Next Article in Journal
An Overview of Thermal Exposure on Microstructural Degradation and Mechanical Properties in Ni-Based Single Crystal Superalloys
Previous Article in Journal
Mechanical Performance and Cytotoxicity of an Alginate/Polyacrylamide Bipolymer Network Developed for Medical Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Silica-Fiber-Reinforced Composites for Microelectronic Applications: Effects of Curing Routes

by
Imran Haider
,
Iftikhar Hussain Gul
,
Malik Adeel Umer
* and
Mutawara Mahmood Baig
Thermal Transport Laboratory, Department of Materials Engineering, School of Chemical and Materials Engineering (SCME), National University of Sciences and Technology (NUST), Islamabad 44000, Pakistan
*
Author to whom correspondence should be addressed.
Materials 2023, 16(5), 1790; https://doi.org/10.3390/ma16051790
Submission received: 27 April 2022 / Revised: 30 May 2022 / Accepted: 6 June 2022 / Published: 22 February 2023
(This article belongs to the Section Advanced Composites)

Abstract

:
For curing of fiber-reinforced epoxy composites, an alternative to thermal heating is the use of microwave energy, which cures quickly and consumes less energy. Employing thermal curing (TC) and microwave (MC) curing methods, we present a comparative study on the functional characteristics of fiber-reinforced composite for microelectronics. The composite prepregs, prepared from commercial silica fiber fabric/epoxy resin, were separately cured via thermal and microwave energy under curing conditions (temperature/time). The dielectric, structural, morphological, thermal, and mechanical properties of composite materials were investigated. Microwave cured composite showed a 1% lower dielectric constant, 21.5% lower dielectric loss factor, and 2.6% lower weight loss, than thermally cured one. Furthermore, the dynamic mechanical analysis (DMA) revealed a 20% increase in the storage and loss modulus along with a 15.5% increase in the glass transition temperature (Tg) of microwave-cured compared to thermally cured composite. The fourier transformation infrared spectroscopy (FTIR) showed similar spectra of both the composites; however, the microwave-cured composite exhibited higher tensile (15.4%), and compression strength (4.3%) than the thermally cured composite. These results illustrate that microwave-cured silica-fiber-reinforced composite exhibit superior electrical performance, thermal stability, and mechanical properties compared to thermally cured silica fiber/epoxy composite in a shorter time and the expense of less energy.

1. Introduction

Fiber-reinforced composites (FRC) are superior to other structural materials, due to their high specific strength and stiffness, high temperature, and fatigue resistance [1,2]. Glass-fiber-reinforced composites have outstanding properties, easy manufacturing, and reasonable cost compared to other fiber-reinforced composites [3,4]. The mechanical properties depend on the fiber as well as the matrix properties. The physical and mechanical properties of the thermoset resins, such as epoxy, depend on curing conditions [5]. The mechanical performance of FRC is influenced by the properties of the phases and their interactions. The interface and fiber-matrix bonding are quite important for enhancing the effective mechanical properties of FRC [6,7]. Amorphous materials (polymers, glasses, and metals) change their structural state below Tg, as physical aging or structural relaxation affect the properties of the material (mechanical, dielectric, and thermal behavior) [8]. The epoxy resin in a B-level semi-cured state can be cured under specific conditions to obtain high-quality composite products [9]. The curing degree of the resin has a very large effect on the properties of the composite [10,11,12].
In adhesively bonded structures, the quality of bonded composites has a strong relationship with the variabilities caused by process parameters, such as temperature, curing duration, and rate [13,14]. Usually, ambient, thermal curing, or a combination of both is used, depending on the requirement. As the demand for lighter, cheaper, and more compact electronic devices increases, there is a greater need to develop innovative and fast processing techniques with higher energy efficiency and reduced cure time. Longer curing times in thermal curing reduce the production throughput. In the electronics industry especially, the curing of thermoset systems has become a limiting factor in the production time; other processing methods are needed to be explored to reduce manufacturing costs or increase energy efficiency. One such method makes use of microwave radiation to heat epoxy resins [15]. The microwave curing process (MCP) can reduce production costs [16]. The MCP mechanism operates through dipolar loss, conduction loss, hysteresis loss, and Eddy current loss [17]. For thermosetting polymers, microwave curing involves the microwave radiation to heat the sample directly, which leads to efficient, fast, and selective heating, compared to conventional thermal curing, where the samples are heated indirectly. Microwave radiation also heats in a volumetric manner, thus leading to less temperature disparity and therefore a less severe temperature gradient within the material, leading potentially to less internal residual stress [18,19,20]. Microwave curing produces composites with mechanical properties comparable to those produced with the autoclave process while reducing 45% processing time and 3% energy consumption [21]. The curing kinetics has a significant influence on the macroscopic mechanical properties of composite materials [22,23,24]. Comparative studies of microwave-cured and thermally cured composites have reported contradictory results, while many authors have shown a reaction rate enhancement for microwave curing compared to thermal curing [25,26].
For many applications, the advantages of microwave curing must not be outweighed by a loss of thermal, chemical, or mechanical properties. Microwave heating is known as the most efficient volume-heating process due to its excellent depth of penetration in polymers [27]. Its process can be tuned with the use of highly efficient dielectric nanomaterials in polymers to boost micro-level heating at the molecular scale [28,29,30]. Dielectric nanomaterials can efficiently absorb radiation and convert it to molecular vibrations/rotations via dipole moments, mainly because of dipolar polarization. The vibrations can then add to the heating level in the polymer surrounding the nanomaterials by the friction mechanism [29].
The mechanical properties of certain microwave-cured materials are similar or even increased compared to conventional cures [27,28,29,30,31]. Generally, an increase in the glass transition temperature, Tg, has been reported [32]. However, depending on the type of curing agent used, Tg can be seen to decrease compared to thermally cured samples [33]. When composites are used as interconnections, printed circuit boards, and airplane skin materials, their dielectric properties become very important and must be determined before they can be used in these applications [34]. For polymer composites, the dielectric properties are associated with the component fractions [35,36] and they are widely used in electronics applications because of their good dielectric and mechanical properties. Silica fibers are used as reinforcement due to their small coefficient of thermal expansion, low thermal conductivity, superior mechanical strength, and excellent dielectric properties [37]. Compression molding, hand lay-up, spray-up, vacuum infusion, vacuum bagging, resin transfer molding (RTM), autoclave molding, filament winding, automated fiber placement (AFP), pultrusion, injection molding, vacuum forming, and stamp forming are most of the manufacturing techniques used in composite production. Prepegging produces a semi-finished composite by controlling the curing to fulfill the final requirement [38]. The curing process parameters have a crucial influence on the quality of the composite products [39]. Effects of curing degree, of Quartz fiber-boron phenolic composites, were investigated by the mechanical properties test, scanning electron microscope (SEM), and thermogravimetric analysis [40]. Fourier-transform infrared analysis shows no significant difference between the conventional and microwave cured samples [41]. The chemorheology of a filled epoxy system declares isothermal DSC measurements to be inadequate in the case of fast-curing thermosets [42]. Generally, a comparison between literature data is difficult due to the variety of curing agents used and their effects on the curing of the epoxy resin [43]. As preferred by many researchers, one of the most common methods is the mechanical testing of composite properties on different cure regimes, which can optimize the curing parameters [44].
Different composite curing studies have investigated various aspects of functional fiber-reinforced composites, but in microelectronics, low dielectric constant and loss factor of material is a matter of interest. In this study, multiple proportions of silica fiber/epoxy composite (prepregs) were cured thermally, via microwave energy, and their dielectric, structural, morphological, thermal, and mechanical properties were investigated.

