Sandeep K. Markaa,
Bashaiah Sindamb,
K. C. James Rajubc and
Vadali V. S. S. Srikanth*a
aSchool of Engineering Sciences and Technology (SEST), University of Hyderabad, Gachibowli, Hyderabad 500046, India. E-mail: vvsssse@uohyd.ernet.in; Tel: +91 40 23134453
bSchool of Physics, University of Hyderabad, Gachibowli, Hyderabad 500046, India
cAdvanced Centre of Research in High Energy Materials (ACRHEM), University of Hyderabad, Hyderabad 500046, India
First published on 14th April 2015
1 mm thick flexible few-layered graphene (FLG)/poly vinyl alcohol (PVA) composite sheets containing 0.1 and 0.5 vol.% of FLG are prepared using an easy solution mixing process followed by a simple casting process. As-synthesized FLG is used as the filler. The morphology, structure and phase characteristics clearly showed the formation of a composite. A maximum electromagnetic interference (EMI) shielding effectiveness (SE) of ∼19.5 dB (in the X-band, 8.2–12.4 GHz) was obtained in the case of the composite with 0.5 vol.% of FLG. Absorption is found to be the dominant mechanism for EMI shielding. The high EMI SE is attributed to the network-like features formed by FLG in the PVA matrix. EMI SE and absorption are understood by analyzing the dielectric behaviour, ac-conductivity and transparency to visible light of the composite sheets.
Of late, several polymer based composites with carbonaceous fillers have exhibited excellent EMI SE. These composites are light in weight, mechanically stable and flexible, and most importantly they are cost-effective when compared to other shielding materials.1–5 However, in the case of some composites, time consuming and difficult multi-step procedures have to be followed to obtain the final product. In some cases synthesis is easy but scalability is a problem. In view of this, it has been suggested that the disadvantages (namely difficulty in synthesis, presence of impurities, intrinsic bundling, high price etc.) involved in the case of certain carbonaceous fillers could be mitigated by using graphene as a filler material.5–7 Some of the important polymer based composites with carbonaceous fillers and their EMI SE are listed in Table 1.
Composite | Frequency range (GHz) | Maximum EMI SE (dB) |
---|---|---|
a MWCNT = Multi Wall Carbon Nanotubes, rGO = reduced graphene oxide. | ||
MWCNT/poly (vinylidene fluoride)–poly (vinyl pyrrolidone)8 | 0.0–1.5 | ∼20 |
MWCNT/poly (methylmethacrylate)9 | 0.05–13.5 | ∼27 |
Polystyrene–MWCNT–graphite nanoplate microbeads10 | 8.2–12.4 | ∼20.2 |
Carbon nanofibers filled PVA11 | 0.85 | 1.6–4.8 |
Conductive carbon black and short carbon fiber filled natural rubber and ethylene vinyl acetate12 | 8–12 | ≥ 20 |
PANI–graphite composites13 | 0.01–1 | ∼27 |
Functionalized graphene/epoxy14 | 8.2–12.4 | ∼21 |
Graphene/poly (ethylene oxide)15 | 2–18 | −38.8 reflection loss |
Microcellular polyetherimide–graphene foams16 | 8.2–12.4 | ∼44 |
Aligned graphene sheets in wax17 | 8.2–12.4 | ∼12 |
rGO/SiO2 composite18 | 8.2–12.4 | ∼37 |
rGO/wax and graphite nanosheet/wax19 | 2–18 | ∼29.68 and ∼10 |
In this work an easy and cost effective method to obtain flexible sheets of few-layered graphene (FLG)/poly vinyl alcohol (PVA) composite that is useful for EMI shielding has been demonstrated. PVA has been chosen as the matrix material owing to its water solubility, high transparency, very good flexibility and mechanical strength and wide commercial availability. In addition, PVA has excellent film forming, emulsifying and adhesion properties.20 There are only two reports on graphene/PVA composites especially for EMI shielding in the X-band frequency range21,22 while all other reports are on electrical conductivity and mechanical behavior. In these works, synthesis of graphene/PVA composites involves multiple steps such as unzipping MWCNT, in situ reduction of graphene oxide in polymer using hazardous reducing agents like hydrazine hydrate and so on. On the contrary, in this work, the synthesis procedure involves efficient use of ultrasonic waves, simple solution mixing of as-synthesized FLG with PVA to form FLG/PVA composite and simple casting process to form flexible FLG/PVA composite sheets which exhibit an excellent EMI shielding ability.
