Home Single-source-precursor synthesis and characterization of SiAlC(O) ceramics from a hyperbranched polyaluminocarbosilane
Article Open Access

Single-source-precursor synthesis and characterization of SiAlC(O) ceramics from a hyperbranched polyaluminocarbosilane

  • Le Yang , Pei Zhang , Yao Feng and Zhaoju Yu EMAIL logo
Published/Copyright: April 6, 2022

Abstract

Polyaluminocarbosilane (PACS) as a single-source-precursor of SiAlC(O) ceramic was prepared by reacting a hyperbranched allylhydridopolycarbosilane (AHPCS) and aluminum(iii) acetylacetonate (Al(acac)3), and the PACS was characterized using gel-permeation chromatography, Fourier-transform infrared spectroscopy, and nuclear magnetic-resonance spectroscopy. The polymer-to-ceramic transformation of the obtained PACSs was investigated by Fourier-transform infrared spectroscopy and thermogravimetric analysis. The ceramic yield of the PACS was approximately 15% higher than that of the original AHPCS at 1,200°C. The phase composition and microstructure of the final ceramics were studied by X-ray diffraction, energy-dispersive spectroscopy, and scanning electron microscopy. The introduction of aluminum to the SiC(O) ceramics suppressed the β-SiC crystal growth and improved the density of the ceramics that were annealed at 1,800°C, which is advantageous for high-temperature ceramics. The aluminum content of the SiAlC(O) ceramics can be readily controlled by the Al(acac)3 content in the PACS precursors.

1 Introduction

Recently, SiC-based ceramics that contain heterogeneous elements (Al, Zr, Mn, Ti, and B) have received considerable attention because of their combined structural and functional properties [1,2,3,4,5,6,7,8,9]. Intensive investigations into the development of preceramic polymers to SiC-based ceramics that contain heterogeneous elements have been performed [8,9]. The polymer-derived ceramic (PDC) approach is a promising method for preparing SiC-based ceramics because the chemical composition, phase composition, and even the ceramic microstructure can be designed and tailored. A key factor of the PDC method is the synthesis of suited precursors. Among these preceramic precursors, silicon-containing polymers, especially polycarbosilanes (PCSs), have been developed most vigorously [10,11,12].

It is well known that aluminum is a good sintering additive, and incorporating Al into PCS as a sintering additive could achieve high-performance SiC-based ceramics [7,8,13,14,15,16,17,18,19]. Sorarù et al. prepared a Si–Al–O–C fiber by pyrolyzing green polyaluminocarbosilane (PACS) from a reaction between Yajima PCS and either Al(OBus)3 or Al(OBus)2(etac) [13]. Another significantly stable SiC fiber that contains Al was reported by using a similar precursor obtained by reaction between Yajima PCS and aluminum(iii) acetylacetonate [Al(acac)3] at 300°C in a nitrogen atmosphere [14]. A super-high temperature-resistant fiber that was stable at 2,200°C was produced by a procedure analogous to that used for Nicalon, except for the incorporation of Al and a final sintering at 1,800°C. Cao et al. [15,16] and Yu et al. [17] conducted additional in-depth research to study the synthetic procedure and reaction mechanism between PCS and Al(acac)3. To reduce the sublimation of Al(acac)3, PACS was prepared by reaction of a liquid product produced by thermolysis of polysilacarbosilane (PSCS) instead of PCS with Al(acac)3 above 300°C in a nitrogen stream at atmospheric pressure [18,19].

Recently, liquid hyperbranched polycarbosilanes (HBPCSs) have attracted extensive attention as excellent effective precursors, especially as a matrix source because of their unique structures and favorable properties, such as more favorable solubilities and larger number of reactive groups (e.g., Si–H x , C═C, and/or C≡C) [20]. In our previous work, we synthesized a series of HBPCSs [21,22,23,24]. The polymer-to-ceramic conversion of the HBPCSs for SiC ceramics has been studied further [25,26,27]. Our results show that the HBPCSs can be cured at relatively low temperatures (140–170°C) with a high ceramic yield, which is a promising precursor for SiC matrices and coating materials. More recently, we incorporated various metals, such as Ti, Zr, Fe, and Hf into the AHPCS by chemical modification, which provides new access to obtain metal-containing SiC ceramics with a combined structural and functional performance [28,29,30,31].

Based on the findings that an Fe-containing single-source-precursor was synthesized via the reaction between AHPCS and Fe(acac)3, in this study, we prepared an Al-containing single-source-precursor using Al(acac)3 as a source of Al. We prepared a hyperbranched PACS by Al(acac)3 and hyperbranched AHPCS, which involves massive Si–H x groups and C═C groups under argon. For Yajima PCS, a relatively high temperature (310°C) is required to guarantee the chemical reactions between PCS and Al(acac)3 [15]. However, sublimation of Al(acac)3 at such a high temperature is unavoidable. Compared with the Yajima PCS, the AHPCS contains a large amount of Si–H x and C═C groups, which shows higher reactivity, and the AHPCS can react with Al(acac)3 at a much lower temperature (140°C). Herein we report our first results on the reactivity of the Al(acac)3 vs the AHPCS and the application of resultant polyaluminocarbosilanes (PACSs) to the preparation of the SiAlC(O) ceramics.

2 Experimental methods

2.1 Raw materials

All manipulations were carried out using standard high-vacuum or inert-atmosphere techniques described by Shriver [33]. Liquid AHPCS with a general empirical formula [SiH1.26(CH3)0.60(CH2CH═CH)0.14CH2] n was prepared by a one-pot synthesis with CH2═CHCH2Cl, Cl2Si(CH3)CH2Cl, and Cl3SiCH2Cl as the starting materials [21]. AHPCS used in this work has a number-average molecular weight of approximately 600 and a polydispersity index of 4.50. Al(acac)3 (99%) was from Alfa Aesar. Dimethylbenzene was distilled for use. Other commercially available reagents were used as received.