2. Materials and Methods

2.1. Materials

Commercial silica fiber fabric BWT260-82 (ρ = 2.25 g/cc, avg. fiber dia = 6.86 μm, UTS = 1.8 GPa, SiO2 ~95%, thickness = 0.21 mm) was obtained from Business and Engineering Trends, Punjab, Pakistan (BET Pakistan). The bonding adhesive used was a two-component commercial Epoxy RER160 and curing agent REH160 (RESSICHEM, Karachi, Pakistan). Commercial ethanol was used as a solvent in the resin matrix preparation.

2.2. Preparation of Composite Prepregs

Unidirectional woven silica fabric was cut into fabric plies (200 × 200 mm), washed with commercial ethanol, air-dried (30 min), and then oven-dried (30 min) at 120 °C. Lab environmental conditions were 25 °C and 36% RH. In a 1000 mL glass beaker, epoxy (500 gm) and hardener (50 gm) were weighed (SHIMADZU-UW 3400 g, Kyoto, Japan), and the resin matrix was prepared with dropwise addition ethanol (30 mL). The resin matrix was poured on fabric plies and distributed with an applicator. By hand lay-up, the impregnated plies were stacked layer by layer between clean surfaces of mold plates and retained for 2 h. Stacked laminates (2 No.) were simultaneously processed to form composite prepregs (S0.3E0.7, S0.4E0.6, S0.5E0.5, S0.6E0.4, S0.7E0.3) where “S” refers to silica fabric, “E” refers to epoxy resin and the subscript refers to the wt. fraction. The fabrication process (raw material to prepreg to composite) is shown in Figure 1.

2.3. Curing of Composite Prepregs

Curing is an irreversible time-dependent and progressive molecular reaction process that requires heat energy, either by conventional heating or radiation, to raise molecular mobility [45]. Epoxy curing is an exothermic process that strengthens the material by cross-linking polymer chains [46]. Curing can be achieved by conventional heating, electron beams, chemical additives, or accelerated curing (e.g., microwave, radiofrequency, and ultra-violet radiation) [47]. The electromagnetic radiation cure mechanisms differ from thermal mechanisms in that curing is initiated by ionic or free radical intermediates formed by high-energy electromagnetic radiation [41].
The mechanism of microwave heating is shown in Figure 2. Epoxy resin is electrically neutral but dipolar in nature, as it possesses partial (+δ, −δ) charges. Microwave energy penetrated in a volumetric manner and increased the molecular mobility of resin due to periodic changing electrical fields (Figure 2). In this manner, the kinetic energy of molecules was increased and created a temperature gradient. Keeping in view the previous studies, the curing time and temperature were experimentally designed.
Five composite prepregs (of each proportion) were cured. “T″ refers to thermally cured, “M” to microwave cured, and “tc” to cure time. In the thermal cure method, the heating oven was raised to 40 °C then composite prepregs were put in (heating rate was 10 °C/min) to reach the curing temperature (160 °C). Curing times were 240 s, 300 s, 360 s, 420 s, 480 s, 600 s, and 660 s. In the microwave cure cycle, a set of composite prepregs was heated (300 W) for 20 s, 30 s 45 s, 60 s, 75 s, 90 s, and 100 s.

2.4. Characterization

The dielectric constant (Ɛr) and dielectric loss (δ) were measured (S-band) at PNA Network 8362B (Agilent) with a 3 mm inner bore circular disc (2 × 6 mm). The spectrums of thermally and microwave-cured silica fiber/epoxy composites (5 mg) were recorded at a resolution of 4 cm−1 using an FTIR spectrometer (Spectrum 100, Perkin-Elmer, Waltham, MA, USA). The microscopic morphology was obtained through scanning electron microscopy (SEM) (JSM-6490A, EOL, Tokyo, Japan) at an accelerating voltage of 20 kV. Calorimetric measurements were conducted on a DSC 6000, Perkin-Elmer, USA differential scanning calorimeter (DSC). 7 mg powder was placed in hermetic sample pan in the DSC cell which was raised from 50 °C to 250 °C (nitrogen atmosphere and heating rate of 10 °C/min). The storage modulus (E′), loss modulus (E″), and damping factor (Tan D) were obtained through a dynamic mechanical analyzer Q800 DMA (1 Hz and the heating rate at 5 °C/min. The weight loss was determined (30–800 °C) by thermogravimetric analysis on TGA Q600 SDT, (TA instruments, SHIMADZU, Kyoto, Japan at a heating rate of 20 °C under nitrogen atmosphere. The tensile and compression strength were measured (ASTM D3039) on the universal testing machine AGX-Plus (SHIMADZU, Kyoto, Japan) test speed of 2 mm/min using a 50 N load.

