Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Journal of Natural Fibers

ISSN: (Print) (Online) Journal homepage: www.tandfonline.com/journals/wjnf20

Mechanical, Thermal Properties and Void


Characteristics of Bamboo Fiber-Reinforced Epoxy
Resin Composites Prepared by Vacuum-Assisted
Resin Transfer Molding Process

Jiangjing Shi, Wenfu Zhang, Jian Zhang, Shaofei Yuan & Hong Chen

To cite this article: Jiangjing Shi, Wenfu Zhang, Jian Zhang, Shaofei Yuan & Hong Chen (2023)
Mechanical, Thermal Properties and Void Characteristics of Bamboo Fiber-Reinforced Epoxy
Resin Composites Prepared by Vacuum-Assisted Resin Transfer Molding Process, Journal of
Natural Fibers, 20:1, 2187919, DOI: 10.1080/15440478.2023.2187919

To link to this article: https://doi.org/10.1080/15440478.2023.2187919

© 2023 The Author(s). Published with


license by Taylor & Francis Group, LLC.

Published online: 17 Mar 2023.

Submit your article to this journal

Article views: 723

View related articles

View Crossmark data

Citing articles: 3 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=wjnf20
JOURNAL OF NATURAL FIBERS
2023, VOL. 20, NO. 1, 2187919
https://doi.org/10.1080/15440478.2023.2187919

Mechanical, Thermal Properties and Void Characteristics of Bamboo


Fiber-Reinforced Epoxy Resin Composites Prepared by
Vacuum-Assisted Resin Transfer Molding Process
Jiangjing Shia,b,c, Wenfu Zhanga,c, Jian Zhanga,c, Shaofei Yuana,c, and Hong Chenb
a
Zhejiang Academy of Forestry, Hangzhou, China; bCollege of Furnishings and Industrial Design, Nanjing Forestry
University, Nanjing, China; cKey Laboratory of Bamboo Research of Zhejiang Province, Hangzhou, China

ABSTRACT KEYWORDS
This study aimed to examine the effect of epoxy resin (EP) systems on the Bamboo fiber; epoxy resin;
impregnation of bamboo fibers (BFs) by EP and the properties of BF/EP vacuum-assisted resin
composites prepared by vacuum-assisted resin transfer molding (VARTM) transfer molding; void
process. BF preforms obtained by solution suspension method were used characteristic
as reinforcement, and epoxy/anhydride system (EP-1), epoxy/amine system 关键词
(EP-2) and two-component epoxy system (EP-3) were employed as matrix; on 竹纤维; 环氧树脂;
this basis, BF/EP-1, BF/EP-2 and BF/EP-3 composites were prepared by 真空辅助树脂传递模塑成
VARTM. The physical, mechanical and thermal properties and void character­ 型;
istics of BF/EP composites were characterized. The results showed that EP-1 空隙特性
with a viscosity of 265 mPa·s at 23°C and an operating time of room
temperature without curing was beneficial for the EP-1 to impregnate BFs
adequately and uniformly. The incorporation of BFs into EP-1 significantly
improved the flexural, shear and impact properties, thermal properties and
interface properties of EP-1. Moreover, the properties of BF/EP-1 composites
were better than those of BF/EP-2 and BF/EP-3 composites. Three-
dimensional X-ray microscopy scans revealed that the volume and distribu­
tion of voids in the BF/EP composites differed significantly depending on the
EP systems, and the percentage of void volume in the BF/EP-1 composites
was only 0.14% in 3D reconstruction models.
摘要
本研究旨在考察环氧树脂 (EP) 体系对EP浸渍竹纤维 (BFs) 以及真空辅助树脂
传递模塑 (VARTM) 工艺制备的BF/EP复合材料性能的影响. 采用溶液悬浮法
制备的BF预成型体作为增强体, 环氧/酸酐体系(EP-1)、环氧/胺体系(EP-2)和
双组分环氧体系(EP-3)作为基体; 在此基础上, 利用VARTM制备了BF/EP-1、
BF/EP-2和BF/EP-3复合材料. 对BF/EP复合材料的物理、力学、耐热性能和孔
隙特征进行了表征. 结果表明, EP-1在23°C下的粘度为265mPa·s, 操作时间为
室温下不固化的EP-1有利于EP-1充分、均匀地浸渍BFs. BFs的加入显著改善
了EP-1的弯曲、剪切和冲击性能、耐热性能和界面性能. 此外, BF/EP-1复合
材料的性能优于BF/EP-2和BF/EP-3复合材料. 三维X射线显微镜扫描显示, BF/
EP复合材料中孔隙的体积和分布因EP体系的不同而显著不同, 在3D重建模
型中, BF/EP-1复合材料中的孔隙体积百分比仅为0.14%.