2.2 Synthesis of precursors

A series of PACS precursors with different mass ratios of Al to AHPCS (1, 3, and 5%) was synthesized, and the precursors were denoted PACS-1, PACS-3, and PACS-5, respectively. In a typical synthesis, a 150 mL Schlenk flask was equipped with a magnetic stirrer and a refluxing condenser that was connected to a line adapter for argon and vacuum. Al(acac)3 (3.6 g) and 50 mL of dry dimethylbenzene were introduced into the flask to obtain a clear yellow solution with stirring, and then 10.0 g AHPCS was introduced into the Schlenk flask. Air was evacuated and the flask was filled with argon steam. The flask was heated to 140°C for 24 h. A viscous yellow product was obtained after the solvent had been removed under vacuum. A control test was performed under the same condition without Al(acac)3 addition and the synthetic sample was termed control AHPCS.

2.3 Synthesis of SiAlC(O) ceramics

Cross-linking, pyrolysis, and annealing were used to obtain the SiAlC(O) ceramics. The control AHPCS and PACSs (5 g samples) were cross-linked in a 170°C oil bath for 6 h. The control AHPCS was transformed slowly into a viscous and yellow colloid, whereas the PACSs were transformed rapidly into a stiff yellow solids. To study the structural transformation during the polymer-to-ceramic conversion, cross-linked samples were pyrolyzed as follows: samples were heated to the desired temperature (300, 600, or 900°C) at 5°C·min−1, held for 2 h, and cooled to room temperature. The amorphous ceramics (pre-pyrolyzed at 900°C) were annealed at a predetermined temperature (1,200, 1,400, 1,600, and 1,800°C) at 40°C·min−1 and kept at this temperature for 2 h in argon. After annealing, the resulting ceramic was cooled naturally to room temperature.

2.4 Characterization

The molecular-weight distributions were collected by gel-permeation chromatography (GPC) measurements using an Agilent 1100 system (Agilent, Palo Alto, CA) at 35°C with tetrahydrofuran as the eluant at 1.0 mL·min−1 and narrow polystyrene standards for calibration. Fourier-transform infrared (FT-IR) spectra were recorded from 4,000 to 400 cm−1 on a Nicolet Avatar 360 apparatus (Nicolet, Madison, WI) in transmission mode by making KBr plates for the liquid samples and KBr disks for the solid samples. Nuclear magnetic-resonance (NMR) spectroscopy experiments were carried out on a Bruker AV 300 MHz spectrometer (Bruker, Germany) at 300.13 MHz for hydrogen-1 and 75.46 MHz for carbon-13 (1H-decoupling) with a delay time of 30 s. The solvent used in the NMR was CDCl3. The 1H and 13C chemical shifts were referenced to tetramethylsilane as the external standard. Thermogravimetric analysis (TGA) was carried out in argon using a thermal-analysis device (Netzsch STA 409 EP, Netzsch, Germany) from room temperature to 1,200°C with a ramping rate of 10°C·min−1. X-ray diffraction (XRD) patterns of the ceramic powders were obtained on a PAN-alytical X’Pert PRO diffractometer (PAN-alytical, Netherlands) with CuKα radiation. Specimens were scanned continuously from 10° to 90° (2θ) at 0.01678° s−1. The ceramic composition was determined by energy-dispersive spectroscopy (EDS, JAX-8100, Japan). Elemental carbon was measured using a Horiba Carbon/Sulfur Analyzer EMIA-320V (Horiba, Japan) and a Horiba Oxygen/Nitrogen Analyzer EMGA-620W (Horiba, Japan) was used to measure elemental oxygen. Scanning electron microscopy (SEM) (Model 1530, LEO, Germany) was used to observe the morphologies of the obtained ceramics.

3 Results and discussion

3.1 Polymer characterization

To investigate the variation in molecular weights, GPC curves of the parent AHPCS and the typical PACS-3 are shown in Figure 1. The number-average molecular weight (M n) of the parent AHPCS and the typical PACS-3 was 600 and 1,000, respectively, the weight-average molecular weight (M w) of the parent AHPCS and the typical PACS-3 was 2,700 and 4,600, respectively, and the polydispersity index (PDI) of the parent AHPCS and PACS-3 was 4.5 and 4.9, respectively. The GPC curve of the PACS-3 exhibits some new high-molecular-weight peaks compared with that of the AHPCS, which suggests that a chemical reaction occurred between the AHPCS and Al(acac)3, and the reaction mechanism will be studied in more detail.

Figure 1 
                  GPC curves of parent AHPCS and PACS-3.
Figure 1

GPC curves of parent AHPCS and PACS-3.

The FT-IR spectra of the resultant precursors are shown in Figure 2. The PACSs exhibit typical AHPCS characteristics and the peaks in these spectra are assigned according to the literature [22,25,26]. A comparison of the FT-IR spectra of the PACSs with those of the AHPCS and Al(acac)3 shows that PACSs contain characteristic peaks of AHPCS and Al(acac)3. The characteristic absorption peaks of Al(acac)3 at 1,535 cm−1 assigned to C═C stretching and 1,592 cm−1 assigned to C═O stretching are also present in the spectra of PACSs [13]. Compared with the parent AHPCS and the control AHPCS, the Si–H stretch (2,130 cm−1) peak of the PACSs decreases, which indicates that a reaction occurred between AHPCS and Al(acac)3.

Figure 2 
                  FT-IR spectra of parent AHPCS, control AHPCS, PACSs, and Al(acac)3.
Figure 2

FT-IR spectra of parent AHPCS, control AHPCS, PACSs, and Al(acac)3.