3. Results and Discussion

Dielectric performance, structural, morphology, thermal properties, and mechanical properties are discussed in this section.

3.1. Dielectric Properties

The reflection of electromagnetic (EM) waves on the surface and the energy loss inside the material is due to the loss of EM waves. Equations (1) and (2) determine the relative dielectric constant (Ɛ) and dielectric loss (δ) of dielectric material toward an electromagnetic field.
ƐI = Ct/ƐoA
δ = C/Coω
where ƐI = dielectric constant, δ = dielectric loss C = capacitance with dielectric, t = sample thickness, Ɛo = permittivity of air (8.85 × 10–12 F/m), A = cross-sectional area of sample, Co = capacitance without dielectric, and ω = angular frequency.
Dielectric permittivity describes how fast an electrical signal can transmit through a dielectric material and a low dielectric constant facilitates signal propagation across it. Dielectric constant and dielectric loss factors (with standard deviation, SD) of thermally cured and microwave-cured composites are mentioned in Table 1 and Table 2, respectively.
The Ɛr and δ of thermally cured composite were 3.90 and 0.053 (S0.3E0.7), 3.89 and 0.054 (S0.4E0.6), 3.80 and 0.050 (S0.5E0.5), 3.80 and 0.051 (S0.6E0.4), and 3.78 and 0.052 (S0.7E0.3). The dielectric properties (Ɛr and δ) decrease as the extent of the cure increases. From the results, both Ɛr and δ were decreased with the increase in cure time and decreased as the reaction progressed; the changes in the dielectric properties are related to the decreasing number of the dipolar groups in the reactants and the increasing viscosity [48].
The Ɛr and δ of microwave-cured composite were 3.81 and 0.046 (S0.3E0.7), 3.80 and 0.045 (S0.4E0.6), 3.77 and 0.043 (S0.5E0.5), 3.79 and 0.043 (S0.6E0.4) and 3.80 and 0.043 (S0.7E0.3). Since microwaves are high-energy waves, they increased the molecular mobility of epoxy molecules and raised the temperate. The rapid curing of epoxy resin was due to the epoxy–amine reaction progress to a greater extent than the epoxy–hydroxyl reaction [32]. Compared to the thermally cured composite, microwave-cured composite exhibited 1% lower Ɛr and 21.5% lower δ. It was due to the fast curing of thermosetting polymer with the progress of the epoxy–amine reaction.
Figure 3 shows the dielectric constant and loss of S0.5E0.5, which was considered an optimum among all proportions. During curing, dielectric properties change due to the disappearance of epoxy, amine groups, and charge migration of dipolar groups [32]. Remarkable dielectric properties of silica fiber (x = 50%) composites were noted (i.e., dielectric loss of microwave cured). Adding silica fiber (x > 50%), the dielectric properties were insignificantly varied; however, decreasing x < 50%, the Ɛr, and δ were marginally increased. Considering the dielectric properties, S0.5E0.5, the structure, morphology, and thermal and mechanical properties were analyzed. Samples were ethanol washed, cleaned, air dried (2 h), and then kept in a desiccator at room temperature until required for testing.

3.2. Structure and Morphology

3.2.1. FTIR Analysis

The FTIR spectral analysis of cured silica fiber/epoxy composites is shown in Figure 4. The spectrum shows characteristic absorption peaks of the epoxide ring between 400 cm−1 and 4000 cm−1. The peak around 916 cm−1 assigned to the C–O deformation of the oxirane ring while the second band located at 1002 cm−1 represents the C–O–C stretching of the epoxy group and another band at 2922 cm−1, which is attributed to the C–H stretching of methylene group in oxirane. Reference peaks around 1631 cm−1 and 1504 cm−1 correspond to the C–C stretching vibration of aromatics, and the C=C stretching vibrations of -CH3, 2922 cm−1 are related to the C–H stretching vibration of CH2 and C–H stretching of –CH3, respectively [27,32]. A hydroxyl linkage is due to the -OH stretching band at 3426 cm−1. There is a decrease in the epoxy ring at 905 cm−1 and shows N–H compression at 1580 cm−1 confirming the reaction of epoxy resin through crosslinking of end epoxy groups with the hardener during curing [38]. The presence of an absorption peak at 1247 cm−1 represents the stretching of the C–N formed by cross-linking of the epoxy ring with an amine group hardener [18]. The peaks at 890 cm−1, 975 cm−1 and 1002 cm−1 are attributed to the presence of Si–OH compression, and Si–O–Si stretching vibrations. The spectra reveal the opening of the epoxide ring by an amine to form OH and CN groups and, conversion of epoxy groups. Comparable IR spectra of thermally and microwave-cured composites are found with similar functional groups, irrespective of the curing route.

3.2.2. Morphology

SEM images in Figure 5 represent the surface morphology, fabric–matrix interaction, and fracture propagation, of S0.5E0.5-T600 (thermally cured). In the thermally cured composite Figure 5a–c, the resin is attached to fiber surfaces; however, fewer fibers are detached. In thermal curing, the heat is transferred through conduction from the outward surface to the inward, and in some portions, there might be different energy available for cure.
The SEM of the S0.5E0.5-M90 composite, as shown in Figure 5d–f, were similar but had a better fiber-matrix interaction than the thermally cured composite. During microwave curing, however, the irradiation and convection result in localized curing (cross-linking) of the thermoset resin. Despite the few voids there, the fibers were seen to be firmly intact with epoxy.

