phenylene ethynylene)s* 7.1 Introduction After having demonstrated that organometallic conjugated polymer networks can exhibit significantly improved charge-transport characteristics (Chapter 5.4) we sought to extend the scope of this general design approach, to conjugated polymer networks that rely on covalent conjugated cross-links. Interestingly, materials with this structural motif have received little attention in the past,1 possibly because of the difficulty of introducing conjugated cross-links and retaining adequate processibility of the resulting material. In this Chapter we report on the synthesis, characterization and optical properties of crosslinked poly(p-phenylene ethynylene)s, and discuss processing options for this general class of materials. We also show here that the challenge regarding the processing of such cross-linked networks can be overcome by synthesizing these materials in the form of (small) spherical particles, which can be processed from (aqueous) dispersions. Applying concepts employed for the preparation of dispersions of linear conjugated polymers2,3,4 and exploiting the fact that many metal-catalyzed cross-coupling reactions are tolerant to the presence of water,5 we demonstrate for the first time that (cross-linked) conjugated polymer particles can be conveniently produced by conducting metal-catalyzed crosscoupling reactions in aqueous emulsions. The size of the resulting particles can be readily tuned over a wide range (nm to mm) by modifying the reaction conditions, and it is apparent that the approach is universally applicable for many polymer systems.
*

Parts of this Chapter have been published: (a) Hittinger, E.; Kokil, A.; Weder, C. Angew. Chem. Int. Ed. 2004, 43, 1807. (b) Hittinger, E.; Kokil, A.; Weder, C. Macromol. Rapid Commun. 2004, 25, 710.

138

7.2 Experimental Section Materials and Methods 2,5-Diiodo-4-[(2-ethylhexyl)oxy]methoxybenzene (16), 1,4-Bis[(2-ethylhexyl) oxy]-2,5-diiodobenzene (10) and 1,4-diethynyl-2,5-bis-(octyloxy)benzene (11) were prepared as described previously. All other reagents, solvents, and catalysts were purchased from Aldrich Chemical Co. or FLUKA (analytical grade quality, if not indicated otherwise), and were used without further purification. The conjugated polymer network particles were re-dispersed by ultrasonicating the polymers for typically 24 h in toluene using a Fisher Scientific FS-60 ultrasonic bath. SEM images were acquired using a Hitachi model S-4500 field emission scanning microscope. Infrared (IR) spectra were measured using KBr pellets on a Bomem MB104 Michelson Series FT-IR. Optical microscopy was done on an Olympus BX-60 microscope with a Spot Insight digital camera (Model 3.2.0). Photoluminescence (PL) spectra were measured under excitation at 366 nm on a SPEX Fluorolog 3 (Model FL3-12). Corrections for the spectral dispersion of the Xe-lamp, the instrument throughput, and the detector response were applied. Elemental analyses were performed by Prevalere Life Sciences, Inc.

Polycondensations All polymerization reactions were conducted under a protective argon atmosphere with degassed, absolute solvents.

139

Reaction 1 11 (40.8 mg, 0.105 mmol), 17 (20.8 mg, 0.066 mmol), Pd(PPh3)4 (3.8 mg, 0.003 mmol), and CuI (2.5 mg, 0.013 mmol) were combined in a mixture of toluene (2.2 mL) and diisopropylamine (1 mL) and the reaction mixture was heated to 70 oC. After 24 h the reaction mixture had turned into a fragile gel and it was quenched by cooling to room temperature. The gel was transferred into slowly stirred toluene (50 mL) and methanol (300 mL) was gradually added over the course of 8 h. The product was collected and dried overnight in vacuum at room temperature to yield polymer O-OPPE-X (18a) as a brown solid (38.2 mg, 84%). FTIR: 3425 (m, broad), 2951 (s), 2930 (s), 2850 (s), 2752 (w), 2599 (w), 2476 (w), 2356 (w), 2205 (m), 1647 (m), 1558 (w), 1506 (s), 1458 (m), 1412 (m), 1377 (m), 1262 (s), 1217 (s), 1147 (w), 1092 (s), 1028 (s), 975 (w), 860 (m), 823 (w), 803 (s) cm-1. Anal. Calcd: C 83.84%, H 8.97%, N 0.00%, Br 0.00%; Found: C 70.89%, H 8.14%, N 0.09%, Br 10.49%.

Reaction 2 This reaction was conducted in analogy to Reaction 1, but using 16 (12.0 mg, 0.025 mmol), 11 (36.4 mg, 0.094 mmol), 17 (14.5 mg, 0.046 mmol), Pd(PPh3)4 (3.8 mg, 0.003 mmol), CuI (2.7 mg, 0.014 mmol), toluene (1.1 mL), diisopropylamine (0.5 mL) and adding an additional portion of toluene (1 mL) after 18 h. Polymer MEHO-OPPEX0.5 (18b) was obtained as a brown solid (28.3 mg, 62%). FTIR: 3447 (m, broad), 2955 (s), 2920 (s), 2856 (s), 2793 (w), 2355 (w), 2274 (w), 2204 (m, broad), 1653 (m), 1635 (w), 1506 (s), 1466 (s), 1420 (s), 1381 (s), 1273 (s), 1217 (s), 1119 (w), 1040 (s), 976

140

(w), 860 (m), 826 (w), 812 (w), 719 (w) cm-1. Anal. Calcd: C 82.50%, H 9.04%, N 0.00%, Br 0.00%; Found; C 76.24%, H 8.74%, N 0.00%, Br 6.37%.

