Synthesis, characterization, thermal properties of silicon(IV) compounds containing guanidinato ligands and their potential as CVD precursors

Liyong Dua, Wenxiang Chua, Hongyan Miaoa, Chongying Xu*ab and Yuqiang Ding*a
aThe Key Laboratory of Food Colloids and Biotechnology, Ministry of Education, School of Chemical and Material Engineering, Jiangnan University, Wuxi 214122, Jiangsu Province, China. E-mail: yding@jiangnan.edu.cn
bJiangsu Nata Opto-Electronic Material Co. Ltd., 7F One Lakepoint, No. 9 Cuiwei Street, Suzhou Industrial Park, Jiangsu Province, China

Received 24th May 2015 , Accepted 18th August 2015

First published on 19th August 2015


Abstract

The title compounds of the type (Me3Si)2N–C([double bond, length as m-dash]N′R)(–N′′RSiMe3) (with R = iPr or Cy) as potential CVD precursors have been synthesized and characterized by X-ray diffraction, 1H NMR, 13C NMR, 29Si NMR and elemental analysis where necessary. Among these characterizations, solid-/liquid-state 29Si NMR were accomplished to study their behavior in the solid and solution. Thermal properties including stability, volatility, transport behavior and vapour pressure were evaluated by thermogravimetric analysis (TGA) to confirm that they are suitable for the CVD procedure. Deposition was accomplished in a hot wall CVD reactor system, which preliminarily verified the ability of these compounds as CVD precursors.


Introduction

Guanidines/amidinates are N-donor bidentate chelating ligands, which demonstrate a great diversity by the variation of substituents on the conjugated N–C–N backbone.1 The steric and electronic properties are easily modified to meet the needs of different metal centers.1 Corresponding compounds have shown enormous potential as catalysts for the oligomerization and polymerization of olefins, hydroamination, intramolecular hydroamination/cyclization and hydrosilylation.2

Meanwhile, the numerous features and widespread applications have made guanidinato/amidinate an important ligand for precursors used in chemical vapor deposition (CVD)/atomic layer deposition (ALD) manufacturing process. For example, Anjana Devi and coworkers obtained nitrogen-doped CVD-grown TiO2 thin films using [Ti(NMe2)3(guan)] (guan = N,N′-diisopropyl-2-dimethylamidoguanidinato) as precursor.3 Tianniu Chen introduced chelating guanidinate ligands improving thermal stability and maintaining the volatility of precursors for CVD tantalum nitride thin films.4 Copper(I)amidinate [Cu(i-Pr-Me-AMD)]2 was investigated to produce copper films in conventional low pressure chemical vapor deposition (LPCVD) using hydrogen as reducing gas-reagent.5

In contrast, the chemistry of silicon compounds containing guanidinato/amidinate ligands is still a largely unexplored field. To date, a few silicon compounds containing guanidinato/amidinate ligands have been reported (Scheme 1).6 All of these researches were merely focused on synthetic methodology or molecular geometry. Few literatures, however, involved the application of these compounds as CVD/ALD precursors and relevant properties.


image file: c5ra09755j-s1.tif
Scheme 1 Reported silicon compounds with guanidinato/amidinate ligands.

As known, many silicon compounds have been investigated to grow silicon-based thin films such as silicon nitride (SiNx),7 amorphous silicon (α-Si),8 silicon carbide (SiC),9 polycrystalline silicon10 and silicon carbonitride (SiCN),11 which have shown a wide range of properties and have various applications in microelectromechanical systems (MEMS) technology. Apart from the industrialized silanes (SiH4 and Si2H6), many hydrosilicons linked with alkyl and/or amino ligands have been reported, such as SiH3CH3, Et3SiH, SiH2(CH3)2, SiH(NMe2)3, SiMe2(NMe2)2 and (SiHMe2)2NH.12 Insufficiently, the introduction of gaseous reactant Me2NH or Si–H units has increased the synthetic difficulty, thus hindering their further application. Furthermore, the toxicity and transportation difficulties of commonly-used SiH4/Si2H6 have always been a big trouble for the users.13

The search for new silicon precursors is still a great challenge not only for the above-mentioned reasons but also because the chemical and physical properties of the Si source can be adjusted by various ligands/substituents in order to obtain films with many particular properties and thereby satisfy different application requirements.14 Thus, we have recently been intrigued by the question as to whether such silicon compounds containing guanidinato/amidinate ligands are capable for a robust CVD/ALD process.