3.3. Thermal Properties

3.3.1. Thermogravimetric Analysis (TGA)

Figure 6 shows the TGA curves where the red line shows a weight loss of thermally cured S0.5E0.5-T600 and the blue line represents microwave cured composite. The degradation was compared to 20% weight loss of the cured composite samples. In the thermally cured composite, degradation temperature was around 535 °C. An abundant weight loss (16.9%) was observed between 285 °C and 538 °C, which refers to pyrolysis and the maximum weight loss, was 27.1%. In the microwave-cured composite (S0.5E0.5-M90) there is a significant mass degradation that begins at 260 °C with a major weight loss (15.4%) from 268 °C to 548 °C. With the final wt. loss of 24.5%, the microwave-cured composite required more degradation energy than thermally cured composites. In a sense, the degradation temperature of microwave cured was slightly higher than that of the thermally cured composite.
Due to high energy radiation, the microwave energy efficiently cured composite than thermal heating. In thermal curing, mainly the heat was transferred to the composite surface, which was reached inside through conduction. Silica fibers are heat-resistance materials in nature that also affect conduction. However, microwave irradiations and convection end in quicker cross-linking of the epoxy in the composite. Due to this (2.6%) lower weight loss, microwave curing can be claimed as a superior route in manufacturing thermally stable composite.

3.3.2. Differential Scanning Calorimeter (DSC)

DSC evaluated the thermal stability and phase transitions in silica fiber/epoxy composites, represented in Figure 7, where an increase in the glass transition temperature (Tg) was observed. Exothermic transitions appeared due to the polymerization of epoxy-amine. In a high-temperature region, more energy was available for the etherification of -OH and epoxy groups, destruction of weak bonds, and homo-polymerization of epoxide rings [26]. DSC curves of S0.5E0.5-T300, S0.5E0.5-T420, and S0.5E0.5-T600 are shown in Figure 7a, where transitions around 80 °C and near 200 °C indicate curing of thermosetting resin conversion of low mol. wt. monomers into a macromolecular cross-linked network through complex transformations [39]. With increased curing time, the height of exothermic transition peaks is more reduced than S0.5E0.5-T300 and S0.5E0.5-T420. The energy evolution is expected due to the possibility of side reactions and homo-polymerization of residual epoxide groups. S0.5E0.5-T600 (10 min cured) is found with the highest degree of increase in Tg, expected to have intermolecular interactions.
DSC of microwave-cured composite, Figure 7b, shows an increasing Tg, where the polymerization is initiated, propagated, and completed rapidly. The thermogram [S0.5E0.5 -M30 and S0.5E0.5 -M60] illustrates transitions around 204 °C to 208 °C, where the composite was cured in a short time as microwave irradiations boosted the reaction rate by more energy penetration and rapid increase in the system’s viscosity [33]. Microwave radiation has a more complex effect on the curing process than the temperature increase. Quick epoxy-amine crosslinking in a short time and microwave energy restricted the mobility of polymer chains [13]. From the thermogram of S0.5E0.5-M90 (cured for 90 s), it looks fully cured, with good dielectric properties, like that epoxy composites with low concentrations of primary amines [36].

3.4. Mechanical Properties

Viscoelastic behavior, ultimate tensile strength (UTS), and ultimate compression strength (UCS) of silica fiber/epoxy composite are discussed in this section.

3.4.1. Dynamic Mechanical Analysis (DMA)

Viscoelastic properties of cured thermosetting polymers are often investigated using DMA as a standard technique suitable for application in a wide temperature range [48]. The thermomechanical spectra (cured composites) were obtained as shown in Figure 8a,b. Storage modulus (E′), loss moduli (E″), and mechanical damping (Tan D) as a function of temperature were presented. Storage moduli evaluate the material’s resistance to deformation while loss modulus quantifies the energy dissipation in the composites [49]. Tg of thermally cured composite (S0.5E0.5-T600) was 72.1 °C, storage modulus started dropping from 51.5 °C, and loss modulus from 64.0 °C.
Glass transition marks the thermal transition between glass and leathery regimes based on the peak of loss modulus [49]. The storage modulus also dropped before Tg; this referred to polymer chain mobility due to the increased thermal energy from the rising temperature. The chains in the epoxy matrix began to slide with higher degrees of freedom than below the first transition, usually referred to as beta transition. An increase in storage modulus, due to a change in Tβ, refers to an increase in the energy dissipation ability of the composite.
DMA of microwave cured S0.5E0.5-M90, in Figure 8b shows 15.5% higher Tg (85.4 °C), and 20% modulus (storage and loss) than thermally cured composite. The EM waves in the microwave oven periodically changed the electric field, which raised the molecular mobility. This molecular motion increased the kinetic energy and the temperature of epoxy resin, where the quick progression of epoxy–amine polymerization caused a good, cross-linked composite. The DMA refers to the fact that the microwave cured offers better resistance against the applied stresses, which revealed superior stability than the thermally cured composite.

3.4.2. Ultimate Tensile Strength (UTS)

Tensile stress–strain curves of thermally and microwave composites are shown in Figure 9. The UTS of S0.5E0.5-T600 was 75.79 MPa and that of S0.5E0.5-M90 was 89.68 MPa. UTS of the thermally cured was 15.4% lower than the microwave-cured composite. From the stress–strain curves in Figure 9, the tough nature of the composite is evident, as the microwave-cured specimen fractured at higher stress. This difference occurred due to the curing method. The strength of the epoxy becomes greater and reaches the maximum upon efficient crosslinking during curing [32]. The higher UTS indicates that the microwave composite structure became comparatively tough compared to the thermally cured composite due to the increased cross-linked composite density.
Composites fracture mechanism describes that epoxy resin transmitted the resistive forces on fibers and fractured as seen in Figure 10. SEM images of the thermally cured composite are shown in the Figure 10a and the microwave-cured composite is shown in Figure 10b. The applied load started cracks from the bonded epoxy resin on the surface, then the fiber–matrix interface, and finally the fibers. This scheme contributed to the final fracture of the composite structure.
The microwave-cured composite broke at a comparatively higher tensile load which can be seen in the Figure 10b that the fracture originated from the area where the fiber-to-matrix bonding was weak. However, a fractured specimen of the microwave-cured specimen showed an even load distribution, which resulted in higher tensile strength, the benefit of microwave curing over thermal heating.