A special procedure was used to prepare TEM specimen from the samples. Initially borosilicate glass was cleaned in acetone and isopropanol alcohol using bath sonication and finally rinsed with deionized water. It was then dried in hot air oven at 100 °C for 1 h to remove residual alcoholic contents. Polystyrene was then spin coated on the glass at 1000 rpm. The polystyrene coated glass was then dried at 130 °C for 1 h. FLG/PVA composite solution was then spin coated at 8000 rpm on the surface of the polystyrene coated glass and subsequently dried at 130 °C. A part of this film (FLG/PVA composite on top of polystyrene film) was then peeled from the glass and placed on the TEM grid which was then dipped in toluene to dissolve polystyrene leaving a free standing FLG/PVA composite film on the TEM grid.
The measured scattering parameters S11 and S12 are related to the reflected and transmitted power, respectively w.r.t the power incident on the sample surface. The incident power (Pi) on a shielding material is divided into reflected power (Pr), absorbed power and transmitted power (Pt) at the output of the shielding. The EMI shielding effectiveness (EMI SE) of a material is defined as SEtotal = −10log(Pt/Pi).7,10 When an electromagnetic radiation is incident on a shielding material the sum of absorption coefficient (A), reflection coefficient (R) and transmission coefficient (T) must be equal to 1. R, T and A can be calculated from S parameters using the formulae R = (ER/EI)2 = |S11|2 = |S22|2, T = (ET/EI)2 = |S12|2 = |S21|2 and A = 1 − R − T, respectively. Total EMI SE (SEtotal) is the sum of reflection from the material surface (SER), absorption of electromagnetic energy inside the material (SEA) and multiple internal reflections (SEM) of electromagnetic radiation, expressed as SEtotal = SER + SEA + SEM. The reflection is related to the impedance mismatch between air and absorber. Absorption is regarded as the energy dissipation of electromagnetic wave in the shielding material over multiple reflections at the interfaces and scattering from inhomogenieties inside the material while the multiple reflections are the consequence of impedance mismatch at the two sample–air interfaces. When SEtotal ≥ 15 dB, it is usually assumed that SEM is negligible and thus, SEtotal ≈ SER + SEA. The effective absorbance (Aeff) can be therefore expressed as
. The shielding effectiveness due to reflection and absorption of the shielding material with respect to power of the effective incident electromagnetic wave inside the shielding material are expressed as SER = −10
log(1 − R) and SEA = −10
log(T/(1 − R)), respectively whilst SEtotal = SER + SEA = −10
log(T).
From Fig. 1 it is clear that the synthesis method used in this work converts large number of stacked graphene layers i.e., graphite flakes (Fig. 1(a)) into FLG (Fig. 1(b)). Cross-sectional micrographs of PVA and FLG/PVA composites are shown in Fig. 2. PVA (Fig. 2(a)) contains a mixture of flat and folded regions. A typical flat region at high magnification is shown in Fig. 2(b). 0.1 vol.% FLG/PVA composite contains (Fig. 2(c)) uniformly distributed FLG (Fig. 2(d)) particles which occupy very small regions in the PVA matrix. 0.5 vol.% FLG/PVA composite contains uniformly distributed network-like features (Fig. 2(e) formed by FLG (Fig. 2(f)) in PVA matrix.
![]() | ||
Fig. 2 Cross-sectional secondary electron micrographs of (a), (b) PVA and FLG/PVA composites with different vol.%, (c), (d) 0.1, and (e), (f) 0.5 vol.% of FLG. |
FLG sheets are transparent to the electron beam in TEM as shown in Fig. 3(a). This indicates the presence of few graphene layers in each FLG construction. Comparison of XRD data of FLG with graphite (Fig. 3(b)) shows typical (002) and (004) Bragg peaks with lower intensity indicating the formation of FLG. The X-ray diffractogram obtained from FLG also indicates that it is highly ordered. Selective area electron diffraction (SAED) pattern of FLG shows the typical six fold symmetry as expected for graphene. The regular outer and inner hexagon patterns with varied intensity of the diffraction spots as shown in the inset of Fig. 4(a) indicates multi-layer system that resembles A-B type of atomic stacking as in graphite.25 PVA is a partially crystalline polymer that exhibits a strong (101) diffraction peak at 2θ = 19.5° and a weak (and broad) peak at 2θ = 40.4°. FLG has a characteristic (002) diffraction peak at 2θ = 26.5° as shown in Fig. 4.
