Next Article in Journal
Quantitation of Flavor Compounds in Refill Solutions for Electronic Cigarettes Using HS-GCxIMS and Internal Standards
Next Article in Special Issue
Can Biomass Mastication Assist the Downstreaming of Polyhydroxyalkanoates Produced from Mixed Microbial Cultures?
Previous Article in Journal
Structural Characterization of Polysaccharide Derived from Gastrodia elata and Its Immunostimulatory Effect on RAW264.7 Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fluorine-Containing Flow Modifier for BN/PPS Composites Enabled by Low Surface Energy

1
School of Materials Science and Engineering, South China University of Technology, Guangzhou 510640, China
2
Key Lab Guangdong High Property & Functional Polymer Materials, and Key Laboratory of Polymer Processing Engineering, Ministry of Education, Guangzhou 510640, China
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(22), 8066; https://doi.org/10.3390/molecules27228066
Submission received: 31 October 2022 / Revised: 14 November 2022 / Accepted: 17 November 2022 / Published: 20 November 2022

Abstract

:
In this study, a fluorine-containing flow modifier (Si-DF) with low surface energy is successfully synthesized, which is applied to fabricate ideal electronic packaging materials (BN/PPS composites) with high thermal conductivity, excellent dielectric properties, processability, and toughness by conventional melt blending. Si-DPF is located at the interface between the BN fillers and the PPS matrix, which not only improves the dispersion of BN fillers but also strengthens the interaction. With the help of 5 wt% Si-DF, BN/PPS/Si-DF (70/25/5) still exhibits the high thermally conductive coefficient (3.985 W/m·K) and low dielectric constant (3.76 at 100 MHz) although BN fillers are loaded as high as 70 wt%. Moreover, the sample processes a lower stable torque value (2.5 N·m), and the area under the stress–strain curves is also increased. This work provides an efficient way to develop high-performance polymer-based composites with high thermally conductive coefficients and low dielectric constants for electronic packaging applications.

1. Introduction

Polyphenylene sulfide (PPS) has been widely used as electronic packaging material in the field of electronic communications due to its outstanding thermal dimensional stability, superior chemical resistance, and high flame retardance. With continuous miniaturization and increased power in electronic devices, the comprehensive performance of electronic packaging materials makes more stringent requirements [1,2,3,4]. An ideal electronic packaging material should demonstrate high thermal conductivity, excellent dielectric properties, processability, and toughness. Unfortunately, the PPS matrix has a relatively low thermally conductive coefficient (<0.4 W/m·K), which needs to be enhanced by incorporating thermally conductive particles [5,6,7,8]. Some studies have shown that introducing a large number of thermally conductive particles leads to a dramatic increase in the thermal conductivity of the PPS-based composites, such as boron nitride (BN), expanded graphite (EG), aluminum nitride (AlN), and carbon nanotubes (CNT) [9,10,11,12].
However, a high loading of particles easily forms a large number of agglomerates, deteriorating the processability and toughness of the composites [13,14]. Moreover, the serious agglomeration of particles also leads to more interfacial thermal barriers, which is not conducive to the fabrication of composites with high thermally conductive coefficients and low dielectric constants [15]. Thus, a major challenge in the fabrication of ideal electronic packaging materials (PPS-based composites) is to achieve the uniform dispersion of thermally conductive particles in the PPS matrix.
Usually, the PPS-based composites with high thermal conductivity are prepared by a hot-compression technique in which the surface-modified particles are premixed with the PPS matrix, which greatly improves the dispersion of the particles [16,17,18,19]. Yang et al. [16] modified BN fillers using a silanization reaction with γ-aminopropyl triethoxy silane/aminopropyllsobutyl polyhedraloligomeric silsesquioxane (f-BN) and found that f-BN fillers had better dispersion in the PPS matrix. The thermally conductive coefficient (λ) value of the f-BN/PPS composite with 60 wt% f-BN was improved to 1.122 W/m·K, four times higher than that of the PPS matrix (0.286 W/m·K). Ryu et al. [19] reported that the grafting of (3-Aminopropyl) triethoxysilane onto the surface of BN fillers reduced agglomeration. The λ value of 3.09 W/m·K was exhibited by the modified BN containing 60 wt%. However, due to the complexity of this method, it cannot be extended to large-scale industrial production. Therefore, the uniformly dispersed particles in the PPS matrix could be easily realized by melt blending technology, which is of prime importance for the development of ideal electronic packaging materials in industry.
In this work, a fluorine-containing flow modifier (Si-DF) is designed and synthesized (Figure 1), which is introduced into BN/PPS composites (the weight ratio of BN to PPS is 60:40 and 70:30) by a conventional melt processing technique. The -CF3 terminal groups endow Si-DF with low surface energy, which enables Si-DF to locate at a two-phase interface [14]. It is expected that the BN fillers will be dispersed in the PPS matrix, and the BN/PPS composites will present high thermal conductivity, excellent processability, and toughness. The composite will also maintain a low dielectric constant with the help of Si-DF. In addition, the dispersion mechanism is explored.