On the basic of these studies, herein, we report the synthesis of two new silicon compounds (Scheme 2) with guanidinato ligands, their thermal properties and further potential as CVD precursors.


image file: c5ra09755j-s2.tif
Scheme 2 Synthesis of compounds 1 and 2.

Compared with the existing CVD precursors mentioned above, many foreseeable advantages can be obtained through the design and synthesis of these compounds as CVD precursors: (i) the synthesis method is well known as a salt elimination approach that is easily accomplished; (ii) the specific –SiMe3 groups are beneficial to enhance the volatility of these compounds;15 (iii) many guanidinato/amidinate ligands are either easily sublimated solid or volatile liquid that will facilitate the desorption of the byproducts from the surface;16 and (iv) they are nontoxic, non-explosive and non-pyrophoric and are thus readily applicable to a robust CVD system.

Results and discussion

Synthesis and characterization

Compounds 1 and 2 were obtained by reaction of lithium intermediates, which originated from the reaction of corresponding carbodiimides and LiN(SiMe3)2, with equivalent SiMe3Cl according to a general procedure. The synthetic approach involved the formation of lithium intermediates and LiCl elimination, which proceeded easily under benign conditions in diethyl ether.

Between these two precursors, compound 1 is low melting solid (mp 48–50 °C), while 2 melts at a higher temperature (mp 150–153 °C). Both of them were recrystallized out of n-hexane at −30 °C for structural determination and purified further through sublimation under vacuum. Furthermore, both elemental analysis of 1 and 2 are consistent with the proposed monomeric structures.

One ingenious idea of the introduction of –N(SiMe3)2 groups is for convenient and unambiguous learning of coordination states of these two central silicon atoms (Si3 for 1 and Si1 for 2).

The single-crystal X-ray diffraction results (Fig. 1 and 2) confirm the monomeric nature of them. Characteristic crystal parameters of these compounds 1 and 2 are listed in Table 1.


image file: c5ra09755j-f1.tif
Fig. 1 ORTEP diagram of 1 showing 40% thermal ellipsoids. Selected bond lengths [Å] and angles [deg]: C1–N1 1.451(4), C1–C5 1.519(4), C2–N1 1.276(3), C2–N2 1.395(4), C2–N3 1.433(3), C3–N2 1.478(4), C6–Si1 1.844(4), C9–Si2 1.841(4), C12–Si3 1.876(4), N2–Si3 1.777(3), N3–Si2 1.754(2), N3–Si1 1.754(2); N1–C2–N2 114.6(3), N1–C2–N3 125.7(3), N2–C2–N3 119.7(2), C2–N1–C1 123.7(3), C2–N2–C3 121.1(2), C2–N2–Si3 111.79(19), C3–N2–Si3 127.1(2), C2–N3–Si2 117.36(18), C2–N3–Si1 117.43(17), Si2–N3–Si1 125.05(14), N3–Si1–C7 110.54(17), N3–Si2–C9 109.69(15), N2–Si3–C12 109.78(16), C14–Si3–C12 103.41(19).