Introduction
With the worsening global energy crisis and ecological environment, widely sourced, recyclable
and degradable plant fibers of high performance are replacing nonrenewable and hard-to-
degrade artificial fibers (carbon fibers, glass fibers, aramid fibers, basalt fibers, etc.) as the

CONTACT Shaofei Yuan fei20008281@126.com Key Laboratory of Bamboo Research of Zhejiang Province,
ZhejiangAcademy of Forestry, Hangzhou, China
© 2023 The Author(s). Published with license by Taylor & Francis Group, LLC.
This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0/),
which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. The terms on which this
article has been published allow the posting of the Accepted Manuscript in a repository by the author(s) or with their consent.
2 J. SHI ET AL.

reinforcement of fiber-reinforced polymer composites (FRPs) (Chokshi et al. 2020; Furtos et al.
2012). Jute fibers, hemp fibers, flax fibers, wood fibers and bamboo fibers (BFs) have been used
as the common reinforcement of plant fiber-reinforced polymer composites (PFRPs) (Furtos
et al. 2021, 2022). BFs, known as “natural glass fiber,” have become a research hotspot of PFRPs
owing to their low density, high strength and high rigidity (Hu et al. 2019). BF-reinforced
polymer composites (BFRPs) are used in green packaging materials, automotive molded linings,
machine product housings and home decoration (Muhammad et al. 2018). Presently, BFRPs are
mainly prepared by compression molding, extrusion molding, filament winding molding and
liquid composite molding. BFRPs with thermoset resins as matrix are usually prepared by liquid
composite molding, including autoclave molding, resin transfer molding (RTM) and vacuum-
assisted resin infiltration molding (VARI) (Chang and Chang 2022). Vacuum-assisted resin
transfer molding (VARTM) was developed from RTM, which used vacuum pressure to drive
resins to impregnate BF preforms and then cured the whole material at room temperature or
under heated conditions so as to prepare BFRPs with a certain BF content (Wang et al. 2019).
BFRPs prepared by VARTM had fewer voids and greater properties than that by other molding
techniques (Xia et al. 2015).
The current research related to BFRP preparation by VARTM mainly focused on the properties of
BFs and the process parameters of VARTM. Studies on BFs were related to BF content (Yang et al.
2019), BF morphology (Rocky and Thompson 2018), BF surface modification (Chen et al. 2021; Kim
et al. 2013; Wang et al. 2020) and BF preform preparation (Chiu and Young 2020). BF preforms were
often prepared by hand lay-up, needle-punched process, textile process and twisting process. Among
them, hand lay-up was convenient and efficient but might lead to unevenly dispersed BFs because of
the unsatisfying skill and proficiency of operators. In contrast, the needle-punched process, textile
process and twisting process could be used to prepare uniformly dispersed BF preforms but might
easily cause secondary damage to BFs. Therefore, it is necessary to explore a suitable method of BF
preform preparation for VARTM. The properties of BFRPs could be affected by numerous process
parameters, such as vacuum level, injection method, injection time, injection temperature, the
existence of diversion network, permeability and pressure. These parameters were interrelated and
interplayed, jointly determining the properties of final BFRPs (Kedari, Farah, and Hsiao 2011; Li, Xie,
and Ma 2015).
BFRP preparation by VARTM was highly dependent on vacuum-driven resin flow to
impregnate BFs compared to other molding processes, making resin systems a critical part.
Different viscosity, operating times, and curing processes of various resin systems had
a significant impact on BF impregnation by resins and properties of BFRPs (Yang et al.
2011). BF impregnation by resins could be reflected by the void content in BFRPs and the
properties of BFRPs. The voids were mainly flow-induced and gas-induced bubbles in BFRPs,
as well as gaps at the BF-resin interface primarily determined by the physical and chemical
properties of the two materials and the VARTM process (Sul, Youn, and Song 2012). However,
at present, resin systems for BFRP preparation by VARTM tend to borrow those used for the
preparation of other PFRPs or artificial FRPs by VARTM. For BFs with unique structures and
properties, the effect of different resin systems on the properties of VARTM-prepared BFRPs is
worth further examination.
In this study, BFs were prepared into BF preforms by solution suspension method to achieve
nondestructive lay-up and uniformity. BFs were used as reinforcement. Epoxy/anhydride system (EP-
1), epoxy/amine system (EP-2) and two-component epoxy system (EP-3) were used as the matrix. BF/
EP-1, BF/EP-2 and BF/EP-3 composites were prepared by VARTM. The physical, mechanical and
thermal properties and void characteristics of BF/EP composites were characterized to investigate the
effect of EP systems on BF impregnation by EP and the properties of BF/EP composites. The research
and optimization of VARTM process parameters are of great significance to the practical application
of BFRP products.
JOURNAL OF NATURAL FIBERS 3