2. Experimental

2.1. Main Materials

Polyphenylene sulfide (PPS) was purchased from Nanjing Deyuan Science and Technology Co. Ltd. (Nanjing, China) Boron nitride (BN, mean diameter of 20–30 μm, thermally conductive coefficient of 33 W/m·K) was received from Qinhuangdao ENO High-Tech Materials Development Co., Ltd. (Qinhuangdao, China) Amino silicone oil was purchased from Shin-Etsu Chemical Industrial Co. Ltd. (Tokyo, Japan) 9,10-dihydro-9-oxa-10-phosphaphenanthrene 10-oxide (DOPO) and 3-(Trifluoromethyl) benzaldehyde were purchased from Shanghai Aladdin Biochemical Co. Ltd. (Shanghai, China).

2.2. Synthesis of Si-DF

Amino silicone oil (17.2 g, 0.02 mol), 3-(Trifluoromethyl) benzaldehyde (6.96 g, 0.04 mol), and 100 mL of ethanol were added sequentially to a 250 mL round-bottom glass flask equipped with a nitrogen inlet, a condensing device, and a magnetic stirrer. The mixture was stirred at 80 °C for 8 h until the complete conversion of amino silicone oil and 3-(Trifluoromethyl) benzaldehyde, as confirmed by FTIR, and then the intermediate (Si-CF) was obtained. DOPO (8.64 g, 0.04 mol) was added into the system and then further reacted for 12 h at room temperature. After the complete conversion of DOPO was tested via FTIR, the ethanol was removed using a rotary evaporator, and the residual solid was recrystallized from ethyl acetate to yield a yellowish powder named Si-DF (30.76 g, yield: 93.8 %). GPC data: Mn = 1398 g/mol, PDI = 1.22.

2.3. Preparation of BN/PPS Composites

BN and PPS were dried in a vacuum oven at 90 °C for 12 h. The BN/PPS composites were prepared by melt blending in a twin-screw extruder, and the temperatures of the extruder were set in the range from 265 °C to 295 °C at a screw speed of 120 rpm. After being granulated and dried, the pellets were injection-molded on an injection-molding machine to obtain different specimens. The molding was conducted at zone temperature profiles of 275-280-290-290-295 °C. The composite preparation process and composition are simply illustrated in Scheme 1, and the compositions and code names of the samples are listed in Table 1.

2.4. Characterization

The FTIR spectra (scanned between 400 and 4000 cm−1) were obtained on a Nicolet 50XC spectrometer (Nicolet, Glendale, WI, USA). The 1H NMR and 31P NMR spectra were recorded on an AV400 unity spectrometer (Bruker, Billerica, MA, USA) operated at 400 MHz with DMSO-d6 as a solvent. The number-averaged molar mass (Mn) and polydispersity (Mw/Mn) of Si-DF were measured by gel permeation chromatography (Waters, Milford, MA, USA). The contact angles were determined on a DSA100 contact angle instrument (KRUSS, Hamburg, Germany) with water and diiodomethane at 25 °C, respectively. Scanning electron microscopy (SEM) images were obtained on a Nova NanoSEM430 scanning electron microscope (FEI, Hillsboro, OR, USA). The torque values of the samples were measured on a Haake torque XSS-300 rheometer (Haake, Vreden, Germany) at a rotation of 50 rpm at 285 °C for 10 min. The thermally conductive coefficient values of the samples were measured on a TPS2200 Hot Disk instrument (AB Corporation, Västerås, Sweden). Dielectric constant values of the samples were measured on a high-frequency Q instrument of QBG-3D (Shanghai Aiyi Electronic Equipment, Shanghai, China). According to ISO 527-2 standard, the tensile strengths of the samples were tested on an AGS-10KNI universal material testing machine (Shimadzu, Kyoto, Japan) at the speed of 10 mm/min.