image file: c5ra09755j-f2.tif
Fig. 2 ORTEP diagram of 2 showing 40% thermal ellipsoids. Selected bond lengths [Å] and angles [deg]: C1–N1 1.457(3), C1–C2 1.508(4), C7–N1 1.278(3), C7–N2 1.396(3), C7–N3 1.436(3), C8–N2 1.475(3), C14–Si1 1.851(4), C17–Si3 1.854(4), C20–Si2 1.847(4), N2–Si1 1.782(2), N3–Si3 1.752(2), N3–Si2 1.759(2); N1–C7–N2 114.7(2), N1–C7–N3 125.9(2), N2–C7–N3 119.37(19), C7–N1–C1 122.9(2), C7–N2–C8 121.13(19), C7–N2–Si1 111.24(15), C8–N2–Si1 127.35(16), C7–N3–Si3 117.79(15), C7–N3–Si2 117.01(15), Si3–N3–Si2 124.99(12), N2–Si1–C14 111.26(15), N3–Si2–C20 112.56(16), –N3–Si2–C21 111.30(16), N3–Si3–C19 109.89(13).
Table 1 Crystal parameters of 1 and 2
  1 2
Empirical formula C16H41N3Si3 C22H49N3Si3
Molecular weight 359.79 439.91
Temperature (K) 296 296
Wavelength Mo Kα (Å) 0.71073 0.71073
Crystal size (mm) 0.27 × 0.30 × 0.32 0.22 × 0.24 × 0.26
Crystal system, space group Triclinic, P[1 with combining macron] Triclinic, P[1 with combining macron]
a (Å) 9.331(4) 9.0667(12)
b (Å) 9.361(4) 11.4557(15)
c (Å) 14.996(7) 14.1193(19)
α (deg) 99.880(11) 100.930(3)
β (deg) 93.805(12) 102.475(2)
γ (deg) 109.819(10) 91.188(2)
Cell volume (Å3)/Z 1203.0(9)/2 1403.0(3)/2
Density δcalcd (g cm−3) 0.993 1.041
Absorption coefficient μ (mm−1) 0.199 0.181
F(000) 400 488
θ range for data collection (deg) 1.4–26.3 1.5–25.0
Completeness to θ = full [%] 98.3 98.6
Index ranges −11 ≤ h ≤ 10, −10 ≤ k ≤ 11, −18 ≤ l ≤ 18 −10 ≤ h ≤ 10, −12 ≤ k ≤ 13, −16 ≤ l ≤ 15
Reflections collected/unique [R(int)] 9120/4807/0.043 9737/4885/0.028
Refinement method Full-matrix least-squares on F2 Full-matrix least-squares on F2
Data/restraints/parameters 4807/0/212 4885/0/262
Goodness-of-fit on F2 (GOF) 1.038 1.089
Observed data [I > 2.0 × sigma(I)] 2592 3867
R, wR2, S 0.0607, 0.1834, 1.04 0.0574, 0.2258, 1.09
Largest difference peak and hole (e Å−3) −0.30, 0.21 −0.43, 0.42
CCDC number 1400899 1033427


For every compound, each focal silicon center (Si3 for 1 and Si1 for 2) is chelated with one nitrogen atom from the amidinate units. The coordination environment around these two silicon centers is approximately tetrahedral (∠N–Si–C and ∠C–Si–C values range from 101.4 to 113.2°). Moreover, they are almost coplanar with NCN atoms of respective amidinate units (dihedral angle N1C2N2Si3 0.8° and N1C7N2Si1 1.4°).

The significant difference between the two N–C distances in both of the amidinate units [1.39 vs. 1.27 Å] indicates the higher degree of localization of the N[double bond, length as m-dash]C double bond compared to the bidentate guanidinato ligands (1.33 vs. 1.34 Å).6d Meanwhile, Si3/Si1–N2 distance (1.777 and 1.782 Å) is significantly shorter than those of compounds with bidentate guanidinato ligands reported by Reinhold Tacke6d (all approximately 1.9 Å), clearly reflecting the different coordination modes. This Si3/Si1–N2 distance, however, is slightly longer than that of –N(SiMe3)2 (approximately 1.75 Å of 1 and 2) and the monodentate ligand (1.76 Å), which implies the interaction between focal silicon center and the ligand. Furthermore, the medium-strong Si–N interactions are emphasized by the Si3–N1/Si1–N1 distances (2.6324 Å for 1 and 2.6242 Å for 2), which are under the sum of the Si and N van der Waals radii of 3.48 Å.17

Besides, we recorded solid-state 29Si CP/MAS NMR spectrums of precipitated powder of them, which show one single peak at −0.489 ppm (1) and −0.543 ppm (2), respectively. These are in agreement with the solid-state structures observed by X-ray diffraction well, where all Si atoms possess almost the same chemical environment.

To elucidate the solution structure in the studied compounds, the conventional NMR parameters (chemical shifts of 1H, 13C, 29Si nuclei) were obtained in CDCl3. For compound 1, the 1H and 13C NMR spectra exhibited chemical shift equivalent N2C2N1 ligand resonances at room temperature. Though 1H NMR analysis of 2 was not distinguishable, 13C NMR spectra of it confirmed its equivalent resonances for N2C7N1 ligand with four singlets of cyclohexyl units. Thus, we concluded that both 1 and 2 behaved as symmetric structures in CDCl3, which disagreed with their solid state formation.