Materials and methods


Materials
BFs (Dendrocalamopsis oldhami (Munro) Keng f.), with a mean length of 2.54 cm, a mean diameter of
0.25 mm and an aspect ratio of 102 were obtained from Fujian Haibosi Chemical Technology Co., Ltd.,
China. The matrix was three epoxy resin (EP) systems: EP-1, EP-2 and EP-3. EP-1 was prepared by
epoxy resin (NPEL127E), methyl tetrahydrophthalic anhydride curing agent (WNY1008) and N,
N-dimethylbenzylamine accelerator with a ratio of 100:95:1. The epoxy resin (viscosity: 10255
MPa·s at 25°C, epoxy equivalent: 182.4 g/Eq) was purchased from Nanya electronic materials
(Kunshan) Co., Ltd., China. The methyl tetrahydrophthalic anhydride curing agent (viscosity: 36
MPa·s at 25°C) was supplied by Dongying Xichen Chemical Co., Ltd., China. N,
N-dimethylbenzylamine accelerator (C6H5CH2N(CH3)2) was obtained from Hangzhou Lanbo
Industrial Co., Ltd., China. EP-2 was prepared by epoxy resin (NPEL127E) and aromatic amine
curing agent (AL-90) with a ratio of 100:56. The aromatic amine curing agent (viscosity: 128 MPa·s
at 18°C) was supplied by Dongying Xichen Chemical Co., Ltd., China. EP-3 was prepared by epoxy
resin group A (Q-302) and epoxy resin group B (Q-302) at a ratio of 3:1. Both epoxy resin group
A (viscosity: 600–800 MPa·s at 25°C) and epoxy resin group B (viscosity: 40–60 MPa·s at 25°C) were
purchased from Suzhou Wuseshi Chemical Co., Ltd., China.

Preparation of bamboo fiber preforms and composites


BF preforms were prepared by the solution suspension method (Figure 1a). First, BFs were selected to
remove impurities and undispersed and lumpy BFs from raw material. Then BFs were evenly scattered
in a self-made sieve with the same shape as the mold. The sieve was placed into the suspension device
and vibrated for even suspension and dispersal of BF in water. Finally, BF preforms were taken out and
dried. The moisture content of BF preforms was controlled to be below 2%.
The process flow of preparing BF/EP composites with different EP systems by VARTM is
shown in Figure 1b. First, after the mold was cleaned, valves and hoses were installed on its left
and right sides, and three layers of release agent were brushed inside it. Then, BF preforms, peel
ply and diversion network were laid in turn. Finally, the sealing tape was applied to the surface of
the female mold, and the vacuum pump was turned on to check the sealing of the mold after
closing the mold. According to the ratio of each of EP systems, raw materials were weighed to

Figure 1. (A) Schematic diagram of BF preforms prepared by solution suspension method. (b) Schematic diagram of BF/EP
composites prepared by VARTM process.
4 J. SHI ET AL.

Table 1. Process parameters of BF/EP composites with different EP systems.


Composite Reinforcement Matrix Operation time Curing process
BF/EP-1 Bamboo fiber Epoxy/anhydride system (EP-1) No curing at 75°C for 1 h, 100°C for 1 h and 130°C for
(BF) room temperature 2h
BF/EP-2 Epoxy/amine system 30 min Room temperature for 24 h
(EP-2)
BF/EP-3 Two-component epoxy system 20 min Room temperature for 24 h
(EP-3)
Notes: The mass fraction of BFs in BF/EP composites was 45 wt%.

prepare the matrix. EP injection was carried out by boundary injection and boundary output with
a vacuum level of 0.080–0.085 MPa. When EP was sucked out of output ports, it continued to be
injected for 20 min so that the air inside the mold was pumped out as much as possible. The valves
on both sides of the mold were closed after EP injection, and BF/EP composites were cured
according to the curing process in Table 1.

Characterization
Physical property tests
The viscosity of the EP systems was tested according to (GB/T 22314–2008) by a rotary viscometer
(NDJ-1, Shanghai Changji Geological Instruments Co., Ltd, China). The density of samples was
measured based on the geometric method in GB/T 1463–2005. The water absorption of samples
was measured according to GB/T 1462–2005: Samples were immersed in boiling water for 30 min and
cooled in distilled water at room temperature for 15 min. Five samples of each type of composites were
selected for each physical property test.

