[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (2)

Search Parameters:
Keywords = polyimide-nanographene sheets composites

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 3920 KiB  
Article
Influence of the Processing Conditions on the Rheology and Heat of Decomposition of Solution Processed Hybrid Nanocomposites and Implication to Sustainable Energy Storage
by Andekuba Andezai and Jude O. Iroh
Energies 2024, 17(16), 3930; https://doi.org/10.3390/en17163930 - 8 Aug 2024
Viewed by 901
Abstract
This study investigates the properties of solution-processed hybrid polyimide (PI) nanocomposites containing a variety of nanofillers, including polyaniline copolymer-modified clay (PNEA), nanographene sheets (NGSs), and carbon nanotube sheets (CNT-PVDFs). Through a series of experiments, the flow behavior of poly(amic acid) (PAA) solution and [...] Read more.
This study investigates the properties of solution-processed hybrid polyimide (PI) nanocomposites containing a variety of nanofillers, including polyaniline copolymer-modified clay (PNEA), nanographene sheets (NGSs), and carbon nanotube sheets (CNT-PVDFs). Through a series of experiments, the flow behavior of poly(amic acid) (PAA) solution and PAA suspension containing polyaniline copolymer-modified clay (PNEA) is determined as a function of the shear rate, processing temperature, and polymerization time. It is shown that the neat PAA solution exhibits a complex rheological behavior ranging from shear thickening to Newtonian behavior with increasing shear rate and testing temperature. The presence of modified clay in PAA solution significantly reduced the viscosity of PAA. Differential scanning calorimetry (DSC) analysis showed that polyimide–nanographene sheet (PI NGS) nanocomposites processed at a high spindle speed (100 rpm) have lower total heat of decomposition, which is indicative of improved fire retardancy. The effect of processing temperature on the specific capacitance of a polyimide–CNT-PVDF composite containing electrodeposited polypyrrole is determined using cyclic voltammetry (CV). It is shown that the hybrid composite working electrode material processed at 90 °C produces a remarkably higher overall stored charge when compared to the composite electrode material processed at 250 °C. Consequently, the specific capacitance obtained at a scan rate of 5 mV/s for the hybrid nanocomposite processed at 90 °C is around 858 F/g after one cycle, which is about 6.3 times higher than the specific capacitance of 136 F/g produced by the hybrid nanocomposite processed at 250 °C. These findings show that the properties of the hybrid nanocomposites are remarkably influenced by the processing conditions and highlight the need for process optimization. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of in situ synthesis of (<b>a</b>) PI/clay nanocomposite, (<b>b</b>) PI/graphene nanocomposite, and (<b>c</b>) solution casting of PAA/CNT-PVDF and (<b>d</b>) electrochemical synthesis of polypyrrole.</p>
Full article ">Figure 2
<p>Dependence of viscosity on the shear rate and temperature for neat PAA solution. The arrows mark the critical shear rates for transition from shear thickening to Newtonian behavior.</p>
Full article ">Figure 3
<p>Effect of temperature on the critical shear rate <math display="inline"><semantics> <mover accent="true"> <mi>γ</mi> <mo>˙</mo> </mover> </semantics></math><sub>c</sub> and steady-state viscosity for the neat poly(amic acid) solution.</p>
Full article ">Figure 4
<p>Effect of shear rate and temperature on the viscosity of poly(amic acid) suspension containing 5 wt.% of polyaniline copolymer-modified (PNEA) Cloisite 30B clay. The arrow marks the critical shear rate for transition from shear thickening to Newtonian behavior.</p>
Full article ">Figure 5
<p>Effect of shear rate and polymerization time on the viscosity of poly(amic acid) after 30 min and 24 h of polymerization. The arrows mark the critical shear rates for transition from shear thickening to Newtonian behavior.</p>
Full article ">Figure 6
<p>Plot of ln viscosity (cP) vs. inverse temperature <math display="inline"><semantics> <mrow> <mfenced> <mrow> <msup> <mi>K</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </mfenced> </mrow> </semantics></math> at a spindle speed rate of 20 rpm for (i) neat PAA sample and (ii) PAA suspension containing 5 wt.% PNEA.</p>
Full article ">Figure 7
<p>Plot of (i) inherent viscosity and (ii) reduced viscosity against concentration for (<b>a</b>) PAA solution and (<b>b</b>) PAA suspension containing 5 wt.% PNEA modified clay.</p>
Full article ">Figure 8
<p>(<b>a</b>) DSC thermograms of (i) neat PI, (ii) PI-10 wt.% nanographene sheet sheared at 100 rpm for 30 min, and (iii) PI-10 wt.% nanographene sheet and (<b>b</b>) DSC curves of (i) neat PI, (ii) PI-40 wt.% nanographene sheet cast after shearing the suspension at 100 rpm for 30 min, and (iii) PI-40 wt.% nanographene sheet cast without additional shearing of the suspension. The DSC test was conducted under a nitrogen atmosphere at a heating rate of 5 °C/min.</p>
Full article ">Figure 9