3.4.3. Ultimate Compression Strength (UCS)

In the compression test, the magnitude of opposing forces pushes inward on the specimen. The stress–strain curves showed a linear increase in the compression stress revealing a tough composite structure. The compression strength of S0.5E0.5-T600 (thermally cured) was 201.1 MPa and of S0.5E0.5-M90 (microwave cured) composite, was 210.25 MPa as shown in Figure 11. The compression strength of microwave composite was seen as 4.3% higher than that of the thermally cured composite. Initially, up to 5% strain, the thermally cured specimen offered more resistive forces, but then true stress–strain curves showed similar increasing trends for both samples. Microwave-cured specimens broke at a higher load than thermally cured composite.
Initially, S0.5E0.5-T600 seemed to be tougher than S0.5E0.5-M90 but as the curve progressed, its behavior was slightly changed. The applied force put a load on the surface and then transmitted it inward to the composite structure. In thermal curing, due to conduction, the composite outer surface was better cured than its inner structure. The strength of the epoxy became greater and reached maximum upon efficient crosslinking during curing [32]. Microwave-cured (S0.5E0.5-M90) was found with an improved compact structure than the thermally cured composite, as shown by the results and images.
SEM images of fractured specimens during the compression test, are shown in the Figure 12a,b. The low compression strength of thermally cured composite indicated that the fracture mechanism changed due to the lowering of the fiber–matrix interfacial shear in this specimen. The higher compression strength of S0.5E0.5-M90 revealed that microwave energy increased penetration and better cross-linking. It is clear from the SEM images that the thermally cured structure was more damaged than the microwave-cured composite. These results showed that microwave-assisted localized heating can improve the tensile and compressive strength of fiber-reinforced composites. However, change in curing conditions can improve mechanical properties with further experimentation.

4. Conclusions

The present work compared the effect of the curing route on the properties of silica fiber/epoxy composites. FTIR spectra revealed a similar chemical structure irrespective of the cure mechanism; however, microwave curing took a shorter time and less energy than thermal heating. A compact composite structure was seen where silica fibers were firmly embedded in epoxy resin. Composite cured through microwave energy was obtained with superior dielectric properties (1% lower dielectric constant, 21.5% lower dielectric loss factor), higher thermal stability (2.6% lower wt. loss), and better mechanical properties (higher Tg, storage, loss modulus, 15.4% higher tensile strength, and 4.3% higher compression strength) than thermally cured composite. In microwave heating, high-energy electromagnetic radiations increase the molecular mobility of dipolar epoxy molecules and facilitate efficient curing. The increased kinetic energy of molecules provides a high temperature to put a complex effect on curing compared to conduction in thermal curing. This study suggests that microwave curing is a superior alternative to thermal curing for achieving low dielectric constant, dielectric loss factor, fair thermal stability, and mechanical properties of silica fiber/epoxy composites for microelectronics. Further research is recommended to optimize microwave curing conditions to attain even better results.