Mechanical property tests


The flexural, shear and impact properties of the samples were investigated according to (GB/T 1449–
2005), (ASTM D2344/D2344M–16, 2016) and (GB/T 1451–2005), respectively. Both flexural and
shear tests were performed with a mechanical testing machine (5582, Instron, USA). The impact
test was conducted with an impact experimental machine (XJJ-5, Chengde Kesheng Testing Machine
Co., Ltd., China). Five samples of each type of composites were selected for each mechanical property
test.
2.3.3 Microtopography analysis
The section images and fracture morphology of samples after the flexural test were observed with
a scanning electron microscope (SEM, Quanta 200, FEI, USA).

Thermal property analysis


The thermal stability of samples was analyzed by a thermogravimetric analyzer (TGA, STA409PC,
Netzsch, Germany). All measurements were conducted under a nitrogen atmosphere and heated up
from 25°C to 800°C at a constant heating rate of 20°C/min. The dynamic mechanical properties of
samples were measured in the dual-cantilever mode by a dynamic mechanical analyzer (DMA, Q800,
TA instruments Inc., USA). Tests were conducted at temperatures ranging from 30°C to 120°C at
a constant heating rate of 3°C/min with an amplitude of 15 µm and a frequency of 1 Hz.

Void characterization
The samples were scanned by a three-dimensional (3D) X-ray microscope (Micro-CT, nanoVoxel
3000, Tianjin Sanying Precision Instrument Co., Ltd., China) to observe the impregnation effect of
different EP systems on BFs and the void characteristics of samples with a tube voltage of 80 kV, a tube
current of 50 μA, a scanning time of 2 h, and a resolution of 6.60 μm. A threshold segmentation
method was used to extract voids in samples to reconstruct 3D models.
JOURNAL OF NATURAL FIBERS 5

Results and discussion


Physical properties
The viscosities of different EP systems EP-1, EP-2 and EP-3 were 265, 275 and 450 MPa·s at 23°C,
respectively (Figure 2a). During BF/EP composite preparation with consistent VARTM process
parameters, the higher the viscosity of EP systems, the more challenging the EP injection, making
no contributions to pumping out air from the mold. As can be seen in Figure 2b, the densities of BF/
EP-1 and BF/EP-2 composites were higher than those of EP-1 and EP-2, while the density of BF/EP-3
composites was lower than that of EP-3. During BF/EP composite preparation by VARTM, the density
of BF/EP composites was related to BF and EP system density and was determined by the impregna­
tion effect of EP on BFs. When the BF content of BF/EP composites was identical, EP-1 and EP-2 with
a low viscosity were beneficial to EP flow to every corner in the mold during molding, significantly
reducing voids in BF/EP composites, thus increasing the density of BF/EP-1 and BF/EP-2 composites.
The section images of BF/EP composites are presented in Figure 2d. Compared with BF/EP-1 and BF/
EP-2 composites, BF/EP-3 composites had significant voids, which were mainly due to the difficulty of
flowing EP-3 with high viscosity during molding.
BF/EP composites could absorb water through three primary pathways: BF, EP and BF/EP interface
(Jena, Pradhan, and Pandit 2014). As shown in Figure 2c, EP-2 presented higher water absorption than
the other neat EP samples, mostly attributed to the component composition of EP systems. When
water entered the interior of BF/EP composites through EP, it would fill BF/EP composites along BF/
EP interface by capillary action. BFs could absorb water through physical and chemical ways. Physical
water absorption was mainly because of the certain water retention capacity of the BF lumen, while
chemical water absorption was mainly due to a large number of hydroxyl groups on the surface of
cellulose in BFs (Depuydt et al. 2019). These groups were extremely easy to contact water molecules
and form hydrogen bonds, resulting in strong water absorption of BFs. The water absorption of
different EP systems greatly increased by incorporating BFs. In this study. BF/EP-1 composites showed
the lowest water absorption among the three BF/EP composites, which was because EP-1 with low
viscosity and long operation time impregnated BFs adequately and effectively, BF/EP-1 composites
had fewer voids, and EP-1 could wrap around BFs to form a barrier. The shorter operation time of EP-

Figure 2. (A) Viscosity of EP systems. BF/EP composites with different EP systems: (b) density, (c) water absorption and (d) section
images.
6 J. SHI ET AL.

2 and the higher viscosity of EP-3 tended to lead to inadequate and poorer impregnation of BFs with
EP, resulting in higher and different water absorption of BF/EP-2 and BF/EP-3 composites.