To confirm the architecture of the resultant polymers, 1H and 13C NMR analysis was carried out on the parent AHPCS, control AHPCS, Al(acac)3, and the resultant PACS. The 1H NMR spectra are shown in Figure 3. Compared with the AHPCS, the PACSs show two obvious additional peaks at 1.99 ppm and 5.47 ppm because of the –CH3 and –CH═C═O from Al(acac)3, respectively; however, these two peaks from PACSs show a chemical shift compared with Al(acac)3. With the increase in Al(acac)3 in the feed, the chemical shift moves to a higher nuclear magnetic field. Compared with that of the control AHPCS, the Si–H x peaks of the PACSs decrease, which matches the FT-IR results. A new peak occurred at 2.3 ppm in PACSs, which shows the existence of a new hydrogen environment assigned to hydrogen protons of residual acetylacetone chain segment, possibly because of the condensation reaction of Si–H bonds in AHPCS and the ligands of Al(acac)3 [12,32,34].

Figure 3 
                  
                     1H NMR spectra in CDCl3 of Al(acac)3, parent AHPCS, control AHPCS, and PACSs.
Figure 3

1H NMR spectra in CDCl3 of Al(acac)3, parent AHPCS, control AHPCS, and PACSs.

The 13C NMR spectra of the parent AHPCS, control AHPCS, PACS-5, and Al(acac)3 are shown in Figure 4. The PACS-5 contains characteristic peaks of AHPCS and Al(acac)3. A chemical shift at 191.5 ppm is attributed to C═O, 101.2 ppm is attributed to CH═C, and 26.9 ppm is attributed to –CH3 in Al(acac)3. A new peak at 25 ppm occurred in PACS-5, which is assigned to the carbon protons of a residual acetylacetone chain segment. The 13C NMR results agree well with the FT-IR and 1H NMR results. Therefore, it is believed that the change in chemical shift and the existence of a residual acetylacetone chain segment can confirm the successful introduction of Al(acac)3 into AHPCS chains to form a single-source-precursor.

Figure 4 
                  
                     13C NMR spectra in CDCl3 of parent AHPCS, control AHPCS, PACS-5, and Al(acac)3.
Figure 4

13C NMR spectra in CDCl3 of parent AHPCS, control AHPCS, PACS-5, and Al(acac)3.

Based on the GPC, FTIR, and NMR analyses and a study of the reaction mechanism between AHPCS and Fe(acac)3 reported in our previous work [32], the reaction mechanism between AHPCS and Al(acac)3 is shown in Figure 5. The chemical shift in NMR of –CH3 and –CH═C═O from Al(acac)3 most likely results from the formation of Si–O–Al bonds. However, because of the existence of steric hindrance, it is difficult for all three Al(acac)3 ligands to perform a condensation reaction, which is confirmed by the finding that the characteristic peaks of Al(acac)3 in the FTIR and NMR spectra of the PACS precursors obtained at 140°C result from remaining ligands of Al(acac)3.

Figure 5 
                  Reaction mechanism between AHPCS and Al(acac)3.
Figure 5

Reaction mechanism between AHPCS and Al(acac)3.

3.2 Polymer-to-ceramic transformation

Cross-linking treatment of polymeric precursors can improve the ceramic yield. Referencing our previous work [26,27], the resultant PACSs and control AHPCS were cured at 170°C for 6 h under argon before pyrolysis. The resultant PACSs were converted rapidly into compact and stiff yellow solids within 20–30 min, whereas the control AHPCS was converted slowly into viscous and yellow colloid within 3 h. This finding suggests that the cross-linking of the PACSs is enhanced significantly via the introduction of Al(acac)3, which is confirmed by FT-IR (Figure 6).

Figure 6 
                  FT-IR spectra of control AHPCS, cured control AHPCS, PACSs, and cured PACSs.
Figure 6

FT-IR spectra of control AHPCS, cured control AHPCS, PACSs, and cured PACSs.

The intensities of the Si–H peaks at 2,130 cm−1 of the cured control AHPCS and PACSs decrease compared with that of the control AHPCS and PACSs. The reaction degrees of Si–H (P Si–H) of the cured control AHPCS and PACSs can be calculated according to the following equation on the basis of a semi-quantitative method [29]:

P Si H = ( A Si H / A Si CH 3 ) uncured ( A Si H / A Si CH 3 ) cured ( A Si H / A Si CH 3 ) uncured ,

Because the Si–CH3 bond (1,253 cm−1) should not participate in the chemical reaction below 600°C, the intensity of the Si–CH3 bond was selected to determine the variation in Si–H bond intensities [23,30]. The P Si–H of the cured control AHPCS, PACS-1, PACS-3, and PACS-5 were 7.8, 18.5, 28.9, and 29.2%, respectively. The increased introduction of Al(acac)3 into the PACSs resulted in an increased number of Si–H groups taking part in the reaction and shows that further reaction occurred between the remaining Si–H groups of the PACSs with Al–O groups for further cross-linking, as described in Figure 5. The decrease in characteristic peaks of Al(acac)3 can also confirm the conclusion. The C═C peaks of the cured control AHPCS and cured PACSs decreased after cross-linking, which indicates that hydrosilylation between AHPCS and AHPCS also occurred during cross-linking.

Structural evolution during the polymer-to-ceramic conversion was researched with the PACS-5 by FT-IR (Figure 7). The intensities of active groups, such as Si–H and C═C in PACS-5, decrease gradually with an increase in curing temperature from room temperature to 600°C, because of further cross-linking via the condensation reaction between Si–H and Al–O and hydrosilylation between Si–H and C═C. The C═C peaks and the Si–H peaks almost vanish up to 600°C. The intensities of the Si–CH3 and Si–CH2–Si decrease with an increase in pyrolysis temperature; this is attributed to the decomposition of organic groups. The disappearance of C═O groups at 1,592 cm−1 in the ligands of Al(acac)3 at 600°C is attributed to the ligand decomposition. At 900°C, only one broad peak at ∼780 cm−1 belonged to the retention of an amorphous SiC framework structure, which indicates that the polymer-to-ceramic conversion is completed. From 1,200 to 1,600°C, the SiC peak band sharpens, and a shift in its position occurred from 780 to 830 cm−1 after careful examination, which indicates the formation of crystalline SiC [30]. Si–O–Si absorption occurred because of the introduction of oxygen from Al(acac)3 at 1,200–1,400°C, and the intensity of Si–O–Si absorption decreases with the increase in the temperature. Further heating to 1,600°C resulted in the Si–O–Si peak disappearing with an increase in crystallized SiC absorption.

Figure 7 
                  FT-IR spectra of PACS-5 heat treated at different temperatures in Ar.
Figure 7

FT-IR spectra of PACS-5 heat treated at different temperatures in Ar.

Figure 8 shows the thermal behavior of the cured control AHPCS and PACSs from room temperature to 1,200°C. The 1,200°C ceramic yield of the cured control AHPCS, PACS-1, PACS-3, and PACS-5 reached 49.0, 54.1, 63.6, and 57.4%, respectively, which indicates that the introduction of Al(acac)3 to AHPCS improves the ceramic yield. The onset of thermal decomposition for PACSs is ∼190°C, which is consistent with the melting point of Al(acac)3. At 315°C, which is the boiling point of Al(acac)3, the mass loss of cured-control AHPCS was 15.3%, whereas those of the cured PACS-1, PACS-3, and PACS-5 were 18.3, 14.7, and 23.2%, respectively. The cured PACS-3 has the least mass loss. By combining the study of PSi–H, the PSi–H of PACS-3 and PACS-5 were similar at 28.9 and 29.2%, respectively, which indicates that the cross-linking degree was maximum with an increase in Al(acac)3 in the feed, whereas an increase in Al(acac)3 in the feed would lead to increased Al(acac)3 volatilization. Therefore, a relatively complete reaction and less volatilization of Al(acac)3 in PACS-3 yield the least mass loss. In the 315–500°C region, the mass loss of the cured-control AHPCS, PACS-1, PACS-3, and PACS-5 was 27.9, 19.9, 11.1%, and 10.4%, respectively. In the 500–900°C region, the mass loss of the cured control AHPCS (7.3%) and those of the cured PACSs (7.0–10.0%) was similar. There was no obvious mass loss for the cured control AHPCS and PACSs in the 900–1,200°C region, which indicates that a polymer-to-ceramic conversion was achieved. Differences in mass loss resulted between the cured PACSs with control AHPCS for 315–500°C, which may yield differences in final ceramic yields. Based on the literature [29,35], the evolution of volatile gases CH3CH3, SiH4, and CH3SiH3 may result in the AHPCS mass loss over 300–500°C. The evolution of these three gas types is inhibited in the PACSs, which leads to less mass loss.

Figure 8 
                  TGA curves of the cured control AHPCS and PACSs.
Figure 8

TGA curves of the cured control AHPCS and PACSs.

3.3 Phase composition and ceramic microstructure

XRD was used to examine the crystallization behavior of the PACS-derived ceramics. The variation in crystallization behavior of the pyrolytic products of cross-linked PACS-5 at 900, 1,200, 1,400, 1,600, and 1,800°C is shown in Figure 9. The sample pyrolyzed at 900°C was amorphous and highly disordered. The sample annealed at 1,200 and 1,400°C exhibited a broad peak at 35.6°, because of incomplete crystallization and the formation of local-order SiC4. Heating at 1,600°C yielded three major peaks at 2θ = 35.6°, 60.0°, and 72°, which is attributed to the (111), (220), and (311) lattice planes of β-SiC. The shoulder at 34° corresponds to stacking faults like α-SiC in β-SiC [34]. With an increase in annealed temperature, the β-SiC peaks sharpen and the crystal size increases, which indicates an increased degree of crystallization, as expected.

Figure 9 
                  XRD patterns of PACS-5-derived ceramics at different temperatures.
Figure 9

XRD patterns of PACS-5-derived ceramics at different temperatures.

The detailed chemical compositions of PACS-derived ceramics at different temperatures were determined by bulk chemical analysis, and the result is shown in Table 1. Ceramics derived from PACS-5 at 900, 1,200, 1,400, 1,600, and 1,800°C were denoted PACS-5-900°C, PACS-5-1,200°C, PACS-5-1,400°C, PACS-5-1,600°C, and PACS-5-1,800°C, respectively. The oxygen content decreases significantly when the ceramic was annealed at 1,400°C, which indicates a decomposition of the Si–O–Si phase. With decomposition, the crystallization of resultant ceramics starts at 1,400°C, as shown in Figure 9.

Table 1

Chemical composition of the 1,600°C ceramics from bulk chemical analysis

Sample Si element contenta C element contentb O element contentc Al element contentd Average formula
wt% at% wt% at% wt% at% wt% at%
PACS-5-900°C 49.03 32.87 24.32 38.04 22.12 25.94 4.53 3.15 SiC1.16O0.789Al0.096
PACS-5-1,200°C 47.33 31.17 26.42 40.61 21.91 25.25 4.34 2.97 SiC1.30O0.810Al0.095
PACS-5-1,400°C 48.46 32.57 26.35 41.32 17.87 21.01 7.32 5.10 SiC1.27O0.645Al0.156
PACS-5-1,600°C 55.69 39.15 26.18 42.94 9.36 11.51 8.77 6.40 SiC1.10O0.294Al0.163
PACS-5-1,800°C 57.28 41.47 25.88 43.72 4.21 5.33 12.63 9.48 SiC1.05O0.128Al0.228

aSi element content = 100% − (C element content + O element content + Al element content). bC element content was measured by detecting CO2 content in oxygen atmosphere and high temperature using Carbon/Sulfur Analyzer. cO element content was measured by detecting CO content in argon atmosphere and high temperature using Oxygen/Nitrogen Analyzer. dAl element content was measured by EDS.

The effect of the aluminum content in the feed on the 1,600 and 1,800°C ceramics was also investigated. Figures 10 and 11 show the XRD patterns of the 1,600 and 1,800°C ceramic samples and Figure 12 shows the average crystal size of the 1,600 and 1,800°C ceramic samples with different aluminum contents.

Figure 10 
                  XRD patterns of 1,600°C ceramics derived from (a) AHPCS, (b) PACS-1, (c) PACS-3, and (d) PACS-5.
Figure 10

XRD patterns of 1,600°C ceramics derived from (a) AHPCS, (b) PACS-1, (c) PACS-3, and (d) PACS-5.

Figure 11 
                  XRD patterns of 1,800°C ceramics derived from (a) AHPCS, (b) PACS-1, (c) PACS-3, and (d) PACS-5.
Figure 11

XRD patterns of 1,800°C ceramics derived from (a) AHPCS, (b) PACS-1, (c) PACS-3, and (d) PACS-5.

Figure 12 
                  Effect of aluminum content in feed on the average crystal size of the final ceramic at different temperatures.
Figure 12

Effect of aluminum content in feed on the average crystal size of the final ceramic at different temperatures.

The width of the (111) diffraction peak at mid-height could be applied to calculate the apparent mean grain size of the β-SiC crystalline phase in the sample according to the Scherrer equation [36]. As the aluminum content increases, the apparent mean grain size of the β-SiC decreases prominently, which indicates that SiC crystal growth was inhibited. Based on the previous study, the inhibition to crystal growth was found advantageous with respect to the thermal stability of the PDC at high temperature for the metal-modified SiC-based ceramics [7].

To analyze the 1,600°C ceramic composition, the EDS elemental analysis of the ceramics was measured (Figure 13). The characteristic peaks of Si, Al, O, and C in the EDS spectrum of the PACS-5-derived ceramic confirm the ceramic composition.

Figure 13 
                  EDS elemental analysis of 1,600°C ceramics derived from (a) AHPCS and (b) PACS-5.
Figure 13

EDS elemental analysis of 1,600°C ceramics derived from (a) AHPCS and (b) PACS-5.

The calculated Al contents of the Si/C/Al ceramics from EDS are shown in Figure 14. As the Al content in the PACS precursors increased, the Al content in the final ceramics increased linearly, which indicates that the Al content in the ceramic could be readily controlled by changing the Al(acac)3 feed in the precursor.

Figure 14 
                  Dependence of Al content in 1,600°C ceramics on Al content in PACS precursors.
Figure 14

Dependence of Al content in 1,600°C ceramics on Al content in PACS precursors.

Finally, the typical microstructures of the ceramics derived from AHPCS and PACS-5 annealed at 1,800°C were investigated by SEM (Figure 15). The AHPCS-derived ceramic surface is porous and unconsolidated. In contrast, the PACS-derived ceramic surface was dense. These results indicate that the final ceramic densification improved markedly by the introduction of aluminum into the AHPCS, because the aluminum and SiC forms a solid solution to activate sintering [7,14].

Figure 15 
                  SEM micrographs of ceramics derived from (a) AHPCS and (b) PACS-5 annealed at 1,800°C.
Figure 15

SEM micrographs of ceramics derived from (a) AHPCS and (b) PACS-5 annealed at 1,800°C.

4 Conclusion

We prepared a series of hyperbranched PACS precursors via the condensation reaction between AHPCS and Al(acac)3, which were characterized by GPC, FT-IR, 1H NMR, and 13C NMR. SiAlC(O) ceramics were obtained by a PDC route from the as-synthesized PACSs. Thermogravimetric analysis shows that the ceramic yields of the PACSs is enhanced significantly, possibly because the evaporation of low-molecular-weight oligomers decreased significantly after cross-linking between the AHPCS and Al(acac)3. The FT-IR results suggest an organic-to-inorganic transition and an amorphous to crystalline SiC conversion during PACS pyrolysis with cross-linking. By varying the Al(acac)3 content in the preceramic PACSs, the composition of the final ceramics can be controlled easily. The contribution of aluminum to SiC(O) ceramics includes: (i) suppressing β-SiC crystal growth and (ii) improving the density of final ceramics annealed at 1,800°C, which suggests that SiAlC(O) ceramics from polyaluminocarbosilane show excellent high-temperature behavior.

Acknowledgements

We thank the National Natural Science Foundation of China (No. 51872246), Creative Research Foundation of Science and Technology on Thermostructural Composite Materials Laboratory (No. 6142911040114).

  1. Funding information: National Natural Science Foundation of China (No. 51872246 ), Creative Research Foundation of Science and Technology on Thermo-structural Composite Materials Laboratory (No. 6142911040114).

  2. Author contributions: Yang Le: performed the experiment and data analyses, wrote the manuscript; Zhang Pei: performed the experiment; Feng Yao: performed the experiment and edited the English text of a draft of manuscript; Yu Zhaoju: conducted the experiments, contributed significantly to analysis and manuscript preparation.

  3. Conflict of interest: The authors state no conflict of interest.

  4. Data availability statement: All authors can confirm that all data used in this article can be published in High Temperature Materials and Processes.

References

[1] Yajima, S., J. Hayshi, M. Omori, and K. Okamura. Development of a silicon carbide fibre with high tensile strength. Nature, Vol. 261, 1976, pp. 683–685.10.1038/261683a0Search in Google Scholar

[2] Amorós, P., D. Beltrán, C. Guillem, and J. Latorre. Synthesis and characterization of SiC/MC/C ceramics (M = Ti, Zr, Hf) starting from totally non-oxidic precursors. Chemistry Materials, Vol. 14, 2002, pp. 1585–1590.10.1021/cm011200sSearch in Google Scholar

[3] Calombo, P., G. Mera, R. Riedel, and G. D. Sorarù. Polymer-derived ceramics: 40 years of research and innovation in advanced ceramics. Journal American Ceramics Society, Vol. 93, 2010, pp. 1805–1837.10.1111/j.1551-2916.2010.03876.xSearch in Google Scholar

[4] Gui, M. M., Y. H. Fang, and Z. J. Yu. Single-source-precursor synthesis and high-temperature behavior of SiC ceramics containing boron. High Temperature Materials and Processes, Vol. 33, 2014, pp. 563–570.10.1515/htmp-2013-0123Search in Google Scholar

[5] Ishikama, T., Y. Kohtoku, and K. Kumagawa. Production mechanism of polyzirconocarbo-silane using zirconium (IV) acetylacetonate and its conversion of the polymer into inorganic materials. Journal Materials Science, Vol. 33, 1998, pp. 161–166.10.1023/A:1004362116892Search in Google Scholar

[6] Zaheer, M., T. Schmalz, G. Motz, and R. Kempe. Polymer derived non-oxide ceramics modified with late transition metals. Chemical Society Reviews, Vol. 41, 2012, pp. 5102–5116.10.1039/c2cs15326bSearch in Google Scholar PubMed

[7] Tanaka, H., Y. Inomata, K. Hara, and H. Hasegawa. Normal sintering of Al-doped β-SiC. Journal of Materials Science Letters, Vol. 4, 1985, pp. 315–317.10.1007/BF00719801Search in Google Scholar

[8] Yu, Y. X., X. D. Li, F. Cao, X. Xing, and C. X. Feng. Advances in the polymer-derived SiC fibers including hetero-element. Journal Chinese Ceramic Society, Vol. 31, 2003, pp. 371–375.Search in Google Scholar

[9] Rice, R. W. Ceramics from polymer pyrolysis, opportunities and needs- a materials perspective. American Ceramic Society Bulletin, Vol. 62, 1983, pp. 889–892.Search in Google Scholar

[10] Birot, M., J. Pillot, and J. Dunogues. Comprehensive chemistry of polycarbosilanes, polysilazanes, and polycarbosilazanes as precursors of ceramics. Chemistry Reviews, Vol. 95, 1995, pp. 1443–1477.10.1021/cr00037a014Search in Google Scholar

[11] Liu, Q., H. J. Wu, R. Lewis, G. E. Maciel, and L. V. Interrante. Investigation of the pyrolytic conversion of poly (silylenemethylene) to silicon carbide. Chemistry of Materials, Vol. 11, 1999, pp. 2038–2048.10.1021/cm981067cSearch in Google Scholar

[12] Ionescu, E., H. J. Kleebe, and R. Riedel. Silicon-containing polymer-derived ceramic nanocomposites (PDC-NCs): preparative approaches and properties. Chemical Society Reviews, Vol. 41, 2012, pp. 5032–5052.10.1039/c2cs15319jSearch in Google Scholar PubMed

[13] Sorarù, G. D., M. Mercadini, R. D. Maschio, F. Taulelle, and F. Babonneau. Si-Al-O-N fibers from polymeric precursor: synthesis, structural, and mechanical characterization. Journal of the American Ceramic Society, Vol. 76, 1993, pp. 2595–2600.10.1111/j.1151-2916.1993.tb03986.xSearch in Google Scholar

[14] Ishikawa, T., Y. Kohtoko, K. Kumagawa, T. Yamamura, and T. Nagasawa. High-strength alkali-resistant sintered SiC fibre stable to 2,200 C. Nature, Vol. 391, 1998, pp. 773–775.10.1038/35820Search in Google Scholar

[15] Cao, F., D. P. Kim, X. D. Li, C. X. Feng, and Y. C. Song. Synthesis of polyaluminocarbosilane and reaction mechanism study. Journal of Applied Polymer Science, Vol. 85, 2002, pp. 2787–2792.10.1002/app.10781Search in Google Scholar

[16] Cao, F., X. D. Li, P. Peng, C. X. Feng, J. Wang, and D. P. Kim. Structural evolution and associated properties on conversion from Si-CO-Al ceramic fibers to Si-C-Al fibers by sintering. Journal of Materials Chemistry, Vol. 12, 2002, pp. 606–610.10.1039/b106868gSearch in Google Scholar

[17] Yu, Y. X., Y. Zhang, J. M. Yang, and M. Tang. Synthesis and characterization of ceramic precursor aluminum-containing polycarbosilane and its pyrolysis. Journal of Inorganic and Organometallic Polymers and Materials, Vol. 17, 2007, pp. 569–575.10.1007/s10904-007-9143-2Search in Google Scholar

[18] Yu, Y. X., X. D. Li, and F. Cao. Synthesis and characterization of polyaluminocarbosilane. Journal of Materials Science, Vol. 40, 2005, pp. 2093–2095.10.1007/s10853-005-1243-1Search in Google Scholar

[19] Yu, Y. X. and X. Y. Tang. Ceramic precursor aluminum-containing polycarbosilane: Preparation and structure. Journal of Inorganic and Organometallic Polymers and Materials, Vol. 19, 2009, pp. 389–394.10.1007/s10904-009-9284-6Search in Google Scholar

[20] Interrante, L. V. and Q. H. Shen. Silicon-containing dendritic polymers, P.R. Dvornic and Owen, M. J., (ed.), Springer Science + Business Media B.V., Dordrecht, 2009, pp. 315–343.Search in Google Scholar

[21] Huang, T. H., Z. J. Yu, X. M. He, M. H. Huang, L. F. Chen, H. P. Xia, et al. One-pot synthesis and characterization of a new, branched polycarbosilane bearing allyl groups. Chinese Chemical Letters, Vol. 18, 2007, pp. 754–757.10.1016/j.cclet.2007.04.006Search in Google Scholar

[22] Huang, M. H., Y. H. Fang, R. Li, T. H. Huang, Z. J. Yu, and H. P. Xia. Synthesis and properties of liquid polycarbosilanes with hyperbranched structures. Journal of Applied Polymer Science, Vol. 113, 2009, pp. 1611–1618.10.1002/app.30071Search in Google Scholar

[23] Fang, Y. H., M. H. Huang, Z. J. Yu, H. P. Xia, L. F. Chen, and Y. Zhang. Synthesis, characterization, and pyrolytic conversion of a novel liquid polycarbosilane. Journal of the American Ceramic Society, Vol. 91, 2008, pp. 3298–3302.10.1111/j.1551-2916.2008.02603.xSearch in Google Scholar

[24] Yu, Z. J., R. Li, J. Y. Zhan, C. Zhou, L. Yang, and G. M. He. Synthesis and characterization of a propargyl-substituted polycarbosilane with high ceramic yield. Journal of Applied Polymer Science, Vol. 121, 2011, pp. 3400–3406.10.1002/app.33838Search in Google Scholar

[25] Li, H. B., L. T. Zhang, L. F. Cheng, Y. G. Wang, Z. J. Yu, M. H. Huang, et al. Polymer–ceramic conversion of a highly branched liquid polycarbosilane for SiC-based ceramics. Journal of Materials Science, Vol. 43, 2008, pp. 2806–2811.10.1007/s10853-008-2539-8Search in Google Scholar

[26] Li, H. B., L. T. Zhang, L. F. Cheng, Y. G. Wang, Z. J. Yu, M. H. Huang, et al. Effect of the polycarbosilane structure on its final ceramic yield. Journal of the European Ceramic Society, Vol. 28, 2008, pp. 887–891.10.1016/j.jeurceramsoc.2007.07.020Search in Google Scholar

[27] Li, H. B., L. T. Zhang, L. F. Cheng, Z. J. Yu, M. H. Huang, H. B. Tu, et al. Effect of curing and pyrolysis processing on the ceramic yield of a highly branched polycarbosilane. Journal of Materials Science, Vol. 44, 2009, pp. 721–725.10.1007/s10853-008-3176-ySearch in Google Scholar

[28] Yu, Z. J., H. Min, J. Y. Zhan, and L. Yang. Preparation and dielectric properties of polymer-derived SiCTi ceramics. Ceramics International, Vol. 39, 2013, pp. 3999–4007.10.1016/j.ceramint.2012.10.250Search in Google Scholar

[29] Yu, Z. J., L. Yang, J. Y. Zhan, C. Zhou, H. Min, Q. Zheng, et al. Preparation, cross-linking and ceramization of AHPCS/Cp2ZrCl2 hybrid precursors for SiC/ZrC/C composites. Journal of the European Ceramic Society, Vol. 32, 2012, pp. 1291–1298.10.1016/j.jeurceramsoc.2011.12.015Search in Google Scholar

[30] Yu, Z. J., L. Yang, H. Min, P. Zhang, C. Zhou, and R. Riedel. Single-source-precursor synthesis of high temperature stable SiC/C/Fe nanocomposites from a processable hyperbranched polyferrocenylcarbosilane with high ceramic yield. Journal of Materials Chemistry C, Vol. 2, 2014, pp. 1057–1067.10.1039/C3TC32088JSearch in Google Scholar

[31] Wen, Q. B., Y. P. Xu, B. B. Xu, C. Fasel, O. Guillon, G. Buntkowsky, et al. Single-source-precursor synthesis of dense SiC/HfC(x)N(1-x)-based ultrahigh-temperature ceramic nanocomposites. Nanoscale, Vol. 6, 2014, pp. 13678–13689.10.1039/C4NR03376KSearch in Google Scholar PubMed

[32] Yu, Z. J., P. Zhang, Y. Feng, S. Li, and Y. X. Pei. Template-free synthesis of porous Fe3O4/SiOC(H) nanocomposites with enhanced catalytic activity. Journal of the American Ceramic Society, Vol. 99, No. 8, 2016, pp. 2615–2624.10.1111/jace.14305Search in Google Scholar

[33] Shriver, D. F. and M. A. Drezdzon. The manipulation of Air-Sensitive Compounds, 2nd edn, Wiley, New York, 1986.Search in Google Scholar

[34] Yamaoka, H., T. Ishikawa, and K. Kumagawa. Excellent heat resistance of Si-Zr-CO fibre. Journal of Materials Science, Vol. 34, 1999, pp. 1333–1339.10.1023/A:1004506300754Search in Google Scholar

[35] Yamamura, T., T. Ishikawa, and M. Shibuya. Transition metal sprinkle polymer containing silicon and titanium or zirconium and its production method. Japan Patent, No. 1699835, 1992.Search in Google Scholar

[36] Mocaer, D., R. Pailler, R. Naslain, C. Richard, J. P. Pillot, J. Dunogues, et al. Si-CN ceramics with a high microstructural stability elaborated from the pyrolysis of new polycarbosilazane precursors. Journal of Materials Science, Vol. 28, 1993, pp. 2615–2631.10.1007/BF00356196Search in Google Scholar

Received: 2018-09-20
Accepted: 2019-02-28
Published Online: 2022-04-06

© 2022 Le Yang et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Numerical and experimental research on solidification of T2 copper alloy during the twin-roll casting
  3. Discrete probability model-based method for recognition of multicomponent combustible gas explosion hazard sources
  4. Dephosphorization kinetics of high-P-containing reduced iron produced from oolitic hematite ore
  5. In-phase thermomechanical fatigue studies on P92 steel with different hold time
  6. Effect of the weld parameter strategy on mechanical properties of double-sided laser-welded 2195 Al–Li alloy joints with filler wire
  7. The precipitation behavior of second phase in high titanium microalloyed steels and its effect on microstructure and properties of steel
  8. Development of a huge hybrid 3D-printer based on fused deposition modeling (FDM) incorporated with computer numerical control (CNC) machining for industrial applications
  9. Effect of different welding procedures on microstructure and mechanical property of TA15 titanium alloy joint
  10. Single-source-precursor synthesis and characterization of SiAlC(O) ceramics from a hyperbranched polyaluminocarbosilane
  11. Carbothermal reduction of red mud for iron extraction and sodium removal
  12. Reduction swelling mechanism of hematite fluxed briquettes
  13. Effect of in situ observation of cooling rates on acicular ferrite nucleation
  14. Corrosion behavior of WC–Co coating by plasma transferred arc on EH40 steel in low-temperature
  15. Study on the thermodynamic stability and evolution of inclusions in Al–Ti deoxidized steel
  16. Application on oxidation behavior of metallic copper in fire investigation
  17. Microstructural study of concrete performance after exposure to elevated temperatures via considering C–S–H nanostructure changes
  18. Prediction model of interfacial heat transfer coefficient changing with time and ingot diameter
  19. Design, fabrication, and testing of CVI-SiC/SiC turbine blisk under different load spectrums at elevated temperature
  20. Promoting of metallurgical bonding by ultrasonic insert process in steel–aluminum bimetallic castings
  21. Pre-reduction of carbon-containing pellets of high chromium vanadium–titanium magnetite at different temperatures
  22. Optimization of alkali metals discharge performance of blast furnace slag and its extreme value model
  23. Smelting high purity 55SiCr automobile suspension spring steel with different refractories
  24. Investigation into the thermal stability of a novel hot-work die steel 5CrNiMoVNb
  25. Residual stress relaxation considering microstructure evolution in heat treatment of metallic thin-walled part
  26. Experiments of Ti6Al4V manufactured by low-speed wire cut electrical discharge machining and electrical parameters optimization
  27. Effect of chloride ion concentration on stress corrosion cracking and electrochemical corrosion of high manganese steel
  28. Prediction of oxygen-blowing volume in BOF steelmaking process based on BP neural network and incremental learning
  29. Effect of annealing temperature on the structure and properties of FeCoCrNiMo high-entropy alloy
  30. Study on physical properties of Al2O3-based slags used for the self-propagating high-temperature synthesis (SHS) – metallurgy method
  31. Low-temperature corrosion behavior of laser cladding metal-based alloy coatings on EH40 high-strength steel for icebreaker
  32. Study on thermodynamics and dynamics of top slag modification in O5 automobile sheets
  33. Structure optimization of continuous casting tundish with channel-type induction heating using mathematical modeling
  34. Microstructure and mechanical properties of NbC–Ni cermets prepared by microwave sintering
  35. Spider-based FOPID controller design for temperature control in aluminium extrusion process
  36. Prediction model of BOF end-point P and O contents based on PCA–GA–BP neural network
  37. Study on hydrogen-induced stress corrosion of 7N01-T4 aluminum alloy for railway vehicles
  38. Study on the effect of micro-shrinkage porosity on the ultra-low temperature toughness of ferritic ductile iron
  39. Characterization of surface decarburization and oxidation behavior of Cr–Mo cold heading steel
  40. Effect of post-weld heat treatment on the microstructure and mechanical properties of laser-welded joints of SLM-316 L/rolled-316 L
  41. An investigation on as-cast microstructure and homogenization of nickel base superalloy René 65
  42. Effect of multiple laser re-melting on microstructure and properties of Fe-based coating
  43. Experimental study on the preparation of ferrophosphorus alloy using dephosphorization furnace slag by carbothermic reduction
  44. Research on aging behavior and safe storage life prediction of modified double base propellant
  45. Evaluation of the calorific value of exothermic sleeve material by the adiabatic calorimeter
  46. Thermodynamic calculation of phase equilibria in the Al–Fe–Zn–O system
  47. Effect of rare earth Y on microstructure and texture of oriented silicon steel during hot rolling and cold rolling processes
  48. Effect of ambient temperature on the jet characteristics of a swirl oxygen lance with mixed injection of CO2 + O2
  49. Research on the optimisation of the temperature field distribution of a multi microwave source agent system based on group consistency
  50. The dynamic softening identification and constitutive equation establishment of Ti–6.5Al–2Sn–4Zr–4Mo–1W–0.2Si alloy with initial lamellar microstructure
  51. Experimental investigation on microstructural characterization and mechanical properties of plasma arc welded Inconel 617 plates
  52. Numerical simulation and experimental research on cracking mechanism of twin-roll strip casting
  53. A novel method to control stress distribution and machining-induced deformation for thin-walled metallic parts
  54. Review Article
  55. A study on deep reinforcement learning-based crane scheduling model for uncertainty tasks
  56. Topical Issue on Science and Technology of Solar Energy
  57. Synthesis of alkaline-earth Zintl phosphides MZn2P2 (M = Ca, Sr, Ba) from Sn solutions
  58. Dynamics at crystal/melt interface during solidification of multicrystalline silicon
  59. Boron removal from silicon melt by gas blowing technique
  60. Removal of SiC and Si3N4 inclusions in solar cell Si scraps through slag refining
  61. Electrochemical production of silicon
  62. Electrical properties of zinc nitride and zinc tin nitride semiconductor thin films toward photovoltaic applications
  63. Special Issue on The 4th International Conference on Graphene and Novel Nanomaterials (GNN 2022)
  64. Effect of microstructure on tribocorrosion of FH36 low-temperature steels
Downloaded on 14.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/htmp-2022-0025/html
Scroll to top button