Author Contributions

Conceptualization, writing original draft and data curation, I.H., supervision and review, I.H.G.; validation, M.A.U.; validation and editing, M.M.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge the administrative and financial support of School of Chemical and Materials Engineering (SCME) National University of Science and Technology (NUST, H-12 Campus). We also acknowledge Mustafa Anwar for (US-Center of Advance Energy Systems, (US-CAES), NUST for thermo-mechanical analysis, and RIMMS (NUST) for dielectric properties testing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kistaiah, N.; Kiran, C.U.; Reddy, G.R.; Rao, M.S. Mechanical characterization of hybrid composites: A review. J. Reinf. Plast. Compos. 2014, 33, 1364–1372. [Google Scholar] [CrossRef]
  2. Sathishkumar, T.P.; Satheeshkumar, S.; Naveen, J. Glass fiber-reinforced polymer composites—A review. J. Reinf. Plast. Compos. 2014, 33, 1258–1275. [Google Scholar] [CrossRef]
  3. Stickel, J.M.; Nagarajan, M. Glass Fiber-Reinforced Composites: From Formulation to Application. Int. J. Appl. Glass Sci. 2012, 3, 122–136. [Google Scholar] [CrossRef]
  4. Yıldız, S.; Karaağaç, B.; Güzeliş, S.G. Utilization of glass fiber reinforced polymer wastes. Polym. Compos. 2021, 42, 412–423. [Google Scholar] [CrossRef]
  5. Benedetti, A.; Fernandes, P.; Granja, J.L.; Sena-Cruz, J.; Azenha, M. Influence of temperature on the curing of an epoxy adhesive and its influence on bond behaviour of NSM-CFRP systems. Compos. Part B Eng. 2016, 89, 219–229. [Google Scholar] [CrossRef]
  6. Rutz, B.H.; Berg, J.C. Electrostatic deposition of silica nanoparticles between E-glass fibers and an epoxy resin. J. Appl. Polym. Sci. 2015, 132, 41516. [Google Scholar] [CrossRef]
  7. Godara, A.; Gorbatikh, L.; Kalinka, G.; Warrier, A.; Rochez, O.; Mezzo, L.; Luizi, F.; van Vuure, A.W.; Lomov, S.V.; Verpoest, I. Interfacial shear strength of a glass fiber/epoxy bonding in composites modified with carbon nanotubes. Compos. Sci. Technol. 2010, 70, 1346–1352. [Google Scholar] [CrossRef]
  8. Quinson, R.; Perez, J.; Rink, M.; Pavan, A. Components of non-elastic deformation in amorphous glassy polymers. J. Mater. Sci. 1996, 31, 4387–4394. [Google Scholar] [CrossRef]
  9. Budelmann, D.; Schmidt, C.; Meiners, D. Prepreg tack: A review of mechanisms, measurement, and manufacturing implication. Polym. Compos. 2020, 41, 3440–3458. [Google Scholar] [CrossRef]
  10. Wachsman, E.D.; Frank, C.W. Effect of cure history on the morphology of polyimide: Fluorescence spectroscopy as a method for determining the degree of cure. Polymer 1988, 29, 1191–1197. [Google Scholar] [CrossRef]
  11. Uddin, M.A.; Alam, M.O.; Chan, Y.C.; Chan, H.P. Adhesion strength and contact resistance of flip chip on flex packages––effect of curing degree of anisotropic conductive film. Microelectron. Reliab. 2004, 44, 505–514. [Google Scholar] [CrossRef]
  12. Montserrat, S.; Flaqué, C.; Pagès, P.; Málek, J. Effect of the crosslinking degree on curing kinetics of an epoxy–anhydride system. J. Appl. Polym. Sci. 1995, 56, 1413–1421. [Google Scholar] [CrossRef]
  13. Hart-Smith, L.J. Adhesively bonded joints in aircraft structures. In Handbook of Adhesion Technology; De Silva, L.F.M., Öchsner, A., Adams, R.D., Eds.; Springer: Berlin/Heidelberg, Germany, 2011; pp. 1103–1149. [Google Scholar]
  14. So, S.; Rudin, A. Effects of resin and curing parameters on the degree of cure of resole phenolic resins and woodflour composites. J. Appl. Polym. Sci. 1990, 40, 2135–2149. [Google Scholar] [CrossRef]
  15. Benson, R.C.; Miragliotta, J.A.; Farrar, D. Polymer adhesives and encapsulants for microelectronic applications. Johns Hopkins Apl Tech. Dig. 2008, 28, 58–71. [Google Scholar]
  16. Park, E.-T.; Lee, Y.; Kim, J.; Kang, B.-S.; Song, W. Experimental study on microwave-based curing process with thermal expansion pressure of PTFE for manufacturing carbon fiber/epoxy composites. Materials 2019, 12, 3737. [Google Scholar] [CrossRef] [Green Version]
  17. Mishra, R.R.; Sharma, A.K. Microwave–material interaction phenomena: Heating mechanisms, challenges and opportunities in material processing. Compos. Part A Appl. Sci. Manuf. 2016, 81, 78–97. [Google Scholar] [CrossRef]
  18. Liu, L.; Yi, S.; Ong, L.S.; Chian, K.S. Finite element analysis for microwave cure of underfill in flip chip packaging. Thin Solid Film. 2004, 462, 436–445. [Google Scholar] [CrossRef]
  19. Malik, S.; Gul, I.H.; Baig, M.M. Hierarchical MnNiCo ternary metal oxide/graphene nanoplatelets composites as high rated electrode material for supercapacitors. Ceram. Int. 2021, 47, 17008–17014. [Google Scholar] [CrossRef]
  20. Baig, M.M.; Gul, I.H. Transformation of wheat husk to 3D activated carbon/NiCo2S4 frameworks for high-rate asymmetrical supercapacitors. J. Energy Storage 2021, 37, 102477. [Google Scholar] [CrossRef]
  21. Li, N.; Li, Y.; Jelonnek, J.; Link, G.; Gao, J. A new process control method for microwave curing of carbon fibre reinforced composites in aerospace applications. Compos. Part B Eng. 2017, 122, 61–70. [Google Scholar] [CrossRef]
  22. Mphahlele, K.; Ray, S.S.; Kolesnikov, A. Cure kinetics, morphology development, and rheology of a high-performance carbon-fiber-reinforced epoxy composite. Compos. Part B Eng. 2019, 176, 107300. [Google Scholar] [CrossRef]