Mechanical properties
As shown in Figure 3a, the incorporated BFs into three EP systems enhanced the flexural strength of
EP-1 but slightly decreased that of EP-2 and EP-3. This was mainly because when the BF content of
BF/EP composites was consistent, the BF impregnation by EP had a greater influence on the flexural
strength of BF/EP composites. EP-1 with low viscosity and long operation time was beneficial to the
void reduction of BF/EP composites and improvement of the interfacial bonding between BFs and EP.
In contrast, the BF impregnation of EP-2 and EP-3 was inferior to that of EP-1. BF incorporation into
EP showed a significant increase in the flexural modulus of different EP systems, with the largest
enhancement in BF/EP-1 composites by 98% compared to that of EP-1. The fracture morphology of
BFs and the interfacial bonding between BFs and EP after the flexural property test of BF/EP
composites are shown in Figure 3d and e. As can be seen, basically no obvious gaps existed between
BFs and EP-1, and the interfacial bonding between BFs and EP-1 was well. BFs with flat ends were not
pulled out when damaged by an external load, and they could play an obvious strengthening role. BFs
in BF/EP-2 and BF/EP-3 composites had obvious gaps with EP. BFs were pulled out when damaged by
an external load, with the ends torn. The interfacial properties of BFs and EP were weak.
The shear strength of BF/EP composites with different EP systems is shown in Figure 3b, presenting
a similar trend to that of flexural strength. Only the shear strength of BF-incorporated EP-1 was

Figure 3. The effect of EP systems on the mechanical properties of BF/EP composites: (a) flexural strength/modulus, (b) shear
strength and (c) impact toughness. (d) Fracture morphologies of BFs. (e) Interfacial bonding of BFs and EP.
JOURNAL OF NATURAL FIBERS 7

improved substantially, which was mainly because shear strength could be used as a mechanical index
to measure the interfacial properties of BF/EP composites. With consistent BF content and VARTM
process, BF/EP-1 composites presented the highest shear strength due to the low viscosity and long
operation time of EP-1 and medium temperature curing of BF/EP-1 composites, and the favorable BF
impregnation of EP-1 during molding to establish a physical and chemical bonding between EP-1
and BFs.
The BF incorporation enhanced the impact toughness of EP-1, EP-2 and EP-3 (Figure 3c). BFs were
prepared through the solution suspension method to produce BF preforms. BFs were interspersed
with each other, and the mechanical interlocking between BFs was enhanced, which could absorb
more impact energy, thus effectively enhancing the impact toughness of the three neat EP samples.
The impact toughness of BF/EP-1, BF/EP-2 and BF/EP-3 composites increased by 56%, 25% and 49%,
respectively, compared to that of EP-1, EP-2 and EP-3. The impact toughness of BF/EP composites
was determined by the impact properties of EP systems and interfacial bonding of BFs and EP. BF-
impregnated EP-1 showed the optimum effect, and thus BF/EP-1 composites represented the largest
improvement in energy absorption during the impact among all BF/EP composites.

Thermal properties
Thermal stability
The thermal stability of three EP systems was higher than that of BFs (Figure 4a,b). The main pyrolysis
stage of BFs was 220–410°C, in which numerous hemicellulose and cellulose in BFs were pyrolyzed

Figure 4. TG-DTG curves of (a), (b) BFs, EP systems and (c), (d) BF/EP composites.
8 J. SHI ET AL.

Figure 5. (A) Storage modulus and (b) loss modulus curves of EP systems and BF/EP composites with different EP systems.

(Yang et al. 2019). Among the three EP systems, EP-1 showed the optimum thermal stability and was
mainly pyrolyzed at 300–500°C. EP-2 presented a large mass loss at 25–300°C and a main pyrolysis
stage at 300–500°C. EP-3 pyrolyzed slightly at 25–250°C and significantly at 250–500°C. As shown in
Figure 4c, when the thermal stability of BFs was lower than that of the three EP systems, the latter
decreased after BF incorporation. BF/EP composites presented a main pyrolysis stage at 250–500°C,
among which BF/EP-1 composites showed a relatively high thermal stability. BF/EP-2 composites
showed a significant mass loss at 25–250°C, about 13.76% at 250°C. In particular, the maximum
pyrolysis rates of three BF/EP composites were lower than that of the corresponding neat EP,
indicating that numerous BF incorporation could effectively reduce the pyrolysis rates of BF/EP
composites in the main pyrolysis stages (Figure 4b,d).