Reaction 3 This reaction was conducted in analogy to Reaction 1, but using 16 (23.7 mg, 0.049 mmol), 11 (46 mg, 0.119 mmol), 17 (11.3 mg, 0.036 mmol), Pd(PPh3)4 (3.0 mg, 0.003 mmol), CuI (2.0 mg, 0.010 mmol), toluene (1.1 mL) and diisopropylamine (0.5 mL), and adding an additional portion of toluene (0.5 mL) after 18 h. Polymer MEHOOPPE-X1.3 (18c) was obtained as a brown solid (58.6 mg, 98%). FTIR: 3471 (m, broad), 2958 (s), 2925 (s), 2902 (m), 2849 (s), 2799 (w), 2348 (w), 2316 (w), 2204 (m, broad), 1721 (w), 1644 (w), 1600 (m), 1533 (w), 1508 (s), 1499 (s), 1465 (s), 1418 (s), 1386 (s), 1273 (s), 1215 (s), 1143 (w), 1120 (w), 1082 (w), 1038 (s), 988 (w), 862 (m), 722 (m) cm-1. Anal. Calcd: C 81.7%, H 9.26%, N 0.00%, Br 0.00%; Found: C 76.85%, H 8.61%, N 0.10%, Br 4.14%.

Reaction 4 This reaction was conducted in analogy to Reaction 1, but using 16 (68.0 mg, 0.139 mmol), 11 (71.5 mg, 0.184 mmol), 17 (9.9 mg, 0.031 mmol), Pd(PPh3)4 (7.5 mg, 0.006 mmol), and CuI (1.0 mg, 0.005 mmol). Polymer MEHO-OPPE-X4.4 (18d) was obtained as a brown solid (43.7 mg, 84%). FTIR: 3438 (m, broad), 3239 (w), 3192 (w), 2955 (s), 2929 (s), 2855 (s), 2738 (w), 2345 (w), 2207 (w, broad), 1870 (w), 1639 (m, broad), 1542 (m), 1508 (s), 1458 (s), 1420 (s), 1388 (s), 1272 (s), 1217 (s), 1150 (w),

141

1092 (m), 1035 (s), 976 (w), 865 (m), 809 (m) cm-1. Anal. Calcd: C 80.81%, H 9.33%, N 0.00%, Br 0.00% Found: C 78.66%, H 8.93%, N 0.01%, Br 1.78%.

Reaction 5 This reaction was conducted in analogy to Reaction 1, but using 16 (67.1 mg, 0.137 mmol), 11 (61.1 mg, 0.158 mmol), 17 (4.0 mg, 0.013 mmol), Pd(PPh3)4 (7.3 mg, 0.006 mmol), CuI (1.4 mg, 0.007 mmol), toluene (1.1 mL) and diisopropylamine (0.5 mL), and adding an additional portion of toluene (1 mL) after 18 h. Polymer MEHOOPPE-X10.5 (18e) was obtained as an orange solid (50.9 mg, 55%). FTIR: 3454 (m, broad), 3243 (w), 3167 (w), 3046 (w), 3001 (w), 2960 (s), 2928 (s), 2856 (s), 2654 (w), 2350 (w), 2039 (w, broad), 1869 (w), 1845 (w), 1716 (w), 1652 (w), 1639 (m, broad), 1540 (m), 1508 (s), 1457 (s), 1419 (s), 1387 (s), 1272 (s), 1214 (s), 1144 (w), 1083 (w), 1033 (s), 930 (w), 863 (m), 804 (m) cm-1. Anal. Calcd: C 80.44%, H 9.41%, N 0.00%, Br 0.00% Found: C 77.58%, H 9.37%, N 0.03%, Br <0.74%.

Reaction 6 A portion (ca. 0.5 mL) of a solution of 16 (93.5 mg, 0.191 mmol), 11 (106.2 mg, 0.274 mmol), 17 (17.2 mg, 0.055 mmol), Pd(PPh3)4 (12.4 mg, 0.011 mmol), and CuI (4.0 mg, 0.021 mmol) in a mixture of toluene (3.5 mL) and diisopropylamine (1.5 mL) was carefully added to a 100 mL round-bottom flask containing degassed water (50 mL). The two-phase system was heated to 70 oC. The reaction was quenched after 24 hrs by cooling to room temperature and the free-standing film of MEHO-OPPE-X3.4 (18f), which had formed on top of the aqueous phase, was carefully removed from the flask.

142

The product was placed in toluene (10 mL) and methanol (100 mL) was gradually added over the course of 3 h.

Millimeter-sized Particles 16 (73.6 mg, 0.151 mmol), 11 (93.0 mg, 0.240 mmol), 17 (22.3 mg, 0.071 mmol), Pd(PPh3)4 (7.3 mg, 0.006 mmol), and CuI (1.7 mg, 0.009 mmol) were combined in a mixture of toluene (2.2 mL) and diisopropylamine (1 mL). The mixture was stirred for 60 seconds, and degassed water (60 mL) was added. The reaction mixture was heated to 70 oC and vigorously stirred for 24 h. The reaction mixture, whose the organic phase had solidified into many individual, millimeter-sized particles, was cooled to room temperature. The product was filtered off, suspended in toluene (15 mL) and methanol (100 mL) was added in 20 min intervals over the course of 3 h. The product was collected and dried overnight in vacuum at room temperature to yield brown spherical polymer particles (127.3 mg, 95.6%). FTIR: 3452 (m, broad), 3056 (w), 2960 (m), 2924 (s), 2854 (s), 2743 (w), 2353 (w), 2208 (m, broad), 1511 (s), 1460 (s), 1420 (s), 1384 (s), 1263 (s), 1215 (s), 1088 (s), 1301 (s), 865 (m), 801 (s), 723 (m) cm-1. Anal. Calcd: C 80.77 %, H 9.40 %, N 0 %, Br 0 %; Found: C 76.17 %, H 8.56 %, N 0.06 %, Br <1.76 %.

Micrometer-size Particles 16 (135.8 mg, 0.278 mmol), 11 (144.1 mg, 0.372 mmol), 17 (19.4 mg, 0.061 mmol), Pd(PPh3)4 (15.0 mg, 0.013 mmol), and CuI (2.3 mg, 0.012 mmol) were combined in a mixture of toluene (4.4 mL) and diisopropylamine (2 mL). The mixture was stirred for 60 seconds, and a degassed aqueous solution (60 mL) of sodium dodecyl sulfate (0.8

143

g) was added. The reaction mixture was heated to 70 oC and stirred vigorously for 24 h. The reaction mixture, which had turned into an orange opaque suspension, was cooled to room temperature and the solvents were evaporated in vacuo. Toluene (100 mL) was added to the remaining solid, and the mixture was placed into an ultrasonic bath for 1 h. The solid content of the resulting suspension was separated by centrifugation and the supernatant clear solution was discarded. This procedure was repeated with

toluene/methanol mixtures of volume rations of 3:1, 1:1 and 1:3, pure methanol, and twice with toluene. The product was collected and dried overnight in vacuum to yield a brown solid (189.8 mg, 89%). FTIR: 3440 (m, broad), 2963 (s), 2917 (s), 2854 (s), 2332 (w), 2187 (w), 1897 (w), 1638 (m, broad), 1514 (s), 1465 (s), 1420 (s), 1387 (s), 1278 (s), 1218 (s), 1094 (w), 1034 (s), 970 (w), 853 (m), 726 (w) cm-1 Anal. Calcd: C 80.77 %, H 9.40 %, N 0 %, Br 0 %, S 0 %; Found: C 76.91 %, H 8.82 %, N 0.12 %, Br 0 %, S 0.17 %.

Nanometer-size Particles (a) 16 (108.0 mg, 0.221 mmol), 11 (114.9 mg, 0.296 mmol), 17 (15.2 mg, 0.048 mmol), Pd(PPh3)4 (14.2 mg, 0.012 mmol), CuI (3.5 mg, 0.018 mmol), and sodium dodecyl sulfate (24.3 mg) were combined in a mixture of toluene (3.3 mL), diisopropylamine (1.5 mL), and water (9 mL). The reaction mixture was heated to 70oC in an ultrasonic bath and kept at this temperature for 24h. The reaction mixture was cooled to room temperature, macroscopic solid polymer particles were filtered off, and the solvents were evaporated in vacuo. Toluene (3 mL) was added to the remaining solid, the mixture was placed into an ultrasonic bath for 24 h, and the suspension was added

144

drop-wise into well-stirred methanol (100 mL). The solid fraction of the resulting suspension was separated by centrifugation and was dried overnight in vacuum to yield a brown solid (7.5 mg, 4.4%).

Nanometer-size Particles (b) 10 (114.1 mg, 0.194 mmol), 11 (100.1 mg, 0.261 mmol), 17 (16.4 mg, 0.052 mmol), Pd(PPh3)4 (11.7 mg, 0.010 mmol), CuI (4.7 mg, 0.024 mmol), and sodium dodecyl sulfate (45.1 mg) were combined in a mixture of toluene (2.8 mL), diisopropylamine (1.3 mL), and water (21 mL). The reaction mixture was heated to 70oC in an ultrasonic bath and kept at this temperature for 24h. The reaction mixture was cooled to room temperature, macroscopic solid polymer particles were filtered off, and the solvents were evaporated in vacuo. Toluene (3 mL) was added to the remaining solid, the mixture was placed into an ultrasonic bath for 24 h, and the suspension was added drop-wise into well-stirred methanol (100 mL). The solid fraction of the resulting suspension was separated by centrifugation and was dried overnight in vacuum to yield a brown solid (82.3 mg, 48.9%).

Nanometer-size Particles (c) 10 (116.9 mg, 0.199 mmol), 11 (107.5 mg, 0.280 mmol), 17 (13.6 mg, 0.043 mmol), Pd(PPh3)4 (11.7 mg, 0.010 mmol), CuI (3.3 mg, 0.016 mmol), and sodium dodecyl sulfate (130.7 mg) were combined in a mixture of toluene (2.8 mL), diisopropylamine (1.3 mL), and water (30 mL). The reaction mixture was heated to 70oC in an ultrasonic bath and kept at this temperature for 24h. The reaction mixture was

145

cooled to room temperature, macroscopic solid polymer particles were filtered off, and the solvents were evaporated in vacuo. Toluene (3 mL) was added to the remaining solid, the mixture was placed into an ultrasonic bath for 24 h, and the suspension was added drop-wise into well-stirred methanol (100 mL). The solid fraction of the resulting suspension was separated by centrifugation and was dried overnight in vacuum to yield a brown solid (114.5 mg, 64.9%).

Nanometer-size Particles (d) 10 (150.9 mg, 0.257 mmol), 11 (106.7 mg, 0.278 mmol), 17 (8.3 mg, 0.026 mmol), Pd(PPh3)4 (13.2 mg, 0.011 mmol), CuI (12.1 mg, 0.061 mmol), and sodium dodecyl sulfate (45.1 mg) were combined in a mixture of toluene (3 mL), diisopropylamine (1.3 mL), and water (21 mL). The reaction mixture was heated to 70oC in an ultrasonic bath and kept at this temperature for 24h. The reaction mixture was cooled to room temperature, macroscopic solid polymer particles were filtered off, and the solvents were evaporated in vacuo. Toluene (3 mL) was added to the remaining solid, the mixture was placed into an ultrasonic bath for 24 h, and the suspension was added drop-wise into well-stirred methanol (100 mL). The solid fraction of the resulting suspension was separated by centrifugation and was dried overnight in vacuum to yield a brown solid (73.4 mg, 37.9%).

7.3 Bulk Synthesis and Characterization of Cross-linked Poly(p-phenylene ethynylenes)s The present study was based on polymers synthesized by the palladium-catalyzed crosscoupling polycondensation of 2,5-diiodo-4-[(2-ethylhexyl)oxy]methoxybenzene (16) and

146

1,4-diethynyl-2,5-bis-(octyloxy)benzene (11) (Scheme 7.1). The linear poly(p-phenylene ethynylene) (PPE) based on only 16 and 11 poly[2,5-dioctyloxy-1,4-diethynylphenylene-alt-5-(2'-ethylhexyloxy)-2-methoxy-1,4-phenylene] (MEHO-OPPE) is highly soluble and its optoelectronic properties are well-documented6 and representative for
OR I R 1O
3

OR2 + Br RO
2

Br Br

I +

16/10

11
toluene diisopropylamine Pd(PPh3)4 CuI

17

OR

OR

OR2

R1O

R2O

m
w

R2O

RO

OR OR2

R2O

18

Scheme 7.1 Synthesis of cross-linked PPEs (18) by the palladium-catalyzed crosscoupling reaction of 2,5-diiodo-4-[(2-ethylhexyl)oxy]methoxybenzene (16) or 1,4-bis[(2-ethylhexyl) oxy]-2,5-diiodobenzene (10), 1,4-diethynyl-2,5bis-(octyloxy)benzene (11) and the tri-functional cross-linker 1,2,4tribromobenzene (17). R1=2-ethylhexyl, R2=n-octyl, R3=methyl (16) or 2ethylhexyl (10).

members of this family of conjugated polymers.7 In case of the new polymer systems described here, 1,2,4-tribromobenzene (17) was added to the reaction mixture as a crosslinker. The reactivity of aryl bromides towards cross-coupling is substantially lower than 147

that of aryl iodides, and it might be expected that the reaction of 16, 11, and 17 would allow linear PPE segments of an appreciable molecular length to grow, before crosslinking would lead to network formation and gelation of the reaction mixture. The polymerization reaction, as represented in Scheme 7.1, was conducted with varying monomer compositions under 'conventional' reaction conditions,6,8 i.e. under Pd0/CuI catalysis in a homogeneous toluene/diisopropylamine mixture at 70 C. In a first series of

Figure 7.1 Photograph of a gel of covalently cross-linked PPE network.

experiments (Reactions 1-5, see Experimental Section) the ratio of monomers 16:17 was varied from 0:1 to 10:1 and the amount of monomer 11 was adjusted to keep a stoichiometrically balanced molar ratio between aryl halide and aryl acetylene groups. The reaction mixtures were found to gel after a relatively short reaction time (ca. 1h), which is in line with the expected network formation (Figure 7.1). In accordance with typical procedures applied for the synthesis of linear PPEs,6,9 we elected to continue all reactions for a total of 24 h. After subsequent work-up, which involved purification by 148

rigorous extraction, polymers O-OPPE-X (the acronym represents cross-linked O-OPPE 18a) and MEHO-OPPE-X0.5 - MEHO-OPPE-X10.5 (the acronyms represent cross-linked MEHO-OPPE with the subscript after X denoting the ratio of monomer 16:17, 18b-18e respectively) were obtained in the form of rather large pieces. Consistent with the anticipated network structure, the products could not be dissolved, but were found to swell substantially (between ca. 300 and 600 % w/w) if immersed in good solvents for

12 10 Bromine Content (% w/w) 8 6 4 2 0

4 6 8 Ratio of Monomers 1/3

10

12

Figure 7.2 Residual content of bromine detected by elemental analysis in the crosslinked PPEs, as function of the ratio of monomers 16:17 ( = 18a, = 18b, = 18c, = 18d, = 18e).

MEHO-OPPE such as toluene or chloroform. Infrared (IR) spectra of the new polymers are consistent with the expected molecular structure, and elemental analysis data further confirm the expected composition. Importantly, however, the analytical data reveal the presence of remaining bromine groups (Figure 7.2). As can be seen from Figure 7.2, the amount of residual bromine is substantial for polymers synthesized from monomer mixtures comprising large fractions of cross-linker 17. However, it decreases with

149

increasing ratio of monomers 16:17. The absence of nitrogen in all samples indicates that the bromine present is covalently attached (presumably as aryl bromide) and is not due to residual diisopropylammonium bromide, which is formed during the reaction. Thus, the elemental analysis data are consistent with the interpretation that after reaching

8 Normalized Equilibrium Weight Increase

10

12

Ratio of Monomers 16/17

Figure 7.3 Normalized equilibrium weight increase for polymers 18a-18e swollen in toluene as function of the ratio of monomers 16:17 ( = 18a, = 18b, = 18c, = 18d, = 18e).

the critical extent of reaction, pc, at the gel point, the translational mobility of unreacted functional groups is severely limited, and the reaction essentially ceases after a certain cross-link density is reached. Of course, pc increases towards unity as the fraction of cross-linker decreases,10 and concomitantly the amount of residual bromine groups is reduced. At the same time, as expected, the equilibrium swelling weight decreases for the materials with a greater proportion of monomer 17 (Figure 7.3). This observation is consistent with the fact that higher cross-link densities result in lower equilibrium swelling ratios. 150

The conjugated polymer networks are highly luminescent if swollen with toluene but similar to linear PPEs6 their luminescence is considerably reduced in the dry solid state. The PL spectra of toluene-swollen cross-linked polymers 18a-18e are shown in Figure 7.4, together with a reference spectrum of a dry film of the linear reference

1.0

0.8 PL intensity (a.u.)

0.6

f c a b

0.4

0.2

e d 450 500 550 600 650 700

0.0

Wavelength (nm)

Figure 7.4 Photoluminescence spectra of the cross-linked PPEs investigated here (18a, open triangles, a; 18b, open circles, b; 18c, open squares, c; 18d, filled triangles, d; 18e, filled circles, e; all swollen in toluene) and a (dry) MEHOPPE reference film (filled squares, f). Spectra are scaled to optimally fit the graph.

polymer MEHO-OPPE (1).6 Polymers 18b-18d display very similar emission spectra that are characterized by rather broad spectral features and emission maxima around ca. 525 nm. The spectra match the one of the solid film of the reference polymer well, suggesting that the close proximity of individual chain segments and the limited translational mobility give rise to electronic properties that better match with the solid state of the reference polymer, rather than a molecular solution. On the other hand the emission 151

spectrum of polymer 18a, which features the highest molar fraction of cross-linker 17, is considerably broadened and it displays a significant bathochromic shift (emission maximum ca. 545 nm). This result seems to be linked to the high level of structural defects present in this material and suggests that energy migration to and emission from

Figure 7.5 Photograph of an image through a cross-linked film of MEHO-OPPE-X3.4, 18f.

low-bandgap states are important processes. We speculate that the residual aryl bromide functions, which are present in high concentration, play an important role in this context. Of course, the potential usefulness of the cross-linked polymers under investigation in actual devices depends on the ability to process these materials into thin films (and possibly other shapes). One approach to accomplish this objective follows the general framework routinely employed for standard thermo-set polymers, and is based on simultaneous polymerization and processing of the material into the desired shape. Thus, in an initial attempt to create cross-linked PPE films, the polymerization of monomers 16, 11, and 17 was carried out in a thin layer, which (for practical purposes) was deposited on

152

top of water. Indeed, this process resulted in a cross-linked coherent thin film (18f). Despite the fact that the thickness homogeneity of our initial samples was somewhat limited (600 300 m), they display reasonably good optical quality, can be seen from Figure 7.5.

7.4 Synthesis of Cross-linked Conjugated Polymer Milli-, Micro- and Nanoparticles While the systematic study addressing the influence of cross-linker concentration is reported in the previous section, a stoichiometrically balanced molar ratio of 6:9:2 (similar to 18d from the previous section) of monomers 16 or 10 , 11, and 17 was chosen for all reactions reported herein. As mentioned in the earlier section if the polymerization reaction shown in Scheme 7.1 was repeated, but the reaction was conducted in a vigorously stirred water/toluene/diisopropylamine mixture. Stabilized by the significant shear forces, the reaction mixture initially formed an "oil in water" emulsion, which comprised all monomers and the Pd-catalyst dissolved in the organic phase. The organic droplets rapidly displayed the characteristic color and photoluminescence (PL) emission characteristics of PPEs, and solidified over the course of the reaction, suggesting that the polymerization reaction proceeded smoothly. The final product was obtained in the form of well-separated, millimeter-sized, and essentially spherical particles (Figure 7.6a) and exhibited similar analytical data as the polymer 18d. While the size of the particles produced by this method is clearly too large for the formation of useful suspensions, the experiment unequivocally proved that the employed Pd-catalyzed cross-coupling polymerization indeed proceeds smoothly in an "oil in water" emulsion.11 In order to better stabilize the droplets and to reduce the particle size, an auxiliary surfactant was

153

Figure 7.6 Photographs (a), optical micrographs (b) and scanning electron micrographs (c) of cross-linked conjugated milli- (a) micro- (b), and nanoparticles (prepared in reaction nanometer-sized particles a) (c). Photographs and optical micrographs were taken in fluorescence mode under excitation at 366 nm and transmission/reflection mode, with the polymer particles dispersed in toluene.

introduced3 and in a next experiment sodium dodecyl sulfate (SDS) was added to the otherwise unchanged reaction mixture. It is evident from the optical micrographs shown in Figure 7.6b that this modification indeed led to a reduction of the particle size to the micrometer level. It should be pointed out that the polymer produced was isolated as a dry powder, but as demonstrated by the pictures shown in Figure 7.6b, the product could readily be re-dispersed into well-separated particles by ultrasonication in solvents such as toluene (without further surfactant addition). A detailed analysis of optical micrographs of re-dispersed, toluene-swollen micrometer-size particles show that their size distribution is relatively narrow with an average diameter of 4.7 m (Figure 7.7a). The chemical composition of the polymer was comparable to the one of the milli- and micrometer-size particles, and elemental analysis revealed that the SDS content in the final product is very low. In order to further reduce the average particle size, the

154

60

60

50

50

Number of Particles

Number of Particles

40

40

30

30

20

20

10

10

0
0 2 4 6 8 10 12 14 16 18 20

200

400

600

800

1000 1200 1400 1600 1800 2000

Diameter (m)

Diameter (nm)

Figure 7.7 Size-distribution of cross-linked PPE micro- (a) and nanoparticles (prepared in reaction nanometer-sized particles a) (b). The size of the individual particles was determined from optical transmission microscopy (a) and scanning electron microscopy (b) images. The particles evaluated in (a) were swollen with toluene.

polymerization reaction was conducted under emulsion conditions as outlined above, but an ultrasonic bath was employed instead of a mechanical stirrer. The initial yield of the isolated product was limited (4.4% ; prepared in reaction nanometer-sized particles a) and calls for further improvement of the protocol, scanning electron microscopy pictures of polymer that was re-dispersed by ultrasonication in toluene confirm that cross-linked nanospheres with a diameter between ca. 50 and 400 nm and a narrow size distribution can be produced by this method (Figures 7.6c, 7.7b). To improve the yield of the emulsion reaction yielding cross-linked nanospheres the polymerization reaction was conducted with twice the original amount of the employed aqueous phase (reaction nanometer-sized particles b). The surfactant concentration for the reaction was kept 155

similar to that used in earlier reactions and indeed the yield of the reaction could be improved by an order of magnitude to 48.9%. It was also observed that upon increasing the surfactant concentration (reaction nanometer-sized particles c) the resulting size of the nanosphere obtained could be further decreased (Figure 7.8). Employing similar reaction conditions cross-linked nanoparticles of EHO-OPPE with higher degree of crosslinking were also synthesized.

Figure 7.8 SEM image of cross-linked EHO-OPPE nanoparticles.

As is evident from the optical fluorescence micrographs shown in Figure 7.6, the conjugated polymer networks are highly photoluminescent (PL) if swollen with toluene, but similar to linear PPEs6 the luminescence is strongly quenched in the dry solid state. The PL emission spectra of toluene-swollen milli, micro- and nanoparticles are shown in Figure 7.9, together with a reference spectrum of MEHO-OPPE (a linear PPE based on monomers 16 and 116) in toluene solution. Gratifyingly, the materials display very similar emission spectra, which feature well-resolved phonon bands. As

156

photoluminescence spectra are in general very sensitive to structural defects and impurities, this result seems to suggest the absence of structural defects and impurities

PL Intensity (a. u.)


400

450

500

550

600

650

700

Wavelength (nm)

Figure 7.9 Photoluminescence spectra of PPE milli (dashed line), micro- (dotted line) and nanoparticles (dash-dotted line), suspended in toluene, and as a reference an MEHO-OPPE solution in toluene (solid line).

and indicates that neither the network structure of the cross-linked particles, nor the remaining bromine functions significantly disturb the principal optoelectronic properties of the conjugated polymers.

7.5 Conclusions In conclusion we have shown that covalently cross-linked poly(p-phenylene ethynylene)s can readily be produced by introduction of trifunctional conjugated crosslinkers such as 1,2,4-tribromobenzene into an otherwise conventional palladiumcatalyzed cross-coupling reaction of difunctional aryl halides and aryl acetylenes. The

157

content of the cross-linker was systematically varied, resulting in materials with different cross-link densities. Particularly if swollen with organic solvents, the new polymers are highly photoluminescent. Large mole fractions of the cross-linker lead to considerable contents of residual aryl bromide functions, which appear to act as low-energy relaxation sites and cause a bathochromic shift and broadening of the emission band. Addressing the processing issues associated with these novel cross-linked polymer systems, initial experiments demonstrated that films can be produced by simultaneous polymerization and processing of the material into a desired shape. We have also demonstrated that covalently cross-linked spherical conjugated polymer particles can readily be produced by introduction of adequate cross-linkers and conducting cross-coupling reactions in aqueous emulsions instead of homogeneous solutions. We have demonstrated that the size of the polymer particles can easily be tuned over a wide range (nm to mm) by modification of the reaction conditions. The optoelectronic properties of the materials are similar to those of the linear reference polymer synthesized under conventional conditions, and confirm the absence of electronic defects.

158

7.6 References 1 (a) Kumar, A; Reynolds, J. R.Macromolecules 1996, 29, 7629. (b) Lavastre, O.; Cabioch, S.; Dixneuf, P. H.; Sedlacek, S.; Vohlidal, J. Macromolecules 1999, 32, 4477; (c) Joo, J.; Lee, J. K.; Lee, S. Y.; Jang, K. S.; Oh, E. J.; Epstein, A. J. Macromolecules 2000, 33, 5131. 2 Groenendaal, B. L.; Jonas, F.; Freitrag, D.; Pielartzik, H.; Reynolds, J. R. Adv. Mater. 2000, 12, 481. 3 Landfester, K.; Montenegro, R.; Scherf, U; Gntner, R.; Aswapirom, U.; Patil, S.; Neher, D.; Kietzke, T. Adv. Mater. 2002, 14, 651. 4 Marie, E.; Rothe, R.; Antonietti, M.; Landfester, K. Macromolecules. 2003, 36, 3967. 5 For a review on palladium(0) catalyzed reactions in homogeneous aqueous medium see: Genet, J. P.; Savinac, M. J. Organomet. Chem. 1999, 576, 305. 6 (a) Weder, C.; Wrighton, M. S. Macromolecules 1996, 29, 5157. (b) Dellsperger, S.; Dtz, F.; Smith, P.; Weder, C. Macromol. Chem. Phys. 2000, 201, 192. 7 (a) Weder, C. Ed. Poly(arylene ethynylene)s From Synthesis to Applications; Advances in Polymer Science Series Vol. 177; Springer, Heidelberg, 2005. (b) Bunz, U. H. F. Chem. Rev. 2000, 100, 1605. 8 Sonogashira K. in Metal-Catalyzed Cross-Coupling Reactions (Diederich, F.; Stang, P. J . Eds.), Wiley-VCH, Weinheim 1998, p. 203. 9 (a) Giesa, R.; Schulz, R. C. Macromol. Chem. 1990, 191, 857. (b) Swager, T. M.; Gil, C. J.; Wrighton, M. S. J. Phys. Chem. 1995, 99, 4886. (c) Egbe, D. A. M.; Klemm, E. Macromol. Chem. Phys. 1998, 199, 2683.

159

10

Odian, G. Principles of Polymerization, 3rd Edition, John Wiley & Sons Inc., New York 1991, p. 203.

11

We have also successfully synthesized a linear PPE based on monomers 11 and 16 under similar conditions.

160

Chapter 8: Conclusions and Outlook The experimental investigation presented in this thesis was focused in the area of conjugated polymer networks. These are conjugated polymer networks that can be either covalently or non-covalently crosslinked. In this orthogonal approach we postulated that the structural feature of conjugated cross-links would improve the electronic properties of these systems. In Chapter 3, we have presented the in-depth investigation of the charge transport characteristics of -conjugated semiconducting polymer EHO-OPPE. Before this study was undertaken the charge transport behaviour in poly(p-phenylene ethynylene)s was mainly unknown and these polymers were for the most part considered extremely useful for their optical properties. Using time-of-flight technique it was found that EHO-OPPE exhibits high ambipolar mobility (~210-3 cm2V-1s-1). A detailed study on the charge transport behaviour in this polymer was conducted, and we were able to calculate the positional (off-diagonal) and energetic (diagonal) disorder parameters from the experimental temperature and field dependences of the hole and electron mobilities. An interesting materials property of negative field dependence of the mobility was observed, which was explained within the Gaussian disorder formalism to originate from high offdiagonal disorder. The experimental temperature and field dependent mobilities were consistent with a Gaussian disorder transport formalism applied to systems exhibiting both positional and energetic disorders. Charge transport layers are employed in conjugated polymer based devices (e.g. light-emitting diodes, LEDs etc.) to facilitate the transport of a desired electronic species. The good ambipolar charge transport observed in EHO-OPPE, is promising and find use various applications in the the field of plastic

161

electronics. The oxidative stablilty of these materials is relatively high which thus introduces a challenge for doping these systems. Thus, design and synthesis of PPEs with incorporated moieties that are easily oxidized or reduced, holds the promise of a new class of synthetic metals. In order to investigate the influence of presumably non-conjugated cross-links on charge transport characteristics of EHO-OPPE, organometallic polymer networks were synthesized via ligand-exchange reactions as discussed in Chapter 4 by employing EHOOPPE on one hand and the bridged low-molecular-weight organometallic complex [Pt(-Cl)Cl(PhCH=CH2)]2 on the other. Importantly these organometallic networks can be processed easily, using conventional drop casting technique into thick films. The films thus produces were insoluble in common organic solvents, however were found to dissolve in styrene, thus suggesting the formation crosslinked polymer networks. The charge transport properties of the synthesized EHO-OPPE-PtII were investigated using the time-of-flight technique. The transient time (ttr) extracted from the obtained photocurrent transients, however did not scale with the film thickness, suggesting that the photocurrent transients were range limited. Large scattering was also observed in the calculated charge carrier mobilities, thus signifying that the PtII incorporated in the EHOOPPE leads to significant trapping of both electrons and holes. For investigating the effect of conjugated cross-links on the charge transport in EHO-OPPE, we decided to employ a zerovalent metal as the cross-link in-between the conjugated polymer chains. En route to our goal, we also demonstrated two new and convenient routes for the synthesis of bis(2-alkyne)-platinum(0). Both methods developed depart from readily accessible, chemically stable intermediates, are based on

162

straight forward and undemanding reaction conditions, and afford the desired product in high purity. Extending the synthetic protocols developed for the model compounds, we have also demonstrated that organometallic conjugated polymer networks can be synthesized and processed by ligand exchange between a linear conjugated polymer and a labile metal complex. Gratifyingly, we observed that the introduction of zerovalent metal and thus conjugated cross-links between EHO-OPPE leads to substantial increase of the ambipolar charge carrier mobility (1.610-2 cm2V-1s-1 for electrons and 1.410-2 cm2V-1s-1 for holes). The charge carrier mobilities of the dicussed EHO-OPPE-Pt0 networks represent the highest mobilities yet observed in disordered conjugated polymers and also compare well to the hole mobilities of ordered materials (10-1 cm2V-1s-1). The ease of processing and the ambipolar characteristics of the new materials are particularly intriguing and may lead to a new generation of high-performance semiconducting devices. The reduction of the conjugated length of EHO-OPPE upon complexation with Pt0 might also limit the observed improvement in the charge carrier mobilities. In order to overcome this challenge, incorporation of the ethynyl moiety as a side chain, would help in retaining the conjugation length of the polymer backbone and thus promises improved charge carrier transport. Auxiliary ligands can be incorporated in the polymer main chain to gain access to a plethora of useful metal centers as cross-links. With this approach in mind we have shown in Chapter 6 at the example of poly(p-phenylene ethynylene) derivatives that the introduction of 2,2-bipyridine moieties leads to conjugated polymers, which are formidable precursors for metallo-supramolecular PPE networks that are accessible via ligand-exchange reactions. The resulting three-dimensional cross-linked, conjugated

163

BipyPPE1-metal and BipyPPE2-metal networks display interesting optoelectronic properties. As expected, the coordination of the metal markedly influences the photophysical characteristics of the polymer. It is obvious that the approach is broadly applicable to other conjugated polymer platforms and also to other ligands. The (pseudo)tetrahedral coordination of four-coordinated metals (as in case of some of the complexes explored here) might ultimately not represent the most desirable geometry from a chargetransfer point of view. The study discussed in Chapter 4 as well suggests that the ionic nature of the metal centers used here might lead to charge trapping. The counter-ions on the other hand might also lead increased ionic transport in the films resulting in severe dark currents. After having employed these complexes in the present study to demonstrate the possibility to prepare a new family of -conjugated organometallic polymer hybrid systems, we envision to extend the present concepts to the use of square planar complexes and conduct experiments related to the charge-carrier transport in such systems. One of the organometallic complex that can be employed is

bis(cyclooctadiene)Ni(0). This complex however is highly air sensitive and thus the film processing needs to be carried out in the absence of oxygen and moisture. Focusing on another design principle for obtaining conjugated polymer networks, we have demonstrated in Chapter 7 that covalently cross-linked poly(p-phenylene ethynylene)s can readily be produced by introduction of trifunctional conjugated crosslinkers such as 1,2,4-tribromobenzene into an otherwise conventional SonogashiraHagihara reaction. The number of equivalents of the cross-linker was systematically varied, resulting in materials with different cross-link densities. Particularly if swollen with organic solvents, the new polymers are highly photoluminescent. Large mole

164

fractions of the cross-linker lead to considerable contents of residual aryl bromide functions, which appear to act as low-energy relaxation sites and cause a bathochromic shift and broadening of the emission band. Addressing the processing issues associated with these novel cross-linked polymer systems, initial experiments demonstrated that films can be produced by simultaneous polymerization and processing of the material into a desired shape. For the first time, we have also demonstrated that covalently crosslinked spherical conjugated polymer particles can readily be produced by introduction of adequate cross-linkers and conducting cross-coupling reactions in aqueous emulsions instead of homogeneous solutions. Gratifyingly, the size of the polymer particles can easily be tuned over a wide range (nm to mm) by modification of the reaction conditions. The optoelectronic properties of the materials are similar to those of the linear reference polymer synthesized under conventional conditions, and confirm the absence of

10 Mobility (x10 cm /Vs) 8 6 4 2 0 0 10 20 30 4 Field (10 V/cm) 40

Figure 8.1 Electron (circles) and hole (squares) mobilities for covalently cross-linked nano-particles (9 % mol cross-linker, filled symbols) and EHO-OPPE (open symbols).

-3

165

electronic defects. Promising preliminary results have been obtained regarding the charge carrier mobility in these covalently cross-linked conjugated polymer networks (Figure 8.1), and a detailed study is underway. The design and successful synthesis of these materials as (easily tuned) processable conjugated polymer network particles can be extended for other conjugated polymer platforms and opens the door for the design and synthesis of new and novel network architectures, which promise a new generation of useful high charge carrier mobility materials.

166

Acknowledgements Many individuals have contributed to the present thesis and have made the past years an enriching and enjoyable experience. First of all I would like to thank Dr. Kenneth D. Singer, from Case Physics Department, for his help, guidance and various simulating discussions that have resulted in fruitful culmination of many joint projects and for being one of the co-examiners on my thesis defense commitee. I would also like to express my gratitude to Dr. Irina Shiyanovskaya, who taught me many important skills in the physics laboratory. I would not miss thanking Dr. Walter Caseri from ETH, Zrich for simulating discussions and successful collaborations. I also gratefully acknowledge Dr. Stuart Rowan and Dr. David Schiraldi for their help, various interesting discussions, pleasant times and for being co-examiners on my thesis defense committee. I would especially like to thank people from the Weder group for all their help, great times and friendship (listed in alphabetical order): Elisa Beekman, Mark Burnworth, Brent Crenshaw, Simon Dellsperger, Susan Given, Eric Hittinger, Christian Huber, Dr. Param Iyer, Maki Kinami, Christoph Kocher, Dr. Daniel Knapton, Jill Kunzelman, Dr. Christiane Lwe, James Mendez, Gallia Painter, Claire Rademaker, Dr. Michael Schrrs, Katarina Sigg, Kara Smith, Ravisubhash Tangirala, Dale Wilger, Elliot Yamaguchi, Peter Yao and Sven Zimmerman. During the time of the present thesis I have met many people to whom I am grateful for all the interesting and memorable times (listed in alphabetical order): Ravi Ayyer, Suneel Bandi, Ben Beck, Marco Camesasca, Casey Check, Maddalena Fanelli, Sergio Granados-Focil, Prasad Gopalkrishnan, Shakambhari Gupte, Subramanian

167

Prakash Iyer, Akshay Kamdar, Ashish Kulkarni, Shiwangi Kulkarni, Angelique Lundberg, Sharon Ohba, Amol Patankar, Abhiraj Purandare, Aditya Ranade, Sona Sivakova and Alberto Scurati. I am deeply indebted to my parents, my sister and my family for their love and support for the past 26 years. Last but not the least I would like to thank Dr. Christoph Weder for giving me the opportunity to work with him. This work would not have been possible without his able guidance. I have learned not only a great number of concepts concerning the present thesis and science in general from him, but also a great deal about life. Vielen Dank Wedi!

168

List of Publications Based on this Thesis Huber, C.; Kokil, A.; Caseri, W.; Weder, C. Two alternative, convenient routes to bis(diphenylacetylene)platinum(0); Organometallics 2002, 21, 3817. Kokil, A.; Shiyanovskaya, I.; Singer, K.D.; Weder, C. Charge Transport in -Conjugated Organometallic Polymer Networks; J. Am. Chem. Soc. 2002, 124, 9978. Kokil, A.; Huber, C.; Caseri, W.; Weder, C. Synthesis of -Conjugated Organometallic Polymer Networks; Macromol. Chem. Phys. 2003, 204, 40. Kokil, A.; Shiyanovskaya, I.; Singer, K.D.; Weder, C. Charge Carrier Transport in Poly(2,5-dialkoxy-p-phenylene ethynylene)s; Synth. Met. 2003, 138, 513. Hittinger, E.; Kokil, A.; Weder, C. Synthesis and Characterization of Cross-Linked Poly(p-phenylene ethynylene)s; Macromol. Rapid Commun. 2004, 25, 710. Hittinger, E.; Kokil, A.; Weder, C. Synthesis and Characterization of Cross-Linked Conjugated Polymer Milli-, Micro-, and Nanoparticles; Angew. Chem. Int. Ed. 2004, 43, 1807. Kokil, A.; Yao, P.; Weder, C. Organometallic Networks Based on 2,2-BipyridineContaining Poly(p-phenylene ethynylene)s Macromolecules 2005, 38, 3800.

169

Vous aimerez peut-être aussi