3. Results and Discussion

3.1. Characterization of Si-DF

During a typical synthetic process, the intermediate Si-CF is first synthesized via a Schiff base reaction between amino silicone oil and 3-(Trifluoromethyl) benzaldehyde in an EtOH solution. The flow modifier Si-DF is then performed through the addition reaction of DOPO and Si-CF. Figure 2 shows the infrared spectra of Si-CF and Si-DF. The synthesis of the Si-CF intermediate is performed by the condensation of amino silicone oil and 3-(Trifluoromethyl) benzaldehyde. Obviously, the strong peak of -CH=N- (Schiff base) appears at 1650 cm−1, and the peak of -CF3 appears at 1650 cm−1, which indicates the triggered aldimine polymerization between amino silicone oil and 3-(Trifluoromethyl) benzaldehyde. The Si-DF is performed through the addition reaction of DOPO and Si-CF. Additionally, the peak of -CH=N- is absent at the same wavenumber for Si-DF, indicating that DOPO has successfully carried out the addition reaction with the intermediate containing the -CH=N- group [14,20,21,22].
The 1H NMR spectrum of Si-DF is displayed in Figure 3a. The peaks at −0.12–0.16 ppm, 0.32–0.46 ppm, and 1.16–1.42 ppm are attributed to the protons in -Si-CH3, Si-CH2-, and -CH2-CH2-, respectively. The peaks at 5.29–5.32 ppm and 5.07–5.11 ppm are attributed to the protons in -NH- and P-CH. The peaks at 6.5–8.2 ppm are attributed to the protons in the benzene ring. The 31P NMR spectra of DOPO and Si-DF are shown in Figure 3b. Based on the 31P NMR spectrum of DOPO, the peaks at 14.40–15.75 ppm are observed, while the peaks are absent at the same wavenumber for Si-DF, and new peaks are observed at 31.50–34.85 ppm. The results from FTIR, 1H NMR, and 31P NMR show that the flow modifier Si-DF has been successfully synthesized.

3.2. Phase Morphology

For highly filled polymer-based composites, how to disperse the fillers in the matrix is a key issue presenting a challenge in the fabrication of high-performance composites. To display the effect of Si-DF on the dispersion of BN fillers, the morphology of the BN fillers in the samples is observed by SEM (Figure 4).
There exists serious agglomeration of BN fillers in the sample BN/PPS (60/40) (Figure 4a1) and BN/PPS (70/30) (Figure 4b1). With the addition of Si-DF, the morphology of the BN fillers presents a striking contrast. It is obviously found that BN fillers exhibit a homogeneous dispersion in the samples BN/PPS/Si-DF (60/35/5) (Figure 4a2) and BN/PPS/Si-DF (70/25/5) (Figure 4b2).
In our previous work, we demonstrated that the fluorine-containing flow modifiers (PMFs) with low surface energy were located at the MH/LLDPE interface in the highly filled MH/LLDPE composites (80:20 by weight), which improved the dispersion of MH particles [23]. Meanwhile, the distribution of flow modifiers can be estimated by evaluating the wetting coefficient (ωa) [24]. Herein, Young’s model is used to calculate the ωa of Si-DF. Additionally, when −1 < ωa < 1, Si-DF is preferred to locate at the interface. The surface energy of the individual components (PPS, BN, and Si-DF) is examined by measuring the contact angles, and the interfacial energies are calculated using the surface energy of the individual components using the geometric-mean equation or harmonic-mean equation [25,26,27]. All of the data are listed in Table 2 and Table 3, respectively.
As seen, the values of ωa for Si-DF are −0.37 and −0.31, respectively, as calculated from the geometric-mean equation, the harmonic-mean equation, and Young’s model, indicating that Si-DF tends to locate at the interface between the BN fillers and PPS matrix. Combining with the phase morphology and theoretical calculations, it is concluded that the distribution of Si-DF at the two-phase interface is effectively able to reduce the formation of agglomeration in the matrix.

3.3. Thermal Conductivity

Figure 5 presents the thermally conductive coefficient (λ) values of the PPS matrix and BN/PPS composites. The λ value of the PPS matrix is 0.372 W/m·K. The composites reveal a dramatic enhancement in thermal conductivity in comparison with the PPS matrix. The λ values of BN/PPS (60/40) and BN/PPS (70/30) are 3.379 W/m·K and 3.569 W/m·K, which are approximately 9 and 9.6 times higher than that of the PPS matrix, respectively. With the addition of Si-DF, the samples present higher thermal conductivity, and the corresponding λ values are increased to 3.873 W/m·K of BN/PPS/Si-DF (60/35/5) and 3.985 W/m·K of BN/PPS/Si-DF (70/25/5), respectively. This implies that efficient thermal transfer pathways could be formed at a high loading of BN fillers in the PPS matrix by the introduction of Si-DF [16,28]. Additionally, it is believed that the homogeneous dispersion of BN fillers in the PPS matrix is the key to the formation of thermal transfer pathways. Furthermore, Si-DF forms a protective layer on the BN surfaces, which shows relatively lower interfacial thermal barriers with the PPS matrix compared to the samples without Si-DF, resulting in higher λ values.
Table 4 summarizes the reported thermally conductive coefficient and thermal conductivity enhancement for the BN/PPS composites. It is noted that the BN/PPS composites with higher BN filler contents in this work show a high thermally conductive coefficient and the highest thermal conductivity enhancement. This work provides a relatively more efficient and facile method to improve the thermal conductivity of BN/PPS composites.

3.4. Dielectric Property

As a key parameter of electronic packaging materials, a low dielectric constant is conducive to reducing the signal propagation time in electronic components, which is very important in practical applications. The dielectric constant-frequency curves of the PPS matrix and BN/PPS composites are displayed in Figure 6. The PPS matrix possesses a dielectric constant value of 3.32 at 100 MHz. It can be seen that the dielectric constant value of the samples is increased as the loading of BN fillers, and the corresponding values are increased to 3.75 for BN/PPS (60/40) and 3.98 for BN/PPS (70/30). It is worth noting that the introduction of Si-DF is beneficial in decreasing the dialectic constant of the samples. For samples BN/PPS/Si-DF (60/35/5) and BN/PPS/Si-DF (70/25/5), the dielectric constant values decreased to 3.55 and 3.76, respectively. The lower dielectric constant should be attributed to the weakening of interface polarization [29,30]. Si-DF not only improves the dispersion of BN fillers but also strengthens the interaction between the BN fillers and PPS matrix, weakening the polarization between the BN fillers and PPS matrix, resulting in relatively lower dielectric constant values of the BN/PPS composites.

3.5. Processability

The processability of BN/PPS composites is evaluated by the torque rheology test [31]. Figure 7 shows the torque vs. time curves for the BN/PPS composites. As expected, the stable torque values of the samples are increased with the loading of BN fillers, and a pronounced decrease in the torque values of the BN/PPS composites with the addition of Si-DF is observed. For samples BN/PPS (60/40) and BN/PPS (70/30), the stable torque values are 3.0 N·m and 11.7 N·m. When 5 wt% of Si-DF is added, the corresponding values are decreased to 2.1 N·m of BN/PPS/Si-DF (60/35/5) and 2.5 N·m of BN/PPS/Si-DF (70/25/5), respectively, indicating that Si-DF has an obvious advantage in improving the processability of the BN/PPS composites. Si-DF tends to be located at the interface and effectively reduces the melt viscosity, which prevents BN fillers from aggregation and decreases the friction between the BN fillers and between the BN fillers and the PPS matrix.

3.6. Mechanical Properties

Mechanical properties are critical for composites, especially tensile strength and toughness. In this work, tensile testing of the PPS matrix and BN/PPS composites is conducted, and the results are shown in Figure 8. The tensile strength of the PPS matrix is 38 MPa. With the addition of BN fillers, the tensile strength of the samples increased, and the corresponding strength increased to 48 MPa for BN/PPS (60/40) and 43 MPa for BN/PPS (70/30). The tensile strength is slightly decreased for the samples with Si-DF. The toughness of the samples is evaluated by calculating the fracture work derived from the area under the stress–strain curve [32,33]. The work of the fracture of the PPS matrix is 13.66 MJ/m3, and the work of the fracture is decreased with the loading of BN fillers. For samples BN/PPS (60/40) and BN/PPS (70/30), the work of the fracture decreased to 7.46 MJ/m3 and 5.53 MJ/m3, respectively. This is understandable because BN fillers aggregate easily and form agglomerations in the PPS matrix. As expected, after adding 5 wt% of Si-DF, the work of the fracture of the samples increased to 11.05 MJ/m3 of BN/PPS/Si-DF (60/35/5) and 8.97 MJ/m3 of BN/PPS/Si-DF (70/25/5), indicating that Si-DPF has an obvious advantage in improving the toughness of the highly filled BN/PPS composites. This result is mainly ascribed to the distribution of Si-DF; that is, Si-DPF tends to be located at the interface and forms a “barrier layer” for BN fillers. Si-DF not only reduces the agglomeration but also transfers the external force received by the samples to the small aggregate of BN filler, which is able to act as the stress concentration points to dissipate the external energy, and the toughness of the sample is achieved.

4. Conclusions

A fluorine-containing flow modifier Si-DF was successfully synthesized and applied to the fabrication of ideal electronic packaging materials (BN/PPS composites) by conventional melt blending techniques. Si-DF with low surface energy was dominantly located at the BN/PPS interface, leading to the relatively homogeneous dispersion of the BN fillers, thus successfully achieving simultaneously high thermal conductivity, excellent dielectric properties, processability, and toughness in BN/PPS composites, solving the well-known problem that polymer-based composites with high loading thermally conductive fillers usually present high thermal conductivity but poor processability and toughness due to the unavoidable aggregation of the fillers. The Si-DF located at the interface effectively constructed thermal conductive paths and reduced the interfacial heat resistance between the BN fillers and the PPS matrix as well. Meanwhile, the carefully developed additive weakened the polarization between the fillers and the matrix, which reduced the dielectric constant of the composites. In addition, the BN/PPS exhibited excellent processability and toughness. Most importantly, this work paved the way for constructing ideal electronic packaging materials through the rational design of flow modifiers.

Author Contributions

Conceptualization, B.C.; methodology, B.C. and X.H.; software, W.Z.; validation, X.H. and P.W.; investigation, B.C. and X.H.; data curation, X.H.; writing—original draft preparation, B.C.; writing—review and editing, B.C.; visualization, P.W.; supervision, W.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Guangdong Basic and Applied Basic Research Foundation (No. 2021A1515110134) and the Opening Project of Key Laboratory of Polymer Processing Engineering (South China University of Technology), Ministry of Education (No. KFKT2102).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the authors upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Shahil, K.M.; Balandin, A.A. Graphene-Multilayer Graphene Nanocomposites as Highly Efficient Thermal Interface Materials. Nano Lett. 2012, 12, 861–867. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Han, Z.; Fina, A. Thermal Conductivity of Carbon Nanotubes and Their Polymer Nanocomposites: A Review. Prog. Polym. Sci. 2011, 36, 914–944. [Google Scholar] [CrossRef] [Green Version]
  3. Lee, G.W.; Park, M.; Kim, J.; Lee, J.I.; Yoon, H.G. Enhanced Thermal Conductivity of Polymer Composites Filled with Hybrid Filler. Compos. Part A Appl. Sci. Manuf. 2006, 37, 727–734. [Google Scholar] [CrossRef]
  4. Song, S.H.; Park, K.H.; Kim, B.H.; Choi, Y.W.; Jun, G.H.; Lee, D.J.; Kong, B.S.; Paik, K.W.; Jeon, S. Enhanced Thermal Conductivity of Epoxy-Graphene Composites by Using Non-Oxidized Graphene Flakes with Non-Covalent Functionalization. Adv. Mater. 2013, 25, 732–737. [Google Scholar] [CrossRef] [PubMed]
  5. Xu, L.; Chen, G.; Wang, W.; Li, L.; Fang, X. A Facile Assembly of Polyimide/Graphene Core–Shell Structured Nanocomposites with Both High Electrical and Thermal Conductivities. Compos. Part A Appl. Sci. Manuf. 2016, 84, 472–481. [Google Scholar] [CrossRef]
  6. Kim, H.S.; Jang, J.U.; Yu, J.; Kim, S.Y. Thermal Conductivity of Polymer Composites Based on the Length of Multi-Walled Carbon Nanotubes. Compos. Part B Eng. 2015, 79, 505–512. [Google Scholar] [CrossRef]
  7. Gu, J.; Yang, X.; Lv, Z.; Li, N.; Liang, C.; Zhang, Q. Functionalized Graphite Nanoplatelets/Epoxy Resin Nanocomposites with High Thermal Conductivity. Int. J. Heat Mass Transf. 2016, 92, 15–22. [Google Scholar] [CrossRef]
  8. Huang, X.; Zhi, C.; Jiang, P.; Golberg, D.; Bando, Y.; Tanaka, T. Polyhedral Oligosilsesquioxane-Modified Boron Nitride Nanotube Based Epoxy Nanocomposites: An Ideal Dielectric Material with High Thermal Conductivity. Adv. Funct. Mater. 2013, 23, 1824–1831. [Google Scholar] [CrossRef]
  9. Han, Y.; Shi, X.; Yang, X.; Guo, Y.; Zhang, J.; Kong, J.; Gu, J. Enhanced Thermal Conductivities of Epoxy Nanocomposites Via Incorporating in-Situ Fabricated Hetero-Structured SiC-BNNS Fillers. Compos. Sci. Technol. 2020, 187, 10944. [Google Scholar] [CrossRef]
  10. Guo, Y.; Xu, G.; Yang, X.; Ruan, K.; Ma, T.; Zhang, Q.; Gu, J.; Wu, Y.; Liu, H.; Guo, Z. Significantly Enhanced and Precisely Modeled Thermal Conductivity in Polyimide Nanocomposites with Chemically Modified Graphene Via in Situ Polymerization and Electrospinning-Hot Press Technology. J. Mater. Chem. C 2018, 6, 3004–3015. [Google Scholar] [CrossRef]
  11. Xie, S.H.; Zhu, B.K.; Li, J.B.; Wei, X.Z.; Xu, Z.K. Preparation and Properties of Polyimide/Aluminum Nitride Composites. Polym. Test. 2004, 23, 797–801. [Google Scholar] [CrossRef]
  12. Jiang, Q.; Wang, X.; Zhu, Y.; Hui, D.; Qiu, Y. Mechanical, Electrical and Thermal Properties of Aligned Carbon Nanotube/Polyimide Composites. Compos. Part B Eng. 2014, 56, 408–412. [Google Scholar] [CrossRef]
  13. Guan, X.; Cao, B.; Cai, J.; Ye, Z.; Lu, X.; Huang, H.; Liu, S.; Zhao, J. Design and Synthesis of Polysiloxane Based Side Chain Liquid Crystal Polymer for Improving the Processability and Toughness of Magnesium Hydrate/Linear Low-Density Polyethylene Composites. Polymers 2020, 12, 911. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Cao, B.; Wang, O.; Cai, J.; Xie, W.; Liu, S.; Zhao, J. Silicone/Fluorine-Functionalized Flow Modifier with Low Surface Energy for Improving Interfaces in Highly Filled Composites. Compos. Sci. Technol. 2021, 214, 108984. [Google Scholar] [CrossRef]
  15. Gu, J.; Guo, Y.; Yang, X.; Liang, C.; Geng, W.; Tang, L.; Li, N.; Zhang, Q. Synergistic Improvement of Thermal Conductivities of Polyphenylene Sulfide Composites Filled with Boron Nitride Hybrid Fillers. Compos. Part A Appl. Sci. Manuf. 2017, 95, 267–273. [Google Scholar] [CrossRef]
  16. Yang, X.; Tang, L.; Guo, Y.; Liang, C.; Zhang, Q.; Kou, K.; Gu, J. Improvement of Thermal Conductivities for Pps Dielectric Nanocomposites Via Incorporating NH2-POSS Functionalized nBN Fillers. Compos. Part A Appl. Sci. Manuf. 2017, 101, 237–242. [Google Scholar] [CrossRef]
  17. Pan, C.; Kou, K.; Jia, Q.; Zhang, Y.; Wu, G.; Ji, T. Improved Thermal Conductivity and Dielectric Properties of hBN/PTFE Composites Via Surface Treatment by Silane Coupling Agent. Compos. Part B Eng. 2017, 111, 83–90. [Google Scholar] [CrossRef]
  18. Jiang, Y.; Liu, Y.; Min, P.; Sui, G. BN-MWCNT/PPS core-shell Structure Particles and Their 3D Segregated Architecture Composites with High Thermal Conductivities. Compos. Sci. Technol. 2017, 144, 63–69. [Google Scholar] [CrossRef]
  19. Ryu, S.; Kim, K.; Kim, J. Silane Surface Modification of Boron Nitride for High Thermal Conductivity with Polyphenylene Sulfide Via Melt Mixing Method. Polym. Adv. Technol. 2017, 28, 1489–1494. [Google Scholar] [CrossRef]
  20. Chao, P.; Li, Y.; Gu, X.; Han, D.; Jia, X.; Wang, M.; Zhou, T.; Wang, T. Novel Phosphorus–Nitrogen–Silicon Flame Retardants and Their Application in Cycloaliphatic Epoxy Systems. Polym. Chem. 2015, 6, 2977–2985. [Google Scholar] [CrossRef]
  21. Ding, C.; Hu, D.; He, X.; Lai, Y.; Shao, G. Fabrication and microstructure evolution of monolithic bridged polysilsesquioxane-derived SiC ceramic aerogels. Ceram. Int. 2022, 48, 25833–25839. [Google Scholar] [CrossRef]
  22. Shao, G.; Shen, X.; Huang, X. Multilevel Structural Design and Heterointerface Engineering of a Host–Guest Binary Aerogel toward Multifunctional Broadband Microwave Absorption. ACS Mater. Lett. 2022, 4, 1787–1797. [Google Scholar] [CrossRef]
  23. Cao, B.; Zhou, Y.; Wu, Y.; Cai, J.; Guan, X.; Liu, S.; Zhao, J.; Zhang, M. Simultaneous Improvement of Processability and Toughness of Highly Filled MH/LLDPE Composites by Using Fluorine-Containing Flow Modifiers. Compos. Part A Appl. Sci. Manuf. 2020, 134, 105900. [Google Scholar] [CrossRef]
  24. Zhang, W.; Li, X.; Shan, Z.; Wang, S.; Xiao, Y. Surface Modification of Magnesium Hydroxide by Wet Process and Effect on the Thermal Stability of Silicone Rubber. Appl. Surf. Sci. 2019, 465, 740–746. [Google Scholar] [CrossRef]
  25. Jalali Dil, E.; Favis, B.D. Localization of Micro- and Nano-Silica Particles in Heterophase Poly(Lactic Acid)/Poly(Butylene Adipate-Co-Terephthalate) Blends. Polymer 2015, 76, 295–306. [Google Scholar] [CrossRef]
  26. Zha, X.J.; Pu, J.H.; Ma, L.F.; Li, T.; Bao, R.Y.; Bai, L.; Liu, Z.Y.; Yang, M.B.; Yang, W. A Particular Interfacial Strategy in PVDF/OBC/MWCNT Nanocomposites for High Dielectric Performance and Electromagnetic Interference Shielding. Compos. Part A Appl. Sci. Manuf. 2018, 105, 118–125. [Google Scholar] [CrossRef]
  27. Wang, X.; Peng, S.; Chen, H.; Yu, X.; Zhao, X. Mechanical Properties, Rheological Behaviors, and Phase Morphologies of High-Toughness PLA/PBAT Blends by in-Situ Reactive Compatibilization. Compos. Part B Eng. 2019, 173, 107028. [Google Scholar] [CrossRef]
  28. Yuan, C.; Duan, B.; Li, L.; Xie, B.; Huang, M.; Luo, X. Thermal Conductivity of Polymer-Based Composites with Magnetic Aligned Hexagonal Boron Nitride Platelets. ACS Appl. Mater. Interfaces 2015, 7, 13000–13006. [Google Scholar] [CrossRef]
  29. Guo, Y.; Ruan, K.; Shi, X.; Yang, X.; Gu, J. Factors Affecting Thermal Conductivities of the Polymers and Polymer Composites: A Review. Compos. Sci. Technol. 2020, 193, 108134. [Google Scholar] [CrossRef]
  30. Kim, K.; Kim, M.; Kim, J. Thermal and Mechanical Properties of Epoxy Composites with a Binary Particle Filler System Consisting of Aggregated and Whisker Type Boron Nitride Particles. Compos. Sci. Technol. 2014, 103, 72–77. [Google Scholar] [CrossRef]
  31. Hidalgo, J.; Jiménez-Morales, A.; Torralba, J.M. Torque Rheology of Zircon Feedstocks for Powder Injection Moulding. J. Eur. Ceram. Soc. 2012, 32, 4063–4072. [Google Scholar] [CrossRef] [Green Version]
  32. Zhou, Y.; Chen, C.; Zhu, S.; Sui, C.; Wang, C.; Kuang, Y.; Ray, U.; Liu, D.; Brozena, A.; Leiste, U.H.; et al. A Printed, Recyclable, Ultra-Strong, and Ultra-Tough Graphite Structural Material. Mater. Today 2019, 30, 17–25. [Google Scholar] [CrossRef]
  33. Shah, D.; Maiti, P.; Jiang, D.D.; Batt, C.A.; Giannelis, E.P. Effect of Nanoparticle Mobility on Toughness of Polymer Nanocomposites. Adv. Mater. 2005, 17, 525–528. [Google Scholar] [CrossRef]
Figure 1. Synthesis route of Si-DF.
Figure 1. Synthesis route of Si-DF.
Molecules 27 08066 g001
Scheme 1. Schematic illustration of PPS composites preparation.
Scheme 1. Schematic illustration of PPS composites preparation.
Molecules 27 08066 sch001
Figure 2. FTIR spectra of Si-CF and Si-DF.
Figure 2. FTIR spectra of Si-CF and Si-DF.
Molecules 27 08066 g002
Figure 3. (a) 1H NMR spectrum of Si-DF. (b) 31P NMR spectra of Si-DF and DOPO.
Figure 3. (a) 1H NMR spectrum of Si-DF. (b) 31P NMR spectra of Si-DF and DOPO.
Molecules 27 08066 g003
Figure 4. SEM images of BN/PPS (60/40) (a1), BN/PPS/Si-DF (60/35/5) (a2), BN/PPS (70/30) (b1) and BN/PPS/Si-DF (70/25/5) (b2).
Figure 4. SEM images of BN/PPS (60/40) (a1), BN/PPS/Si-DF (60/35/5) (a2), BN/PPS (70/30) (b1) and BN/PPS/Si-DF (70/25/5) (b2).
Molecules 27 08066 g004
Figure 5. Thermal conductivity of PPS matrix and BN/PPS composites.
Figure 5. Thermal conductivity of PPS matrix and BN/PPS composites.
Molecules 27 08066 g005
Figure 6. Dielectric constant-frequency curves of PPS matrix and BN/PPS composites.
Figure 6. Dielectric constant-frequency curves of PPS matrix and BN/PPS composites.
Molecules 27 08066 g006
Figure 7. Torque versus time of BN/PPS composites.
Figure 7. Torque versus time of BN/PPS composites.
Molecules 27 08066 g007
Figure 8. Strain–stress curves of PPS matrix and BN/PPS composites.
Figure 8. Strain–stress curves of PPS matrix and BN/PPS composites.
Molecules 27 08066 g008
Table 1. Compositions and code names of the samples.
Table 1. Compositions and code names of the samples.
SamplesBN (g)PPS (g)Si-DF (g)
BN/PPS (60/40)6040-
BN/PPS/Si-DF (60/35/5)60355
BN/PPS (70/30)7030-
BN/PPS/Si-DF (70/25/5)70255
Table 2. Contact angles and surface energies of PPS, BN and Si-DF.
Table 2. Contact angles and surface energies of PPS, BN and Si-DF.
SamplesContact Angle (°)γ
(mJ/m2)
γd
mJ/m2)
γp
(mJ/m2)
WaterDiiodomethane
PPS90.3724.7646.9946.630.36
BN89.1264.8926.822.54.3
Si-DF98.4551.2034.0833.780.3
Table 3. Preferential distribution of Si-DF in the BN/PPS composites according to interfacial energies γ and wetting coefficients ωa.
Table 3. Preferential distribution of Si-DF in the BN/PPS composites according to interfacial energies γ and wetting coefficients ωa.
Interfacial EnergyBased on Geometric-Mean Equation (mJ/m2)Based on Harmonic-Mean Equation (mJ/m2)ωGωHFlow Modifiers Distribution
γBN/PPS6.5211.75−0.37−0.31BN/PPS interface
γSi-DF/PPS1.042.06
γSi-DF/BN3.475.74
Table 4. Comparison of thermally conductive coefficient and thermal conductivity enhancement in BN/PPS composites.
Table 4. Comparison of thermally conductive coefficient and thermal conductivity enhancement in BN/PPS composites.
SamplesFiller Loading [wt%]Thermally Conductive Coefficient
[W/m·K]
Thermal Conductivity Enhancement
[%]
StrategyYear [Ref]
BN/PPS602.638822.3Hot-compression2017 [15]
BN/PPS601.122292.3Surface-modification, Hot-compression2017 [16]
BN/PPS40 (vol)2.45880Hot-compression2017 [18]
BN/PPS603.1785.8Surface-modification2017 [19]
BN/PPS602.7610.5Melt blending2017 [19]
BN/PPS603.873941.1Melt blendingThis work
BN/PPS703.985971.2Melt blendingThis work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cao, B.; Huang, X.; Zhang, W.; Wu, P. Fluorine-Containing Flow Modifier for BN/PPS Composites Enabled by Low Surface Energy. Molecules 2022, 27, 8066. https://doi.org/10.3390/molecules27228066

AMA Style

Cao B, Huang X, Zhang W, Wu P. Fluorine-Containing Flow Modifier for BN/PPS Composites Enabled by Low Surface Energy. Molecules. 2022; 27(22):8066. https://doi.org/10.3390/molecules27228066

Chicago/Turabian Style

Cao, Bo, Xiaodan Huang, Wenxiang Zhang, and Peng Wu. 2022. "Fluorine-Containing Flow Modifier for BN/PPS Composites Enabled by Low Surface Energy" Molecules 27, no. 22: 8066. https://doi.org/10.3390/molecules27228066

Article Metrics

Back to TopTop