Dynamic mechanical properties


The maximum values of storage modulus for the neat EP samples followed the order of EP-1 > EP-3 >
EP-2 (Figure 5a). BF incorporation into EP could enhance the rigidity of EP. However, the maximum
values of storage modulus for BF/EP composites showed a trend of BF/EP-1 > BF/EP-2 > BF/EP-3,
which was mainly because the dynamic mechanical properties of BF/EP composites were associated
with the properties of BFs, EP systems and the interfacial bonding between BFs and EP. Although the
storage modulus of EP-3 was higher than that of EP-2, the storage modulus of BF/EP-3 composites was
lower than that of BF/EP-2 composites due to EP-3 with a higher viscosity, more voids in BF/EP-3
composites and poor interfacial bonding between BFs and EP-3. In this paper, the temperature at the
maximum loss modulus was used as the glass transition temperature (Tg) of BF/EP composites. BF
incorporation into EP-1 decreased Tg of EP-1 (Figure 5b), while BF incorporation into EP-2 and EP-3
increased Tg of EP-2 and EP-3, which was mainly related to Tg of EP systems. Previous TGA indicated
that EP-2 and EP-3 were pyrolyzed at low temperatures.

Void characteristics
BF/EP composites prepared from different EP systems were scanned by Micro-CT, and the 3D models
of BF/EP composites were reconstructed (Figure 6a). The internal slices of BF/EP composites are
shown in Figure 6b, where the white-gray is BFs, the gray is EP, and the black is voids. Voids in BF/EP
composites were mainly bubbles in BF/EP composites and gaps at the interface between BFs and EP.
BF/EP-1 composites were basically free of obvious voids in the slices, while BF/EP-2 and BF/EP-3
composites had more voids, particularly the latter. In the 3D models (Figure 6c), the white was BFs and
EP, and the blue was the voids in BF/EP composites. The percentages of void volume in BF/EP-1, BF/
JOURNAL OF NATURAL FIBERS 9

Figure 6. Micro-CT scan of BF/EP composites with different EP systems: (a) schematic diagram, (b) Micro-CT slices of BF/EP
composites and (c) void distribution of BF/EP composites.

EP-2 and BF/EP-3 composites were 0.14%, 5.18% and 11.82%, respectively. BF/EP-1 composites had
relatively few and evenly distributed voids, mainly a small number of interfacial bonding gaps
generated at the interface bonding of BFs and EP. More bubbles occurred in BF/EP-2 composites.
Although EP-1 and EP-2 had similar viscosity, the shorter operation time of EP-2 was not conducive
to the adequate BF impregnation by EP-2, and BFs inside BF preforms were more difficult to be
impregnated by EP-2 for a short time due to the high BF content of BF/EP-2 composites. Bubbles and
gaps were more distributed in the interior of BF/EP-2 composites than on the surface layer. They were
most abundant in BF/EP-3 composites in both interior and surface layers. High viscosity and short
operation time significantly increased voids in BF/EP composites.
As shown in Figure 7, BF/EP composites comprised BFs, EP and BF/EP interface. EP systems
exerted a significant impact on the preparation of BF/EP composites by VARTM. The properties of EP
systems laid the basis and had a direct effect on those of BF/EP composites. Moreover, When the
VARTM process was identical, the viscosity of EP systems determined the impregnation effect of EP
10 J. SHI ET AL.

Figure 7. Schema of the mechanism for the preparation of BF/EP composites with different EP systems via the VARTM process.

on BFs. The lower the viscosity of EP systems, the less difficult the EP impregnation into BFs. EP could
fully impregnate BFs and penetrate into the internal BF bundles. Bubbles in the BF/EP composites
became less, and the interfacial gaps between BFs and EP got smaller. The strengthening of physical
and chemical bonding of the BF/EP interface was beneficial for EP reinforcement by BFs. EP systems
with the same viscosity and shorter operation time would significantly lessen EP filling time, and EP
failed to impregnate BFs adequately and effectively. Moreover, more bubbles would appear in BF/EP
composites. The medium temperature curing of BF/EP composites often needs a shorter time than
room temperature curing, conducive to promoting VARTM preparation of BF/EP composites. Before
the epoxy system gelation in BF/EP composites cured at medium temperature, the temperature would
promote the movement of epoxy molecular chains to reduce the viscosity of EP systems and enhance
the impregnation effect of EP on BFs (Francucci et al. 2012). Moreover, in the preparation of BF/EP
composites with high BF content, pressurization of BF preforms is of significance, and the combined
effect of temperature and pressure was more favorable to the formation of hydrogen bonds and van
der Waals forces between functional groups in the EP systems and free hydroxyl groups on the BF
surface (Liu et al. 2012), which enhanced the BF/EP interfacial bonding and thus improved the
properties of BF/EP composites. Taken together, VARTM preparation of BF/EP composites required
EP systems with low viscosity, long operation time and medium temperature curing.