<p>Cyclic voltammograms of PI/CNT-PVDF composites processed at (i) 90 °C and (ii) 250 °C followed by electrodeposition of PPy using a Ag/AgCl reference electrode and a graphite rod counter electrode for (<b>a</b>) 1 cycle and (<b>b</b>) 10 cycles; (<b>c</b>) composites processed at 90 °C for (i) 1 cycle and (ii) 10 cycles; and (<b>d</b>) composites processed at 250 °C for (i) 1 cycle and (ii) 10 cycles. CV was run at 5 mV/s.</p>
Full article ">Figure 10
<p>Cyclic voltammograms of PI/CNT-PVDF composites processed at (i) 90 °C and (ii) 250 °C followed by electrodeposition of PPy using a Ag/AgCl reference electrode and graphite rod counter electrode for (<b>a</b>) 1 cycle and (<b>b</b>) 10 cycles; (<b>c</b>) composites processed at 90 °C for (i) 1 cycle and (ii) 10 cycles; and (<b>d</b>) composites processed at 250 °C for (i) 1 cycle and (ii) 10 cycles. CV was run at 25 mV/s.</p>
Full article ">Figure 11
<p>Transient i–t curves obtained during potentiostatic polymerization of 0.5 M pyrrole in a 0.0225 M toluene sulphonic acid solution at an applied potential of 2 V onto PI/CNT-PVDF composite working electrodes processed at (i) 90 °C and (ii) 250 °C.</p>
Full article ">Scheme 1
<p>Synthesis of polyimide (PI).</p>
Full article ">
14 pages, 4462 KiB  
Article
Decomposition and Flammability of Polyimide Graphene Composites
by Caroline Akinyi, Jimmy Longun, Siqi Chen and Jude O. Iroh
Minerals 2021, 11(2), 168; https://doi.org/10.3390/min11020168 - 5 Feb 2021
Cited by 11 | Viewed by 2741
Abstract
Polyimide-graphene composites were synthesized by in-situ condensation polymerization and the thermal stability and decomposition behavior of the composites were studied. Polyimides, because of their aromatic backbone, are a class of fire-retardant polymers. Their high char retention ≥50% at testing temperatures ≥600 °C makes [...] Read more.
Polyimide-graphene composites were synthesized by in-situ condensation polymerization and the thermal stability and decomposition behavior of the composites were studied. Polyimides, because of their aromatic backbone, are a class of fire-retardant polymers. Their high char retention ≥50% at testing temperatures ≥600 °C makes them thermally stable polymers. The effect of nanographene sheets on the decomposition behavior of polyimide is presented in this paper. It is shown that the reinforcement of polyimide with nanographene sheets significantly decreased the rate of decomposition of polyimide and increased the char retention of the composite. Thermogravimetric analysis data were used to assess the thermal stability, rate of mass loss and predicted limiting oxygen index of the neat polyimide and composites. Results obtained showed around a 43% decrease in the rate of polyimide degradation at 50 wt.% graphene loading. The limiting oxygen index of the polyimide nanocomposite was calculated by using the char retention, and it was found to increase by up to 24% at 50 wt.% graphene loading over that for the neat matrix. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Raman spectra of (a) neat polyimide (PI) and PI containing (b) 1 wt.% (c) 10 wt.% and (d) 30 wt.% graphene.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) Neat PI, (<b>b</b>) Nano-graphene sheets and (<b>c</b>) PINGS-10 (×25,000).</p>
Full article ">Figure 3
<p>SEM micrographs showing cross-sectional morphology of (<b>a</b>) PINGS-30 and (<b>b</b>) PINGS-40.</p>
Full article ">Figure 4
<p>Thermogravimetric mass loss curves of (I) Graphene, (II) PINGS-50, (III) PINGS-30, (IV) PINGS-10 and (V) Neat PI (in nitrogen; β = 30 °C/min).</p>
Full article ">Figure 5
<p>Degradation mechanism of polyimide in nitrogen.</p>
Full article ">Figure 6
<p>Predicted limiting oxygen index (LOI) versus composition for neat PI and composites containing 10, 20, 30, 40 and 50 wt.% graphene based on Equation (2) ([<a href="#B28-minerals-11-00168" class="html-bibr">28</a>]).</p>
Full article ">Figure 7
<p>Derivative mass loss curves (I) PI, (II) PINGS -10, (III) PINGS-30, (IV) PINGS-50 and (V) Graphene (in nitrogen; β = 30 °C/min).</p>
Full article ">Figure 8
<p>Rate of degradation of the polymer matrix as a function of the percentage of graphene in the nanocomposites (in nitrogen; β = 30 °C/min).</p>
Full article ">Figure 9
<p>(<b>a</b>) Thermogravimetric and derivative mass loss curve for PI composite containing 50 wt.% graphene showing (I) the PI matrix degradation peak, (II) the PI char/graphene char degradation peak, (III) the graphene char degradation peak (in air; β = 30 °C/min). (<b>b</b>) Thermogravimetric mass loss curves of (I) Neat PI (II) Graphene (III) PINGS-10 (IV) PINGS-30 (V) PINGS-50 (in air; β = 30 °C/min).</p>
Full article ">Figure 10
<p>(<b>a</b>) TGA and derivative TGA mass loss curves in air and in nitrogen for PINGS (<b>b</b>) Derivative TGA mass loss curves (I) PI, (II) PINGS -50, and (III) Graphene (in air; β = 30 °C/min).</p>
Full article ">Figure 11
<p>Rate of char degradation as a function of graphene composition in the nanocomposites Graphene (in air; β = 30 °C/min).</p>
Full article ">Figure 12
<p>Kissinger-Akahira-Sunose (KAS) plots for PINGS-10 and PINGS-30.</p>
Full article ">Figure 13
<p>Activation energy of imidization and polymer degradation as a function of composition of graphene in the nanocomposites.</p>
Full article ">
Back to TopTop