  23. Baig, M.M.; Gul, I.H.; Ahmad, R.; Baig, S.M.; Khan, M.Z.; Iqbal, N. One-step sonochemical synthesis of NiMn-LDH for supercapacitors and overall water splitting. J. Mater. Sci. 2021, 56, 18636–18649. [Google Scholar] [CrossRef]
  24. Elizbit; Liaqat, U.; Hussain, Z.; Baig, M.M.; Khan, M.A.; Arif, D. Preparation of porous ZIF-67 network interconnected by MWCNTs and decorated with Ag nanoparticles for improved non-enzymatic electrochemical glucose sensing. J. Korean Ceram. Soc. 2021, 58, 598–605. [Google Scholar] [CrossRef]
  25. Joshi, S.C.; Bhudolia, S.K. Microwave-thermal technique for energy and time efficient curing of carbon fiber reinforced polymer prepreg composites. J. Compos. Mater. 2014, 48, 3035–3048. [Google Scholar] [CrossRef]
  26. De Vergara, U.L.; Sarrionandia, M.; Gondra, K.; Aurrekoetxea, J. Polymerization and curing kinetics of furan resins under conventional and microwave heating. Thermochim. Acta 2014, 581, 92–99. [Google Scholar] [CrossRef]
  27. Rangari, V.; Bhuyan, M.; Jeelani, S. Microwave curing of CNFs/EPON-862 nanocomposites and their thermal and mechanical properties. Compos. Part A Appl. Sci. Manuf. 2011, 42, 849–858. [Google Scholar] [CrossRef]
  28. Li, A.; Li, W.; Ling, Y.; Gan, W.; Brady, M.A.; Wang, C. Effects of silica-coated carbon nanotubes on the curing behavior and properties of epoxy composites. RSC Adv. 2016, 6, 23318–23326. [Google Scholar] [CrossRef] [Green Version]
  29. Zhao, B.; Hardiman, M.; Ryan, K.M.; O’Reilly, E.; McCarthy, C. Formation of reworkable nanocomposite adhesives by dielectric heating of epoxy resin embedded Fe3O4 hollow spheres. CrystEngComm 2016, 18, 6096–6101. [Google Scholar] [CrossRef] [Green Version]
  30. Kim, N.D.; Metzger, A.; Hejazi, V.; Li, Y.; Kovalchuk, A.; Lee, S.-K.; Ye, R.; Mann, J.A.; Kittrell, C.; Shahsavari, R. Microwave heating of functionalized graphene nanoribbons in thermoset polymers for wellbore reinforcement. ACS Appl. Mater. Interfaces 2016, 8, 12985–12991. [Google Scholar] [CrossRef]
  31. Alfano, M.; Pagnotta, L. An investigation of the mechanical behaviour of vinyl ester resins cured by microwave irradiation. Mater. Des. 2009, 30, 4537–4542. [Google Scholar] [CrossRef]
  32. Yarlagadda, K.P.; Hsu, S.-H. Experimental studies on comparison of microwave curing and thermal curing of epoxy resins used for alternative mould materials. J. Mater. Process. Technol. 2004, 155, 1532–1538. [Google Scholar] [CrossRef]
  33. Boey, F.; Yap, B. Microwave curing of an epoxy-amine system: Effect of curing agent on the glass-transition temperature. Polym. Test. 2001, 20, 837–845. [Google Scholar] [CrossRef]
  34. Yao, L.; Wang, X.; Liang, F.; Wu, R.; Hu, B.; Feng, Y.; Qiu, Y. Modeling and experimental verification of dielectric constants for three-dimensional woven composites. Compos. Sci. Technol. 2008, 68, 1794–1799. [Google Scholar] [CrossRef]
  35. Mäder, E.; Gao, S.-L.; Plonka, R.; Wang, J. Investigation on adhesion, interphases, and failure behaviour of cyclic butylene terephthalate (CBT®)/glass fiber composites. Compos. Sci. Technol. 2007, 67, 3140–3150. [Google Scholar] [CrossRef] [Green Version]
  36. Ramajo, L.; Reboredo, M.; Castro, M. Dielectric response and relaxation phenomena in composites of epoxy resin with BaTiO3 particles. Compos. Part A Appl. Sci. Manuf. 2005, 36, 1267–1274. [Google Scholar] [CrossRef]
  37. Xie, H.Q.; Zhang, S.; Xie, D. An efficient way to improve the mechanical properties of polypropylene/short glass fiber composites. J. Appl. Polym. Sci. 2005, 96, 1414–1420. [Google Scholar] [CrossRef]
  38. Liu, Z.; Zhang, J.; Tang, L.; Zhou, Y.; Lin, Y.; Wang, R.; Kong, J.; Tang, Y.; Gu, J. Improved wave-transparent performances and enhanced mechanical properties for fluoride-containing PBO precursor modified cyanate ester resins and their PBO fibers/cyanate ester composites. Compos. Part B Eng. 2019, 178, 107466. [Google Scholar] [CrossRef]
  39. Zhu, L.; Wang, Z.; Rahman, M.B.; Shen, W.; Zhu, C. The Curing kinetics of e-glass fiber/epoxy resin prepreg and the bending properties of its products. Materials 2021, 14, 4673. [Google Scholar] [CrossRef]
  40. Fu, H.; Qin, Y.; He, X.; Meng, X.; Zhong, Y.; Zou, Z. Effect of curing degree on mechanical and thermal properties of 2.5 D quartz fiber reinforced boron phenolic composites. e-Polymers 2019, 19, 462–469. [Google Scholar] [CrossRef] [Green Version]
  41. Johnston, K.; Pavuluri, S.; Leonard, M.; Desmulliez, M.; Arrighi, V. Microwave and thermal curing of an epoxy resin for microelectronic applications. Thermochim. Acta 2015, 616, 100–109. [Google Scholar] [CrossRef]
  42. Spoelstra, A.; Peters, G.; Meijer, H. Chemorheology of a highly filled epoxy compound. Polym. Eng. Sci. 1996, 36, 2153–2162. [Google Scholar] [CrossRef] [Green Version]
  43. Boey, F.; Yap, B.; Chia, L. Microwave curing of epoxy-amine system—Effect of curing agent on the rate enhancement. Polym. Test. 1999, 18, 93–109. [Google Scholar] [CrossRef]
  44. Koushyar, H.; Alavi-Soltani, S.; Minaie, B.; Violette, M. Effects of variation in autoclave pressure, temperature, and vacuum-application time on porosity and mechanical properties of a carbon fiber/epoxy composite. J. Compos. Mater. 2012, 46, 1985–2004. [Google Scholar] [CrossRef]
  45. Michels, J.; Cruz, J.S.; Christen, R.; Czaderski, C.; Motavalli, M. Mechanical performance of cold-curing epoxy adhesives after different mixing and curing procedures. Compos. Part B Eng. 2016, 98, 434–443. [Google Scholar] [CrossRef] [Green Version]
