Mihaela Dascalu*,
Valentina E. Musteata,
Loredana Vacareanu,
Carmen Racles and
Maria Cazacu
“Petru Poni” Institute of Macromolecular Chemistry, Aleea Gr. Ghica Voda 41 A, Iasi 700487, Romania
First published on 13th November 2015
Three new poly(siloxane-urethane) crosslinked structures (PSFUs) were prepared from 1,1′-diisocyanatoferrocene (Fc(NCO)2), and two siloxane diols in different proportions, i.e. α,ω-bis(hydroxybutyl)oligodimethylsiloxane (HS-1) and a hybrid diol containing hydrolysable triethoxy (–Si(–O–Et)3) groups (HS-2). The formation of the urethane groups along the polymer backbone was first confirmed by the presence of the specific bands in FTIR spectra. The resulting materials were characterized using differential scanning calorimetry (DSC), thermogravimetric analysis (TGA), scanning electron microscopy (SEM), dielectric spectroscopy (DE) and X-ray photoelectron spectroscopy (XPS). Cyclic voltammograms of the compounds showed quasi-reversible oxidation–reduction waves making them suitable for sensing applications.
The great development of organometallic polymers field resulted from the fact that the incorporation of transition metals into polymeric structures allows access to special materials with unusual and attractive characteristics. Mainly ferrocenyl-based polymers and copolymers are of high importance due to their redox, electrical, conducting/semiconducting, optical, magnetic, catalytic, preceramic and elastomeric properties.6–8 The broad area of application of the ferrocene polymers includes the manufacture of electronic devices such as microelectrochemical diodes,9 formation of redox gels, which show charge transfer properties,10 modification of electrodes,11 construction of amperometric biosensors,12 precursors to ferromagnetic ceramics,13,14 and more recently, applications in the area of non-linear optical (NLO) materials.15–18
In general, the polyurethanes combine rigid hard and flexible soft segments and, depending on the types and compositions of the two phases as well as the preparation procedures, the relationships between structure and properties of polyurethanes are extremely diverse and could be easily controlled.19–22 The hard segment is formed from diisocyanate moieties, while the soft segment results from the chain extenders that generally are flexible polyols23,24 which contribute to flexibility and elasticity. The number of the diols suitable and used for polyurethanes building is extremely larger mostly being organic ones. However, siloxane diols were also used as reagent for diisocyanate and polyurethane extender.24 The combination between polyurethanes and polysiloxanes, both well known as versatile multipurpose materials has attracted special attention due to a series of advantages: a good thermal stability and high flexibility at low temperatures given by the siloxane part and better mechanical strength and abrasion characteristics contributed by polyurethane.25–27
To the best of our knowledge co-existence of ferrocene and siloxane in a polyurethane structure has not been reported to this date, even though, as mentioned above, different combinations of two out of the three main components have been intensively studied. It was the main purpose of our research to obtain combinations of the three main moieties in cross-linked materials and to investigate their properties. Especially the electric properties and electro-chemical behaviour are expected to modify as a result of chemically linked ferrocene. Here we present the synthesis and characterization of new materials based on three main structural elements: ferrocene units and siloxane segments linked by urethane groups. Different from our previous paper,28 where a pure organic monomer was used as urethane precursor, namely isophorone diisocyanate, here we used a less common monomer, 1,1′-diisocyanatoferrocene,29 which permitted the introduction of a metallic element in the form of complexed iron units within the newly formed backbone. Moreover, the resulting poly(siloxane-ferrocene urethane) structures (PSUFs) are crosslinked since a siloxane diol carrying triethoxy groups is used as co-reagent together with an α,ω-bis(hydroxybutyl)oligodimethylsiloxane. The morphology, thermal and dielectric properties of the PSUFs have been investigated. The compounds were analyzed by cyclic voltammetry to evaluate the electrochemical behaviour. In order to determine the precise composition and the oxidation state of the elements present in the resulted materials X-ray photoelectron spectroscopy was used.
The silica domains formed can act as covalent crosslinking points to provide dimensional stability but also act as reinforcement “inorganic filler”. Based on the feeding stoichiometry, the length of the prepolymer differs in PSUF-1 and PSUF-3 (5.25 structural units, theoretical Mn = 5292.11 g mol−1) compared to PSUF-2 (2.77 structural units, theoretical Mn = 2792.22 g mol−1). The lengths of the prepolymers were calculated according to Carothers formula (1) for the case when one of the monomers was used in excess.
![]() | (1) |
The GPC results on the two prepolymers were in good agreement with the theoretical calculations: Mn,PSUF-3 = 5149 g mol−1, PDI = 1.751; Mn,PSUF-2 = 3124 g mol−1, PDI = 1.801.
Due to the oligomeric nature of HS-1, it acts as the softest part of the poly(siloxane-urethanes), while the length of the prepolymer triggers the chain microstructure, with longer “soft” segments (blocks) in PSUF-1 and 3 and a more “equilibrated” composition in PSUF-2. In addition, as mentioned before, PSUF-3 is reinforced with an organic–inorganic filler generated in situ from the HS-2 used in excess; the presence of such particles is not completely excluded in the other samples.
The FTIR spectra of the PSUFs (Fig. S1-ESI†) showed the disappearance of the band assigned to the isocyanate group at 2295 cm−1,31,32 and the presence of the bands typical for polyurethane: CO in urethane group at 1707 cm−1 and –NH (free and bonded) around 3300 cm−1 which is overlapped by the broad band from 3437 cm−1 assigned to OH groups. The hydrolysis of the Si–O–Et groups and their condensation is sustained by the disappearance of the band at 1092 cm−1 assigned to Si–O–C bonds from HS-2 and the appearance of a band around 1198 cm−1 which corresponds to Si–O–Si stretching of crosslinked organo-silica structures resulted by self-condensation of Si-OH groups (formed by hydrolysis of ethoxy groups).33 The band at about 3437 cm−1 and the shoulder at 953 cm−1 are associated with Si-OH groups un-condensed in the described conditions. In the FTIR spectra of the PSUFs (Fig. S1-ESI†), the bands characteristic for the dimethylsiloxane sequences at 1261 and 800 cm−1 are also present. The Si–O–Si band visible in the spectrum of HS-1 at 1020–1080 cm−1 is now overlapped with silica Si–O–Si band and with C–O–C band from the urethane group. The ferrocene sp2 C–H stretching is present at around 3088 cm−1 while asymmetric and symmetric sp3 C–H stretching are assigned at 2963 and 2932 cm−1, respectively.34 The other bands from ferrocene at 1020–1098 cm−1 are overlapped with the Si–O–Si band. The ferrocene Cp–Fe stretching appears at 478 cm−1 in the starting diisocyanate.35 In the PSUFs, a wide band centered at around 460 cm−1 is present. Since in the same region the silica network absorption may overlap,36 we compared the infrared spectra of PSUFs with a previously synthesized poly(siloxane-urethane)28 with similar structure but without ferrocene (Fig. S2-ESI, PU2†). We identified the band for silica at 473 cm−1, which appears as a shoulder in PSUFs, while the band at 460 cm−1 was thus assigned to Cp–Fe bonds in the PSUFs.
The results of the thermogravimetric analysis of the PSUFs in the 20–750 °C temperature range, in nitrogen, are shown in Fig. S3-ESI.† The slight weight loss below 250 °C can be assigned to traces of solvents remained after the drying process (DMF especially) or resulted in the condensation of residual silanol groups (water), plus the eventually adsorbed humidity. The weight loss in this step increases with the increasing amount of HS-2 used in the reaction. The main decomposition stage occurs in the 250–400 °C range. The residue amount at 700 °C increases in the order PSUF-3 < PSUF-2 < PSUF-1.
The introduction of siloxane segments into polyurethane structures creates thermodynamically immiscible systems and contributes to the microphase separation of rigid and flexible segments. In our case, the phase separation that exists in poly(siloxane-urethane) crosslinked structures PSUF-1 and 3 is evidenced in the DSC curves (Fig. S4 and S5-ESI†) in which two glass transitions (Tg) are identified: one at around −40 °C (Table 1) which was assigned to the prepolymer containing the oligosiloxane component (“soft” segment) and the second one in the positive temperature range, which represents the relaxation of “hard” segments containing HS-2 disiloxane with network nodes. Both Tg values are higher in PSUF-3 than in PSUF-1 due to the excess HS-2, forming the organo-silica network, which acts as reinforcing filler. A single glass transition was detected for PSUF-2, which has a more balanced microstructure, with smaller prepolymer chain. One possible explanation could be that in PSUF-2 the two main components are better mixed (alternated) and the prepolymer chains do not organize in separate domains. The melting of the siloxane–ferrocenyl–urethane crystalline domains takes place in the 58–66 °C temperature range for all three samples, rising in line with the HS-2 content of the main chain. The crystallization temperatures are in the same order, with a more pronounced supercooling effect in PSUF-1 compared with the PSUF-3. This difference between PSUF-1 and PSUF-3 could be the result of the presence of silica in PSUF-3, which limits the mobility of the macromolecular chains.
Sample | Tg (°C) (2nd heating) | Tm (°C) (1st heating) | Tc (°C) | |
---|---|---|---|---|
PSUF-1 | −43.67 | 9.85 | 57.73 | 16.59 |
PSUF-2 | — | 10.07 | 66.23 | 58.13 |
PSUF-3 | −38.97 | 20.78 | 60.11 | 39.11 |
SEM micrographs of the cross-section of the prepared samples processed as films are displayed in Fig. 1. Submicronic domains (roughly between 200 and 1000 nm) are observed in all samples. Sample PSUF-2 has the smoothest aspect in cross-section, with very small regular domains and only few larger particles. This aspect correlates with the presence of a single Tg in DSC. The increased HS-2 feeding amount in PSUF-3 leads to formation of large formations, most probably organo-silica particles agglomerated into the network nodes. The EDX analyses proved the presence of all expected elements (C, O, Si, N and Fe).
The presence of the iron in the materials is expected to induce, among others, electrical or conducting/semiconducting properties, so dielectric measurements were carried by sweeping the frequency from 1 Hz to 1 MHz at constant temperatures taken at 5 °C intervals from −140 °C up to 80 °C, which resulted in an effective heating rate of about 1 °C min−1. In order to have an image of dipolar and charge transport processes, the results were discussed as complex dielectric permittivity ε* = ε′ − iε′′ (where ε′ is the storage component and ε′′ is the loss factor which includes the effects of both dielectric loss and conductivity) as well as the calculated ac conductivity:37 σac = ε0ωε′′, where ε0 is the permittivity of free space and ω = 2πf is the angular frequency. The results revealed that the presence of ferrocene units leads to an increase of dielectric constant from ∼2.8 up to 104,38 when the temperature increases in the range −140–80 °C. The curves evidence very well the frequency dependence of dielectric constant. To evidence the effect of temperature on the complex dielectric permittivity for the PSUFs, the isothermal data are re-plotted in Fig. 2 as a function of temperature at several frequencies. At temperature bellow −50 °C, the dependence of dielectric parameters, ε′ and ε′′, highlights regions which are associated to the mobility of the macromolecular chains. ε′′ presents a small amplitude peak shifting to higher temperatures with increasing frequency, which corresponds to the dipolar relaxation associated with the glass transition of amorphous dimethylsiloxane oligomer. This relaxation has been previously evidenced for siloxane polymers.38 With increasing temperature above −40 °C there is an increase with several orders of magnitude both for ε′ and ε′′, especially for PSUF-2 and 3. The steep increase of ε′ at lower frequencies is caused by increased mobility of charge carriers which leads to space charge polarization phenomena.39 The high value of dielectric loss in the same temperature range is also caused by charge migration that determines the dissipation of the field energy in the material. The frequency dependent shoulder (at about −20 and respectively, 0 °C for PSUF-2 and 3) may come from interfacial polarization when charges are trapped within the interfaces of a material, for example between separated phases, amorphous and crystalline regions or at the external electrode–sample interface.40
![]() | ||
Fig. 2 Dielectric permittivity and dielectric loss versus temperature at different frequency (1–10 MHz), for PSUFs. |
The frequency dependence of conductivity at several temperatures is shown in Fig. 3. At low temperatures conductivity has close values for all the samples, with a linear dependence on frequency, in double logarithmic scale. The ac conductivity dispersion with increasing frequency is common in heterogeneous and disordered solids41 and corresponds to short range jumps of charges. On increasing temperature, there is a plateau at low frequencies due to charges jumps over long distances which permit electric conduction throughout the volume of the sample.42 The observed plateaus corresponding to dc conductivity σdc at 25 °C are of the order of 10−10 S cm−1 for PUSF-3 and respectively, 10−7 S cm−1 for sample PUSF-2.
For PSUF-1, the frequency independent region appears only at temperatures above 60 °C, and corresponds to lower values of dc conductivity, with σdc ∼ 10−10 S cm−1. At room temperature, the conductivity of PSUFs is several orders of magnitude higher compared to that of polysiloxanes or polyurethanes (∼10−15 to 10−14 S cm−1).43
The Fc(NCO)2 and the PSUFs films were analysed using X-ray Photoelectron Spectroscopy. The Fe high resolution spectra of Fc(NCO)2 revealed a main peak at 708 eV characteristic for Fe(II) in ferrocene and one peak at 709.3 eV which could be assigned to Fe(III) derived from ferrocene (in literature at 709.9 eV).44 This spontaneous oxidation is normal and could occur during manipulation of the sample. In the PSUFs spectra, the region characteristic for Fe exhibited four peaks (two pairs of peaks). The main maximum was detected at approximately 710 eV (PSUF-3 – 710.5 eV, PSUF-2 – 710.1 eV, PSUF-1 – 709.8 eV) with a satellite at approximately 723 eV which was assigned to Fe(III) in ferrocenium moieties. The smaller maximum at approximately 715 eV (715.2 eV-PSUF-3, 714.5 eV-PSUF-2, 715.3 eV-PSUF-1 (Fig. S6†)), with satellites at around 728 eV could be assigned to the presence of iron oxides. According to the literature, for Fe(III) in iron oxides the binding energy is around 711 eV with a satellite around 719 eV.45 The O 1s binding energy in iron oxides was reported to be “nearly invariant at 530.1 eV”.46 In PSUFs, besides the O 1s main peak at approximately 532 eV, there is a small peak at approximately 530 eV (Fig. S6-ESI†). The amount of iron oxides in the three samples could be estimated from the XPS spectra. For example, in the PSUF-3 sample, based on the atomic concentration of O (16.8%) and Fe (3.24%) from the XPS data (not shown), it results that ca. 1.2% O atoms are in iron oxide. Supposing that Fe2O3 is formed, this would mean that ca. 25% from the total ferrocene moieties transformed into iron oxide. The formation of iron oxides is not surprising, given the experimental conditions for the preparation of the crosslinked films, i.e. thermal treatment and the use of HCl.
The electrochemical characteristics of PSUFs were investigated by cyclic voltammetry (CV) in two ways, as free standing film immersed in chloroform in case of PSUF-3 sample, and as chloroform solutions for PSUF-1, PSUF-2 and Fc(NCO)2. All the chloroform solutions contain 0.1 M TBABF4 as supporting electrolyte. Partial oxidation of Fc(NCO)2 due to the manipulation and processing of the sample in normal atmosphere can be the cause for the presence of multiple redox couples in CVs curves, assigned to the Fe(II) as ferrocene and Fe(III) as ferrocenium units (Fig. S7-ESI†). As previously mentioned, the obtained products are slightly flexible crosslinked films, so the preliminary results of the CV measurements that were done for PSUF-3 film immersed in chloroform solution were not satisfactory. This is the reason why the electrochemical measurements for sample PSUF-1 and PSUF-2 where done in solution before they underwent thermal treatment that leads to crosslinked materials and formation of iron oxides. The potential value applied to the working electrode was swept in a large window potential (−1.0 V to 1.0 V), with a 50 mV s−1 scan rate. The cyclic voltammograms recorded for all three samples are shown in Fig. 4.
![]() | ||
Fig. 4 CVs recorded for (a) PSUF-1 and PSUF-2 and (b) PSUF-3 as film in chloroform solutions containing 0.1 M TBABF4 as supporting electrolyte; scan rate of 50 mV s−1. |
The recorded CV curves for PSUF-1 and PSUF-2 indicate that the samples undergo one redox process and their electrochemical data are listed in Table 2. The voltammograms indicate one reversible redox process with anodic peaks potential, E (V) of 0.409 V for PSUF-1, and 0.210 V for PSUF-2. The electrons removal in the oxidation step is assumed to ferrocene units and these values of ΔE are 5 and 2 times higher, respectively, compared to ferrocene (Fc) single molecule (ΔE = 0.083 V). The synthesized polymers (PSUF-1 and PSUF-2) show good reversibility of redox processes having half-wave potential at E1/2 = 0.201 V, and 0.125 V vs. Ag/AgCl, respectively. The obtained values are close to those obtained for the reversible redox behaviour of ferrocene-modified polysiloxane (E1/2 = 0.260 V).47,48
E/V | Sample | ||||
---|---|---|---|---|---|
Fc | Fc(NCO)2 | PSUF-1 | PSUF-2 | PSUF-3 | |
a Epa – anodic peak potential; Epc – cathodic peak potential, ΔE – the peak separation (ΔE = Epa − Epc), E1/2 – the half-wave potential (E1/2 = (Epa + Epc)/2). | |||||
Epa1 | 0.556 | 0.318 | 0.409 | 0.210 | 0.546 |
Epc1 | 0.473 | 0.224 | −0.007 | 0.040 | — |
ΔE1 | 0.083 | 0.094 | 0.416 | 0.170 | — |
E11/2 | 0.525 | 0.271 | 0.201 | 0.125 | — |
Epa2 | — | 0.530 | — | — | — |
Epc2 | — | 0.462 | — | — | — |
ΔE2 | — | 0.068 | — | — | — |
E21/2 | — | 0.496 | — | — | — |
Epa3 | — | 0.800 | — | — | — |
Epc3 | — | 0.676 | — | — | — |
ΔE3 | — | 0.124 | — | — | — |
E31/2 | — | 0.738 | — | — | — |
For PSUF-3 material, the potential was swept from −1.5 V to 1.5 V, and from the voltammogram it can be observed the appearance of only one large and not well-defined oxidation peak (Epa = 0.546 V). Broad shape of the redox peaks can be explained by the slow electron transfer, which is limited by the slow mass transfer of ions through the film upon redox reactions. The current intensity of the cathodic peak is lower than for the anodic peak (and thus difficult to assign precisely) which can be attributed to the slight enhancement of hydrophilic character by the ferrocenium ion state produced upon oxidation.
Materials with fixed redox site on polymer backbone may be used for sensing applications49 or as electrostatic dissipative polymers.50 The route of the redox charge transport may be controlled by either the electron transport within the conducting chain or the electron hopping between redox center and polymer chain.51 From a practical point of view, we recall that a redox group may act as a switching molecular system for charge transport so that interesting applications of redox polymers to electronic devices are open.52 Based on these results, we assume that the un-crosslinked poly(siloxane-urethane)s with ferrocene moieties could be better candidates for sensing elements than the crosslinked materials. On the other hand, a crosslinking reaction in milder conditions could avoid the formation of iron oxides and thus ensure better defined materials for applications.
The hybrid diol HS-2 was prepared following a previously reported procedure28 starting from 1,3-bis(3-glycidoxypropyl)tetramethyldisiloxane and 3-aminopropyltriethoxysilane in DMF solution using DBTDL as catalyst. Finally a viscous transparent fluid was obtained.
Sample | Reactants | ||
---|---|---|---|
HS-1, mmol | HS-2, mmol | Fc(NCO)2, mmol | |
PSUF-1 | 1.35 | 0.60 | 1.87 |
PSUF-2 | 1.01 | 1.00 | 1.87 |
PSUF-3 | 1.35 | 1.60 | 1.87 |
Footnote |
† Electronic supplementary information (ESI) available: FTIR spectra, TGA and DSC curves, XPS spectra and CV curves of poly(siloxane-urethane) crosslinked structures (PSFUs). See DOI: 10.1039/c5ra15290a |
This journal is © The Royal Society of Chemistry 2015 |