Conclusions
Vacuum-assisted resin transfer molding (VARTM) process could be a suitable method to maximize
the advantages of bamboo fiber reinforced polymer composites and improve their suitability for
JOURNAL OF NATURAL FIBERS 11

industrial manufacturingThe preparation of high-performance BF/EP composites using VARTM


required EP systems with low viscosity, long operation time and medium temperature curing.
Epoxy/anhydride system (EP-1, 265 MPa·s at 23°C, no curing at room temperature and medium
temperature curing) as the matrix for the BF/EP composite preparation by VARTM had relatively high
effectiveness. The flexural strength, flexural modulus, shear strength and impact toughness of BF/EP
composites prepared by EP-1 increased by 9%, 98%, 73% and 56%, respectively, compared with those
of the neat EP-1 samples. The pyrolysis rate of BF/EP composites could be effectively reduced during
the main pyrolysis stage and the BF/EP composites presented more excellent dynamic thermo-
mechanical properties when prepared with EP-1. When the BF content was high, EP-1 was beneficial
to EP impregnation to BFs, reducing the voids in BF/EP composites and improving the BF/EP
interfacial bonding. The percentage of void volume was only 0.14% in 3D reconstruction models of
BF/EP-1 composites.

Acknowledgements
This research was supported by the Cooperation Project of Zhejiang Province Project of Scientific Research Institutes
(Grant No.2021F1065-3).

Disclosure statement
The authors declare that they have no known competing financial interests or personal relationships that could have
appeared to influence the work reported in this paper.

Funding
The work was supported by the the Cooperation Project of Zhejiang Province Project of Scientific Research Institutes
[No.2021F1065-3].

References
ASTM D2344/D2344M-16. 2016. Standard test method for short beam strength of polymer matrix composite materials
and their laminates. West Conshohocken, PA: ASTM International.
Chang, C., and F. Chang. 2022. Physical-mechanical properties of bamboo-textile-reinforced polymer made with
vacuum-assist resin transfer molding. Journal of Natural Fibers, 19 (17):15824–35. Advance online publication.
doi:10.1080/15440478.2022.2133056.
Chen, H., J. Wu, J. Shi, W. Zhang, and H. Wang. 2021. Effect of alkali treatment on microstructure and thermal stability
of parenchyma cell compared with bamboo fiber. Industrial Crops and Products 164:113380. doi:10.1016/j.indcrop.
2021.113380.
Chiu, H. -H., and W. -B. Young. 2020. Characteristic study of bamboo fibers in preforming. Journal of Composite
Materials 54 (25):3871–82. doi:10.1177/0021998320923144.
Chokshi, S., V. Parmar, P. Gohil, and V. Chaudhary. 2020. Chemical composition and mechanical properties of natural
fibers. Journal of Natural Fibers 19 (10):3942–53. doi:10.1080/15440478.2020.1848738.
Depuydt, D. E. C., J. Soete, Y. D. Asfaw, M. Wevers, J. Ivens, and A. W. van Vuure. 2019. Sorption behaviour of bamboo
fibre reinforced composites, why do they retain their properties? Composites Part A, Applied Science and
Manufacturing 119:48–60. doi:10.1016/j.compositesa.2019.01.020.
Francucci, G., A. Vázquez, E. Ruiz, and E. S. Rodríguez. 2012. Capillary effects in vacuum-assisted resin transfer molding
with natural fibers. Polymer Composites 33 (9):1593–602. doi:10.1002/pc.22290.
Furtos, G., L. Molnar, L. Silaghi-Dumitrescu, P. Pascuta, and K. Korniejenko. 2022. Mechanical and thermal properties
of wood fiber reinforced geopolymer composites. Journal of Natural Fibers 19 (13):6676–91. doi:10.1080/15440478.
2021.1929655.
Furtos, G., L. Silaghi-Dumitrescu, M. Moldovan, B. Baldea, R. Trusca, and C. Prejmerean. 2012. Influence of filler/
reinforcing agent and post-curing on the flexural properties of woven and unidirectional glass fiber reinforced
composites. Journal of Materials Science 47:3305–14. doi:10.1007/s10853-011-6169-1.
12 J. SHI ET AL.

Furtos, G., L. Silaghi-Dumitrescu, P. Pascuta, C. Sarosi, and K. Korniejenko. 2021. Mechanical properties of wood fiber
reinforced geopolymer composites with sand addition. Journal of Natural Fibers 18 (2):285–96. doi:10.1080/
15440478.2019.1621792.
GB/T 1449. 2005. Fiber-reinforced plastic composites—determination of flexural poperties. Beijing, China:
Standardization Administration of China.
Gb/T 1451. 2005. Fiber-reinforced plastic composites—Determination of charpy impact properties. Beijing, China:
Standardization Administration of China.
GB/T 1462. 2005. Test methods for water absorption of fiber reinforced plastics. Beijing, China: Standardization
Administration of China.
GB/T 1463. 2005. Test methods for density and relative of fiber reinforced plastics. Beijing, China: Standardization
Administration of China.
GB/T 22314. 2008. Plastics—epoxide resins—determination of viscosity. Beijing, China: Standardization Administration
of China.
Hu, M., C. Wang, C. Lu, N. I. S. Anuar, S. H. S. Yousfani, M. Jing, Z. Chen, S. Zakaria, and H. Zuo. 2019. Investigation on
the classified extraction of the bamboo fiber and its properties. Journal of Natural Fibers 17 (12):1798–808. doi:10.
1080/15440478.2019.1599311.
Jena, H., A. K. Pradhan, and M. K. Pandit. 2014. Studies on water absorption behaviour of bamboo–epoxy composite
filled with cenosphere. Journal of Reinforced Plastics and Composites 33 (11):1059–68. doi:10.1177/
0731684414523325.
Kedari, V. R., B. I. Farah, and K. -T. Hsiao. 2011. Effects of vacuum pressure, inlet pressure, and mold temperature on the
void content, volume fraction of polyester/e-glass fiber composites manufactured with VARTM process. Journal of
Composite Materials 45 (26):2727–42. doi:10.1177/0021998311415442.
Kim, H., K. Okubo, T. Fujii, and K. Takemura. 2013. Influence of fiber extraction and surface modification on
mechanical properties of green composites with bamboo fiber. Journal of Adhesion Science and Technology
27 (12):1348–58. doi:10.1080/01694243.2012.697363.
Liu, D., J. Song, D. P. Anderson, P. R. Chang, and Y. Hua. 2012. Bamboo fiber and its reinforced composites: Structure
and properties. Cellulose 19 (5):1449–80. doi:10.1007/s10570-012-9741-1.
Li, Y., L. Xie, and H. Ma. 2015. Permeability and mechanical properties of plant fiber reinforced hybrid composites.
Materials & Design 86:313–20. doi:10.1016/j.matdes.2015.06.164.
Muhammad, A., M. R. Rahman, S. Hamdan, and K. Sanaullah. 2018. Recent developments in bamboo fiber-based
composites: A review. Polymer Bulletin 76 (5):2655–82. doi:10.1007/s00289-018-2493-9.
Rocky, B. P., and A. J. Thompson. 2018. Production of natural bamboo fibers-1: Experimental approaches to different
processes and analyses. The Journal of the Textile Institute 109 (10):1381–91. doi:10.1080/00405000.2018.1482639.
Sul, I. H., J. R. Youn, and Y. S. Song. 2012. Bubble development in a polymeric resin under vacuum. Polymer Engineering
& Science 52 (8):1733–39. doi:10.1002/pen.23112.
Wang, D., T. Bai, W. Cheng, C. Xu, G. Wang, H. Cheng, and G. Han. 2019. Surface modification of bamboo fibers to
enhance the interfacial adhesion of epoxy resin-based composites prepared by resin transfer molding. Polymers
(Basel) 11 (12):2107. doi:10.3390/polym11122107.
Wang, Q., Y. Zhang, W. Liang, J. Wang, and Y. Chen. 2020. Effect of silane treatment on mechanical properties and
thermal behavior of bamboo fibers reinforced polypropylene composites. Journal of Engineered Fibers and Fabrics
15:1–10. doi:10.1177/1558925020958195.
Xia, C., S. Q. Shi, L. Cai, and J. Hua. 2015. Property enhancement of kenaf fiber composites by means of vacuum-assisted
resin transfer molding (VARTM). Holzforschung 69 (3):307–12. doi:10.1515/hf-2014-0054.
Yang, M., F. Wang, S. Zhou, Z. Lu, S. Ran, L. Li, and J. Shao. 2019. Thermal and mechanical performance of
unidirectional composites from bamboo fibers with varying volume fractions. Polymer Composites
40 (10):3929–37. doi:10.1002/pc.25253.
Yang, J., J. Xiao, J. Zeng, C. Peng, X. Feng, and B. Hou. 2011. An empirical model for resin viscosity during cure in
vacuum infusion molding process. Applied Composite Materials 19 (3–4):573–82. doi:10.1007/s10443-011-9233-8.

You might also like