![]() | ||
Fig. 3 (a) Transmission electron micrograph of FLG (inset shows the corresponding diffraction pattern) and (b) comparison of XRD data of graphite and FLG. |
In the case of composites, it is observed that the intensity of (101) peak of PVA is higher in the case of composites owing to the uniform distribution of FLG (in PVA matrix) which acts as nucleation sites for PVA chains to pack together resulting in large size crystallites in PVA sol–gel26 and thus superior PVA crystallinity in composites. Increase in intensity of FLG diffraction peak in case of composites with increasing FLG loading is due to increase in the volume content of FLG.
XRD results are well complemented with transmission electron microscopy results. As shown in Fig. 5, the transmission electron micrographs show that all composites are transparent to electron beam. Fig. 5(a) and (b) show bright field micrographs of 0.1 vol.% FLG filled PVA composite and high resolution TEM micrograph, respectively. Selected area electron diffraction (SAED) pattern indicates one set of six-fold-symmetric spots which corresponds to six fold symmetry of graphene.27 High resolution TEM micrograph clearly shows 2 to 4 graphene layers which confirms that the synthesized composite contains FLG. Fig. 5(c) and (d) show bright field TEM micrograph and high resolution TEM micrograph, respectively of 0.5 vol.% FLG filled PVA composite. SAED pattern contains family of spots, indicating that field of view contains several grains (individual FLG structures) in different orientations.28 This is well supported by high resolution TEM micrograph (Fig. 5(d)) which shows the presence of several FLG structures in different orientations and the network-like formation by FLG in PVA.
![]() | ||
Fig. 5 Transmission electron micrographs of FLG/PVA composites with (a & b) 0.1 vol.% FLG, (c & d) 0.5 vol.% FLG (insets: (a) and (c) are the corresponding SAED patterns). |
PVA's Raman spectrum typically depicts 17 vibrational modes in (–CH2–CHOH–) monomer. These modes are: 6 CH2 (2 stretching, bending, wagging, twisting and rocking), 3 CH, 3 OH, 3 CO (each consists of a stretching mode plus 2 bending modes, perpendicular and parallel to the chain axis) and 2 CC (skeletal vibrations).29 In each sample's case, an average of six Raman spectra is considered for good accuracy in analysis. A clear explanation for assignment of bands in PVA is available elsewhere.30,31 The Raman spectra corresponding to different samples are shown in Fig. 6. The typical Raman bands at 2840, 2910, and 2942 cm−1 correspond to the stretching modes of CH and CH2 i.e., ν(CH) and νs(CH2) (and νa(CH2)), respectively.29 In a separate work the same are tentatively assigned to undifferentiated C–H stretching vibrations.30 The ambiguity for the assignment was removed in a separate work.31 Based on the previous work,30 the bands at 2835, 2913 and 2935 cm−1 in the present work are assigned to νa(CH2), ν(CH), and νs(CH2), respectively. As observed in Fig. 6(b) there is a small shift in the peak positions with the addition of FLG to the PVA matrix. The shift is attributed to stresses that are induced in the matrix owing the presence of FLG. The symmetric bending mode of the CH2 group, δ(CH2), is observed (Fig. 6(a)) in PVA at ∼1433 cm−1 and in FLG/PVA composites in between ∼1435 and ∼1438 cm−1. The wagging and rocking modes of the CH2 group (γw(CH2), γr(CH2)) are observed (Fig. 6(a)) in PVA matrix at ∼1362 and ∼852 cm−1 and in FLG/PVA composites at ∼1377 cm−1 and in between 848 and 856 cm−1, respectively. The twisting mode of CH2, γt(CH2) is too weak to be distinguished from noise of the spectrum. The stretching and wagging modes of the OH group, ν(OH) and γw(OH) are observed in PVA at 3361 and 639 cm−1, respectively and the same are observed in FLG/PVA composites with slight peak shift. The peak at 1440 cm−1 in PVA and at 1447 cm−1 in FLG/PVA composites are attributed to a combined effect of bending mode vibration of CH and OH groups (δ(CH + OH)). It is not possible to assign the exact position of the peaks in the range 1000–1200 cm−1 due to lack of crystallinity and limitation with instrument's resolution (∼3 cm−1). The band at ∼1096 cm−1 is probably due to ν(CO) in both PVA and FLG/PVA composites. The skeletal CC vibration is observed at 922 cm−1 in PVA and at 914 cm−1 in FLG/PVA composites. A broad band at ∼1723 cm−1 is due to residual acetate groups in PVA.
![]() | ||
Fig. 6 Comparison of Raman spectra of PVA, FLG and FLG/PVA composites in (a) 200–1500 cm−1 and (b) 1500–3800 cm−1 regions. |
The Raman spectra of FLG and FLG/PVA composites (Fig. 6) showed distinctive D and G bands representative of graphitic material.35 In FLG, D, G, M-K scattering, 2D, and 2D′ bands are observed at 1354, 1578, 2442, 2702 and 3234 cm−1, respectively. Various Raman bands in FLG/PVA composites are shown in Table 2.
Raman feature | FLG | 0.1 vol.% FLG/PVA composite | 0.5 vol.% FLG/PVA composite |
---|---|---|---|
D band | 1354/cm | 1356/cm | 1356/cm |
G band | 1578/cm | 1581/cm | 1581/cm |
M-K scattering band | 2442/cm | 2445/cm | 2457/cm |
2D band | 2702/cm | 2712/cm | 2716/cm |
2D′ band | 3234/cm | 3242/cm | 3243/cm |
ID/IG ratio | 0.6 | 0.5 | 0.6 |
IG/I2D ratio | 1.2 | 1.4 | 1.5 |
Raman peak positions pertaining to FLG in the present study may correspond to Raman scattering in folded graphene, single, double, and FLG.32–34 In the case of FLG, the intensity ratio ID/IG which is measure for the extent of disorder was ∼0.6 indicating high order graphitization while IG/I2D was ∼1.2 corresponding to less than 20 graphene layers.35 It is well known that 2D band in the Raman spectrum of graphitic material is a fingerprint for number of graphene layers in the given graphitic material and one constraint for this technique is that, it is hard to distinguish more than 5 to 6 layers from graphite.34 The maximum intensity count of 2D band in FLG was observed at 2702 cm−1 with a good center of symmetry around 2700 cm−1. This type of symmetry will be found for single or bilayer or highly ordered pyrolytic graphite (HOPG).36 But in comparison to a work reported previously,34 the symmetry of 2D band indicated that the sample in the present study (i.e., FLG) contains very few layers and it is definitely not HOPG. For an ideal and defect free graphene it is well known that G and 2D bands in Raman spectrum are observed at 1590 and 2685 cm−1, respectively. As the number of layers increases the G band shifts towards lower wave number side, 2D band shifts towards higher wave number side and shape of the 2D band will change from sharp, symmetric peak to broad, asymmetric peak. For example, for a single layer graphene one sharp 2D band can be observed, for a bilayer graphene 2D mode can be decomposed in four components and for HOPG it is best fitted to two components with less intense 2D1 component and high intense 2D2 component. It was also well observed that as number of layers increase an increment of intensity of the higher frequency 2D2 component compared to the 2D1 component occurs. In case of FLG/PVA composites, the D and G band observed at 1356 and 1581 cm−1 respectively and 2D band shifted to higher wave number side as loading level increased. As the loading level of FLG increased, number of graphene layers also increased (as observed in high resolution TEM micrographs) which is reflected as a change in 2D peak position and shape. G band peak position also shifted towards higher wave number side which is a contradictory result. Recently it has been reported37–39 that such a blue shift in G peak position can be attributed to unintentional doping in graphene by the charge impurities present in the surrounding atmosphere (here due to the PVA). It was also reported40 that defects like edges, dislocations, cracks or vacancies in the sample can cause unintentional doping. Due to this effect, even though the extent of disorder (i.e., ID/IG) increased as the loading of FLG increased, the intensity ratio of G to 2D (i.e., IG/I2D) decreased. It is anticipated that defects in FLG are created during ultrasonication process, which is supported by erosion of FLG edges as observed in Fig. 1(b).