  46. Ellis, B. Chemistry and Technology of Epoxy Resins; Springer: Berlin/Heidelberg, Germany, 1993. [Google Scholar]
  47. Endruweit, A.; Johnson, M.; Long, A. Curing of composite components by ultraviolet radiation: A review. Polym. Compos. 2006, 27, 119–128. [Google Scholar] [CrossRef]
  48. Rusmirović, J.; Galović, J.; Kluz, M.; Perković, S.; Brzić, S.; Bogosavljević, M.; Milojković, A.; Kovačević, T. Using potential of filament-wound carbon/glass polymeric composites as rocket motor thermal insulation. Polym. Compos. 2021, 29 (Suppl. S9), S1541–S1554. [Google Scholar] [CrossRef]
  49. Youssef, G.; Pessoa, G.; Nacy, S. Effect of elevated operating temperature on the dynamic mechanical performance of E-glass/epoxy composite. Compos. Part B Eng. 2019, 173, 106937. [Google Scholar] [CrossRef]
Figure 1. Schematic (raw material to prepreg to composite).
Figure 1. Schematic (raw material to prepreg to composite).
Materials 16 01790 g001
Figure 2. Microwave heating mechanism.
Figure 2. Microwave heating mechanism.
Materials 16 01790 g002
Figure 3. (a) Dielectric constant and (b) loss factor of the cured composites.
Figure 3. (a) Dielectric constant and (b) loss factor of the cured composites.
Materials 16 01790 g003
Figure 4. FTIR of the cured composite.
Figure 4. FTIR of the cured composite.
Materials 16 01790 g004
Figure 5. Morphology of (ac) thermally and microwave (df) cured composites.
Figure 5. Morphology of (ac) thermally and microwave (df) cured composites.
Materials 16 01790 g005
Figure 6. TGA of thermally and microwave-cured composite.
Figure 6. TGA of thermally and microwave-cured composite.
Materials 16 01790 g006
Figure 7. DSC of (a) thermally and (b) microwave-cured composite.
Figure 7. DSC of (a) thermally and (b) microwave-cured composite.
Materials 16 01790 g007
Figure 8. DMA plot (a): thermally cured composite. (b): microwave cured composite.
Figure 8. DMA plot (a): thermally cured composite. (b): microwave cured composite.
Materials 16 01790 g008
Figure 9. Tensile strength of cured composites.
Figure 9. Tensile strength of cured composites.
Materials 16 01790 g009
Figure 10. Fractured in tensile test (a) thermally cured (b) microwave cured.
Figure 10. Fractured in tensile test (a) thermally cured (b) microwave cured.
Materials 16 01790 g010
Figure 11. Compression strength of cured composites.
Figure 11. Compression strength of cured composites.
Materials 16 01790 g011
Figure 12. Compression fractured (a) thermally cured (b) microwave cured.
Figure 12. Compression fractured (a) thermally cured (b) microwave cured.
Materials 16 01790 g012
Table 1. Dielectric properties of thermally cured composite at 160 °C.
Table 1. Dielectric properties of thermally cured composite at 160 °C.
tcS0.3E0.7S0.4E0.6S0.5E0.5S0.6E0.4S0.7E0.3
(sec)DkLossDkLossDkLOSSDkLossDkLoss
2403.980.0593.980.0573.960.0573.960.0573.940.057
SD0.0060.0010.0100.0020.02920.0010.0290.0010.02480.002
3003.960.0563.970.0563.950.0553.950.0543.910.056
SD0.0100.0010.0100.0010.0100.0010.0310.0010.0020.001
3603.940.0563.950.0533.850.0533.890.0533.890.054
SD0.0100.0010.0100.0020.0480.0020.0390.0010.0270.002
4203.910.0533.910.0553.830.0533.850.0533.830.053
SD0.0080.0010.0080.0020.0010.0510.0010.0010.0490.001
4803.910.0523.890.0523.810.0513.820.0523.790.052
SD0.0140.0010.0190.0010.0010.0010.0100.0010.0250.002
6003.900.0533.890.0543.800.0503.820.0513.780.052
SD0.0140.0010.0150.0010.0610.0010.0010.0010.0140.001
6603.900.0533.980.0543.790.0513.800.0513.940.052
SD0.0180.0010.0020.0120.0020.0080.0020.0010.0100.001
Table 2. Dielectric properties of microwave-cured composite at 300 W.
Table 2. Dielectric properties of microwave-cured composite at 300 W.
tcS0.3E0.7S0.4E0.6S0.5E0.5S0.6E0.4S0.7E0.3
(sec)DkLossDkLossDkLossDklossDkLoss
203.980.0503.960.0513.920.0483.940.0483.940.048
SD0.0230.0010.0120.0050.0050.0020.0080.0020.0090.003
303.890.0493.890.0493.890.0463.900.0473.930.046
SD0.0420.0010.0270.0030.0200.0030.0200.0010.0240.002
453.860.0483.850.0483.850.0453.860.0453.880.046
SD0.0360.0010.0130.0010.0430.0040.0290.0020.0340.002
603.840.0483.820.0463.820.0443.820.0433.840.043
SD0.0260.0020.0260.0010.0260.000.0090.0020.0340.002
753.800.0463.790.0453.780.0433.790.0423.800.043
SD0.0170.0010.0300.0020.0530.0040.0360.0020.0450.001
903.810.0463.800.0453.770.0433.790.0433.800.043
SD0.0410.0030.0260.0010.0600.0020.0230.0020.0280.003
1003.800.0463.800.0513.780.0433.790.0413.800.042
SD0.0450.0040.0380.0020.0330.0030.0270.0020.0460.002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Haider, I.; Gul, I.H.; Umer, M.A.; Baig, M.M. Silica-Fiber-Reinforced Composites for Microelectronic Applications: Effects of Curing Routes. Materials 2023, 16, 1790. https://doi.org/10.3390/ma16051790

AMA Style

Haider I, Gul IH, Umer MA, Baig MM. Silica-Fiber-Reinforced Composites for Microelectronic Applications: Effects of Curing Routes. Materials. 2023; 16(5):1790. https://doi.org/10.3390/ma16051790

Chicago/Turabian Style

Haider, Imran, Iftikhar Hussain Gul, Malik Adeel Umer, and Mutawara Mahmood Baig. 2023. "Silica-Fiber-Reinforced Composites for Microelectronic Applications: Effects of Curing Routes" Materials 16, no. 5: 1790. https://doi.org/10.3390/ma16051790

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop