Jaber Nasrollah Gavgania,
Hossein Adelniab,
Davood Zaarei*c and
Mohsen Moazzami Gudarzi*d
aDepartment of Polymer Engineering and Color Technology, Amirkabir University of Technology, P.O. Box 15875-4413, Tehran, Iran
bYoung Researchers and Elite Club, Shahreza Branch, Islamic Azad University, P.O. Box: 86145-311, Shahreza, Iran
cDepartment of Polymer Engineering, Technical Faculty, South Tehran Branch, Islamic Azad University, P.O. Box 11365-4435, Tehran, Iran. E-mail: d_zarei@azad.ac.ir; Tel: +98 21 88830326
dDepartment of Inorganic and Analytical Chemistry, School of Chemistry and Biochemistry, University of Geneva, Geneva, Switzerland. E-mail: mohsen.moazzami@unige.ch
First published on 3rd March 2016
Multifunctional flexible polyurethane (PU)/reduced ultralarge graphene oxide (rUL-GO) composite foams with low density in the range of ∼53–92 kg m−3 were fabricated through the in situ polymerization of PU in the presence of rUL-GO. The incorporation of 1 wt% rUL-GO gave the insular flexible PU composite foams a high electrical conductivity of 4.04 S m−1, and an excellent electromagnetic interference (EMI) shielding efficiency of ∼253 dB (g−1 cm−3) at 8–12 GHz. Achieving such a high specific EMI shielding efficiency as well as a low percolation threshold, combined with the method of foam preparation, results in a uniform dispersion and very high aspect ratio (>20000) for the rUL-GO nanosheets. Furthermore, by the introduction of rUL-GO, the Young’s modulus and tensile strength of the PU composite foams also improved significantly without reducing the flexibility. TGA experiments also indicated the enhanced thermal stability of the composite foams.
A common practice for making conductive composites with suitable EMI shielding efficiencies is to mix a polymer matrix with carbonaceous nanoparticles such as carbon black, carbon nanofibers, carbon nanotubes (CNTs), and graphene and its derivatives. A carbon fiber/polystyrene (PS) composite containing 15 vol% carbon fibers showed an EMI shielding efficiency of 73.9 dB in the frequency range of 8–12 GHz.1 A PS/Fe3O4/thermally reduced graphene oxide (TRG) composite had an EMI shielding effectiveness of 25–35 dB in the range of 9.8–12.0 GHz with 2.24 vol% TRG.2 An EMI shielding effectiveness of approximately 32 dB was obtained for a water-borne polyurethane (PU) composite containing 5 vol% non-covalent modified graphene nanosheets.3 Composite sheets consisting of phenolic resin filled with a mixture of 37 wt% reduced graphene oxide (rGO), 12 wt% γ-Fe2O3 and 1 wt% carbon fibers showed an EMI shielding effectiveness of 45.26 dB in the frequency range of 8.2–12.4 GHz.4 A multilayer graphene/EVA composite film prepared as paraffin-based sandwich structures of 350 μm thickness had a 27 dB EMI shielding efficiency with 69 vol% multilayer graphene nanosheets.5 The EMI shielding effectiveness of a rGO–SiO2 composite containing 20 wt% rGO was found to be 38 dB at 473 K.6 The highly aligned rGO/epoxy composite with 2 wt% rGO exhibited an EMI shielding efficiency of 38 dB in 0.5–4 GHz.7 The porous graphene/PS composite containing 5.6 vol% graphene showed an EMI shielding effectiveness of 25–30 dB in the frequency range of 8–12 GHz.8 The EMI shielding effectiveness of a graphene/polyaniline composite was found to be 34.2 dB with 33 wt% graphene at 8–12 GHz.9 Yang et al.10 prepared PS/CNT composite foams using a chemical blowing agent. The prepared PS composite foam with 7 wt% CNTs exhibited an EMI shielding efficiency of ∼20 dB at 8–12 GHz. Zhang et al.11 reported an EMI shielding efficiency of 13–19 dB for a polymethylmethacrylate/graphene composite foam with 5 wt% graphene in the frequency range of 8–12 GHz. The graphene–Fe3O4/polyetherimide (PEI) composite microcellular foam showed an EMI shielding efficiency of 14–18 dB with a graphene–Fe3O4 loading of 10 wt% at 8–12 GHz.12 A prepared graphene/poly(dimethyl siloxane) composite foam containing 0.8 wt% graphene had a 20 dB EMI shielding effectiveness in the frequency range of 8–12 GHz.13
Among the carbonaceous materials, the latest form, graphene, is attracting a great deal of attention not only due to its atomically thin nature and extraordinary properties but also because graphene is more eco-/environmentally friendly.14,15 Bulk production of graphene and its derivatives is now easily possible using wet chemical methods and it is more cost effective than the tubular form of carbon i.e. CNTs. Despite the encouraging results on the EMI shielding of graphene-based polymer composites,16 one still wonders if the optimal performance of graphenic sheets has been obtained or not. In addition, there are the complicated functions of different parameters such as level dispersion and alignment, the intrinsic properties of the graphenic sheets, interaction with the polymer matrix etc. Let us simplify the problem and just imagine that the optimum performance of graphene for increasing electrical conductivity is the initial goal.
Adding graphene (and other conductive fillers) to a polymer matrix increases the electrical conductivity if an interconnected network of conductive graphenic sheets forms. As soon as this percolative network forms, electrons can pass through the matrix. The threshold for this increase in conductivity depends on the size of the graphenic sheets. In fact, the reason this network forms, is the excluded volume interaction among the sheets which is due to their 2D nature. The larger they are, the higher the chance a network will form. Therefore, it is preferred to use larger graphenic sheets. Another advantage of using large graphenic sheets is the lower resistance for passing electrons through the composites. There is a large resistance at the polymer–graphene interface. Using larger sheets substantially decreases the number density of these undesirable resistances. The remaining challenge is to find an effective way of dispersing giant graphenic sheets in the polymer matrix.
Lightweight PU foams are an excellent candidate as matrices for EMI shielding due to their superior mechanical properties, facile processability, very low density etc. In this study we used PU as the matrix for the fabrication of lightweight, conductive and flexible graphene-based composites. Graphene oxide with a very high lateral size (a few tens of microns, ultra-large graphene oxide, UL-GO) was synthesized according our previous work.17 Based on a novel method, UL-GO was incorporated into a PU matrix and its large size was preserved during the processing. In situ polymerization of PU monomers in the presence of rUL-GO was applied to prepare microcellular PU/rUL-GO composite foams, and rUL-GO was added, with a loading of up to 1 wt%. The properties of the PU/rUL-GO composite microcellular foams were investigated in terms of their microstructure, electrical conductivity, EMI shielding efficiency, thermal stability, and mechanical properties. To the best of our knowledge, few studies have investigated the preparation of lightweight microcellular PU composite foams for their application in electrical conducting and EMI shielding fields. This study provides a reproducible and scalable method which may develop the commercialization of such high performance, lightweight, and multifunctional foams.
![]() | ||
Fig. 2 (a) OM and (b) SEM images of UL-GO deposited on a Si/SiO2 substrate, and (c) TEM micrographs of the as-produced UL-GO nanosheets. |
Fig. 3 shows the XRD patterns of the natural graphite flakes, UL-GO, and rUL-GO samples. A sharp reflection, 2θ = 26.5°, in the XRD scattering pattern of pristine graphite originates from the interlayer (002) spacing (d-spacing = 0.337 nm). Upon oxidation, neighboring layers are ∼0.87 nm apart (2θ = 10.1°) because of the intercalation of oxygen groups and moisture. Moreover, in the case of UL-GO, peak broadening and a shift to a lower angle takes place owing to the disorder introduced by a reduced average crystalline size. For rUL-GO, the characteristic reflections are absent, implying high exfoliation via the chemical reduction of UL-GO.17,25
![]() | ||
Fig. 3 XRD patterns of (a) natural graphite flakes, (b) UL-GO, and (c) rUL-GO samples. The d-spacing is represented by d. |
The results from the XPS analysis are shown in Fig. 4. The C1s spectra of UL-GO and rUL-GO are compared. The spectra were deconvoluted into three peaks corresponding to the following functional groups: carbon sp2 (CC and C–C), carbon sp3, epoxy/hydroxyl (C–O), and carbonyl (C
O), and the summary is presented in Table 1. The C1s XPS spectra of UL-GO indicate a considerable degree of oxidation which indicates the presence of different oxygen functional groups in the UL-GO structure (e.g., carbonyl, epoxy, hydroxyl groups).17,26,27 The peaks corresponding to the covalent bonds of carbon and oxygen atoms are more intense for UL-GO than for rUL-GO. Therefore, it can be said that the reduction process was successful in eliminating oxidized carbon in the form of carbonyl and epoxy/hydroxyl groups.
Binding energy (eV) and assignation | 284.6 eV, C–C and C![]() |
286.6–287.0 eV, C–O | 287.8–288.2 eV, C![]() |
---|---|---|---|
UL-GO | 48.6 | 44.8 | 6.6 |
rUL-GO | 70.0 | 17.5 | 12.5 |
The FTIR spectra of UL-GO before and after reduction by hydrazine are shown in Fig. S1.† UL-GO displayed the typical absorption bands at 1717 (carboxyl CO), 1633 (C
C), 1398 (carboxy C–O), 1225 (epoxy C–O), and 1055 cm−1 (alkoxy/alkoxide C–O). Also, a broad band centered at 3437 cm−1 was seen in the spectrum which is ascribed to the hydroxyl groups located on the surface of UL-GO.28 After the reduction, the intensities of all absorption bands related to the oxygenous groups decreased dramatically, indicating that UL-GO had been almost completely reduced to rUL-GO.
The Raman spectra for UL-GO and rUL-GO were also characterized as shown in Fig. 5. For both samples, distinctive peaks for the D bands and G bands could be observed. It was observed that the D to G band intensity ratio, ID/IG, increased from 0.78 (UL-GO) to 0.83 (rUL-GO) and the G peak position red shifted upon reduction to rUL-GO. The decrease in the size of the newly formed graphene-like sp2 domains is also partly responsible for the increase in the intensity ratio, ID/IG.17,29 Moreover, a prominent D peak is a reflection of disorder in the carbon structure. The above observations are consistent with previous reports on similar rUL-GOs.30,31
![]() | ||
Fig. 7 FTIR spectra of pure PU and PU/rUL-GO1 foams; the highlighted region in the original spectrum is magnified in the right-hand image. |
Fig. 7 shows the FTIR spectra of the pure PU and PU/rUL-GO1 foams. Since the rUL-GO nanosheets derived by chemical synthesis possess remnant oxygenated groups, these active sites may interact with the urethane, carboxyl and isocyanate groups of the PU chains to give strong interfacial adhesion.18,35–37 The pure PU foam showed its characteristic bands at: 3427 and 3520 cm−1 (as a shoulder), which are attributed to the stretching vibrations of hydrogen bonded and free N–H groups, respectively. The absorption bands related to the non-symmetric and symmetric stretching vibrations of CH2 groups were observed at 2930 and 2866 cm−1, respectively. The peaks at 1722, 1228 and 1072 cm−1 are attributed to the stretching vibration of carbonyl (CO) groups, the stretching vibration of aromatic C–O groups, and the non-symmetric stretching vibration of C–O–C groups, respectively. The combination of N–H deformation and C–N stretching vibrations occurs at 1533 and 1228 cm−1, respectively. By comparing the FTIR spectra of pure PU and PU/rUL-GO1 foams, two significant changes are noted. First, the shoulder at 3520 cm−1 completely disappeared for the PU/rUL-GO1 foam. Second, the peak related to the C
O groups of PU was shifted from 1722 to 1699 cm−1. These results indicate that both the free N–H and the C
O groups of PU become involved in hydrogen bonding interactions with the remaining oxygenated groups of the rUL-GO nanosheets.35,37 Besides this, the bands at 1379, 1228 and 1072 cm−1 completely disappeared after the addition of rUL-GO, further corroborating the presence of strong interfacial adhesion and chemical interactions between the PU chains and the rUL-GO nanosheets.
Fig. 8 depicts the electrical conductivity of the PU composite foams plotted as a function of the rUL-GO content. The conductivity of pure PU (∼10−11 S cm−1) is in good agreement with the values reported in the literature.18 On the other hand, the conductivity of thermally exfoliated graphene has been reported to be ∼80 S cm−1.25 As can be seen, the electrical conductivity of the PU composite foams exponentially increased at low rUL-GO contents, followed by a slow increase at higher contents. Thanks to the homogeneous dispersion of monolayer rUL-GO nanosheets in the PU matrix, the electrical conductivity was sharply raised by nearly 7 orders of magnitude when 0.4 wt% rUL-GO was incorporated. Above 1 wt% of rUL-GO, the electrical conductivity increased at a lower rate, then reached a plateau. It is noteworthy to mention that an electrical conductivity of 4.04–9.34 S m−1 was achieved for a rUL-GO content in the range of 1–2 wt%, making these composite foams very promising for consideration in the electronic industry for areas such as electromagnetic interference shielding, electrostatic discharge protection, actuators and photoconductors.
![]() | ||
Fig. 8 Electrical conductivity of PU/rUL-GO composite foams as a function of rUL-GO content (wt%); inset: log![]() |
It should be added that, generally, as the amount of conductive filler gradually increases, the foam undergoes a rapid insulator-to-conductor transition at a certain nanofiller content, referred to as the percolation threshold. The percolation threshold, φc, was calculated based on the power law equation:38,39
σc = σf(φ − φc)n | (1) |
To the best of the authors’ knowledge, up until now, the obtained percolation threshold value is one of the lowest reported in the literature, not only for polymer/graphene composites foams but also for bulk composites. The lowest percolation threshold values of composite foams filled with graphene for different polymer matrices are summarized as follows; PMMA (φc ∼ 0.66 wt%),11 polyvinylidene fluoride (φc ∼ 0.5 wt%),42 and PEI (φc ∼ 0.63 wt%).41 The ultralow percolation threshold value obtained in this work is ascribed to the following features of the PU composite foams: first, the inherently high electrical conductivity of rUL-GO; second, the good state of dispersion; and last but not the least, the ultralarge size of the monolayer graphene nanosheets.
Kim et al.43 proposed an improved analytical model based on the average interparticle distance concept, predicting the percolation threshold of polymer composites filled with nanoplatelets or disk-shaped fillers. According to the results of this model, a percolation threshold of 0.029 vol% is equivalent to an aspect ratio in the order of 104. Assuming that all of the rUL-GO nanosheets were exfoliated in the PU matrix as monolayer sheets (i.e., a thickness of about 1 nm), an aspect ratio of 104 would be equivalent to an average lateral size of 10 μm. This estimation is in the same region as the size of GO sheets used in this study, verifying the molecular-level dispersion of individual rGO nanosheets in the PU matrix.39 The lower estimation of the aspect ratio compared to what was observed experimentally (Fig. 2 and ref. 17) might be due to mere breakage of the nanosheets, the inherent flexibility of them (especially at such high aspect ratio) and also imperfect dispersion. However, these findings further emphasize the high potential of using rUL-GO nanosheets as highly efficient ECNs for the production of highly electrically conductive materials.
The EMI shielding of the PU/rUL-GO composite foams with a thickness of 2.5 mm was determined in the microwave frequency range of 8–12 GHz (Fig. 9). The measured shielding efficiency is defined as
![]() | (2) |
![]() | ||
Fig. 9 (a) EMI shielding efficiency of PU/rUL-GO composite foam at different frequencies. (b) Comparison of the specific EMI shielding efficiency of our material with other reported materials. |
The mechanical strength of the composite foams is of great importance, especially when they are generally exposed to high stress and strain fields. The interfacial interactions, orientation, aspect ratio, and surface functional groups of the nanofiller are some important factors which play a key role in determining the mechanical properties.46–48 The mechanical properties of pure PU and PU composite foams (Fig. S2†) are summarized in Table 2. Even though the changes in the mechanical characteristics of the samples were remarkable, they still exhibited their typical elastomeric behavior. Interestingly, a content of as low as 0.1 wt% of rUL-GO increased the Young’s modulus and tensile strength from 5.5 to 42.6 MPa and from 15.3 to 30.2 MPa, respectively. The elongation at break of the composite foams was lower than that of the pure PU foam, but the obtained values are still in the range of high-elongation-at-break foams. The notable improvement of the modulus in the PU composite foams is due to the well exfoliated rUL-GO nanosheets as well as strong interfacial bonding between the PU chains and rUL-GOs. As demonstrated in the FTIR section, the remaining oxygenated functional groups on the surface of rUL-GO interact with the urethane groups of PU. Such interactions lead to well dispersed and exfoliated monolayer rUL-GO nanosheets, and consequently result in strong interfacial adhesion between PU and the nanosheets.18,37,49 However, these interactions increase the molecular restrictions of the PU chains, and in turn decrease the elongation at break of the composite foams as compared to that of the pure PU foam.
Sample | Tensile strength (MPa) | Young’s modulus (MPa) | Density (kg m−3) | Porosity (%) |
---|---|---|---|---|
Pure PU | 15.3 ± 0.05 | 5 ± 0.1 | 53 ± 6 | 95.58 ± 0.50 |
PU/rUL-GO0.1 | 30.2 ± 0.06 | 42.6 ± 0.3 | 58 ± 5 | 95.16 ± 0.41 |
PU/rUL-GO0.5 | 38.4 ± 0.04 | 62.9 ± 0.1 | 72 ± 7 | 94.00 ± 0.59 |
PU/rUL-GO1 | 54.2 ± 0.2 | 75.2 ± 0.2 | 92 ± 9 | 92.33 ± 0.75 |
The surface morphology of the pure PU and PU/rUL-GO1 foams is shown in Fig. 10. The PU foam exhibits a 3D-hierarchical mesoporous structure with the pore size ranging from 500 to 1200 μm which results from the high expansion ratio of the foaming. The density of the PU composite foams increased with the increasing rUL-GO content which is ascribed to an increase of the polyol viscosity50,51 (Table 2). The increase in the density of the PU composites not only decreased the porosity, but also led to more polyhedral open cell structures. Despite this porosity reduction, the composite foams are still considered as highly porous materials, and therefore possess good compressibility and large internal surface areas. Moreover, the smooth inner and outer layers of these polyhedral open cell structures are free from the interjunctions that enhance the superior electrical conducting path.
![]() | ||
Fig. 10 SEM micrographs of (a) pure PU and (c) PU/rUL-GO1 foams. (b) and (d) are the corresponding higher magnification images of (a) and (c), respectively. |
The thermal stability of the pure PU and PU/rUL-GO1 foams is shown in Fig. 11. It has widely been reported that PU has a three-stage degradation process at 370–400, 400–440, and 470–550 °C. These three characteristic temperatures are referred to as T1st, T2nd, and T3rd, respectively. The first stage is attributed to the dimerization and trimerization reaction of isocyanates;18,52 the second step is due to the de-polymerization of PU to form isocyanates, primary or secondary amines, polyols, olefins, and carbon dioxide; and the last degradation step is assigned to the decomposition of the hard segment.53,54 It is important to note that the slight variations, though negligible, seen in the thermogram curves of the pure PU foam are due to the presence of foaming additive components. Similarly, the typical three-stage degradation process is seen in the PU/rUL-GO1 thermogram curve. The effect of 1 wt% rUL-GO on the degradation steps of the PU foams, in terms of the characteristic temperatures, is shown in Table 3. The addition of 1 wt% rUL-GO, increases T1st, T2nd, and T3rd by about 14, 27, and 37 °C, respectively. Moreover, as can be seen, the addition of 1 wt% rUL-GO improved the heat dissipation as well as the thermal stability of the PU/rUL-GO1 foam as compared to the pure PU one. There are several factors responsible for such a conspicuous thermal stability improvement. First, the interaction of the remaining oxygenated functional groups of rUL-GO with the urethane groups of PU, mentioned in the FTIR section, enhance the interfacial adhesion between the PU chains and rUL-GO.55,56 Well dispersed and monolayer exfoliated rUL-GO nanosheets provide a high surface for direct covalent bonding with the PU matrix. These results continue to support the conclusion that the rUL-GO nanosheets can act as a barrier to reduce the diffusion of volatile compounds and hence an enhancement of the thermal stability is observed.18
Sample | T1st | T2nd | T3rd |
---|---|---|---|
Pure PU foam | 382 | 420 | 521 |
PU/rUL-GO1 foam | 396 | 447 | 557 |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra25374h |
This journal is © The Royal Society of Chemistry 2016 |