![]() | ||
Fig. 7 (a) ac conductivity vs. frequency, (b) transmittance spectra and (c) digital photographs of different samples. |
Since ∼20 dB is the required EMI SE for a material to be useful in most of the commercial applications, 0.5 vol.% FLG/PVA composite fits in this league of materials. To further understand the reasons behind the observed increase in SE of FLG/PVA composites, complex permittivity ε* = ε′ − iε′′ is evaluated from the experimental scattering parameters (S11 and S21) using Nicholson and Ross, and Weir algorithms.41,42 The dielectric constant (ε′) is concomitant with the degree of polarization occurring in the material and the imaginary part (ε′′) is a measure of dissipated energy. In the X-band frequency range, as shown in Fig. 8(c), ε′ and ε′′ values of PVA, 0.1 vol.% FLG/PVA composite and 0.5 vol.% FLG/PVA composite have slightly decreased (encircled regions in Fig. 8(c)) with increase in frequency and their average values are 2.8 and 0.2, 4.0 and 0.6 and 5.1 and 0.82, respectively.
It is well-known that the overall dielectric behavior of a material depends on the degree of ionic, electronic, orientational and space charge (or interfacial) polarization in the material. During the synthesis, combination of microwave irradiation of graphite oxide and subsequent exposure to ultrasonic waves induce residual groups and defects like missing carbon atoms and sheet corrugation (as shown in Fig. 1(b)) in the hexagonal carbon lattice of FLG. Energy transition of microwave band involves the electronic spin, which means greater spin states are required for microwave absorption. It has been reported that localized states near to the Fermi level could be created via introducing lattice defects.43 The existence of defects in FLG (as also indicated by Raman analysis in this work) favors absorption of electromagnetic energy by the transition from contiguous states to Fermi level when the absorbing surface is irradiated with electromagnetic waves.
Cole–Cole plots give an idea about relaxation mechanisms occurring in the material under an external electromagnetic field. In Fig. 8(d) which is a Cole–Cole plot pertaining to 0.5 vol.% FLG/PVA composite, presence of three semi-circles representing three different relaxation processes (in the composite) could be identified. Presence of defects, other residual chemical groups in FLG and interfaces between FLG and FLG and FLG and PVA is the main reason for the relaxation processes.44 As elucidated by Raman scattering and electron microscopy analyses, the defects that are prominent in FLG in the composites can act as polarization centers when they are screened by the free charges. These centers can then generate polarization relaxation under the alternating electromagnetic field and attenuate the field resulting in a profound energy loss. The existence of residual oxygen-containing chemical species such as C–O in the material (as indicated by Raman scattering analysis) can generate electric dipole polarization due to their ability to trap electrons between C and O atoms. Here, under an alternating electromagnetic field, electron motion hysteresis in these dipoles can induce additional polarization relaxation processes which are favorable in enhancing microwave absorption. Interfacial polarization (also known as Maxwell–Wagner polarization) can be another attribute to the absorption process. Interfacial polarization always occurs in a composite material constituted by materials with different dielectric constants (ε) and conductivities (σ). In the case of 0.5 vol.% FLG/PVA composite, the presence of well dispersed and network-like formation by FLG in the insulating PVA matrix (as shown in TEM micrographs) results in the formation conducting/dielectric (FLG/PVA) and conducting/conducting (FLG/FLG) interfaces to induce interfacial polarization. Due to different relaxation/spreading time constants (τ = ε/σ) of the free charge carriers of individual components in the composite material, the migration of these charge carriers through the composite material will be hindered (but differently) at various points (i.e., at interfaces) of the composite material. This hindrance leads to the accumulation of space-charge at the interfaces. Pertaining to the present work, when an external electric field is applied on FLG/PVA composite, redistribution of space-charges at the interfaces can distort the macroscopic field. This distortion appears as polarization of the charges (to an external observer) which can interact with the applied field (here, the microwave field).
This journal is © The Royal Society of Chemistry 2015 |