To further verify this conclusion, 29Si NMR was applied. Analogously, two separated signal for both of them (1: 3.5 and −7.4 ppm; 2: 3.8 and −7.4 ppm; 25 °C) indicated the variation in the chemical environment of focal silicon center (Si3 for 1 and Si1 for 2) and corresponding referential silicon in –N(SiMe3)2 groups. Moreover, in comparison to the tetra-coordinated referential silicon, the fact that resonances of focal silicon were shifted towards higher-field (at least to a small extend) revealed the intramolecular donor-coordination (N1 → Si3 for 1 and N1 → Si1 for 2).18

Based on the similarity of these two compounds, only 29Si NMR solvent effect of 2 was investigated. Compared with the result in CDCl3, there was no obvious difference when d8-toluene was used (3.6 and −7.0 ppm, 25 °C). An unexpected lower-field signal appeared in CD2Cl2, however, with the chemical shift of 7.4 ppm (another two: 3.8 and −7.6 ppm, 25 °C), which might suggest the dynamic equilibrium of the N1 → Si1 interaction mentioned above. Subsequently, VT-29Si NMR of compound 2 was accomplished in d8-toluene. At high temperature (40 °C) compound 2 had an extra silicon chemical shift of 7.0 ppm (another two: 3.6 and −7.0 ppm), suggesting a weakening of the N1 → Si1 coordination. That is, high temperature weakens the interaction which has also been found by other researchers.19 And when given a low temperature of −60 °C, both intrinsic 29Si signal tended to be closer to each other in a slight extent with shifts of 3.4 and −6.7 ppm, which implied the trend of them to merge as solid-state 29Si CP/MAS signal with the decrease of the temperature. Regrettably, lower temperature 29Si NMR failed to obtain because of instrument limit.

Combining all tests above, we conclude that compound 1 and 2 exhibit different nuclear magnetic resonance behaviors by solid state and in solution, which has been widely studied.20 Besides, a dynamic equilibrium exists revealing the intensity change of the interaction between Si and corresponding N.

Thermal properties of compounds 1 and 2

The study of the basic parameters of precursors has not only scientific but also of great practical importance for understanding the chemistry of precursor preparation and their long-term stability as well as for the calculation of proper CVD precursor supply into the reactor for the preparation of the appropriate device structure. In particular, volatility data and a detailed vapour pressure equation are essential for controlled precursor dosimetry.

Both of these two compounds are volatile and thermally stable enough to be purified by vacuum sublimation and their thermal characteristics were first studied and measured by simultaneous thermal analyses (STA) including thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC).

As seen in Fig. 3, both of compounds 1 and 2 are stable up to their melting temperature and exhibit monotonic weight loss with negligible residual mass remaining (0.22% and 0.79% respectively), in a single step. This demonstrates that compounds 1 and 2 have sufficient thermal stability for vaporization without significant decomposition. Besides, the entire weight loss of compounds 1 and 2 finish at 215 and 240 °C respectively, implying the comparable volatility of them.


image file: c5ra09755j-f3.tif
Fig. 3 Thermogravimetric curves of 1 and 2.

The temperature of 50% mass loss derived from TGA data (T50) has been demonstrated to correlate with the volatility of the sample and is used to compare relative volatilities (Table 2).21 The combination of lower T50 with lower onset temperature derived from TGA curve is usually a strong indication of good volatility of a precursor. Clearly, compound 1 possesses lower T50 (197 °C) and onset temperature (179 °C), which suggests its better volatility. Both temperatures increase as the molecular mass increases (1 < 2), which agrees with previous reports by other researchers. For example, the nearly linear relationship between T20 values and molecular weigh in tBu derivatives of gallium chalcogenide cubanes were reported by Barron.22 And the consistent result was obtained by T. Chen, in tantalum amido compounds.4

Table 2 STA data of compounds 1 and 2a
  Molecular weight (g mol−1) T50/°C Onset temperature/°C Residual/%
a Mass residuals were measured at 300 °C.
1 359.79 197 179 0.79
2 439.91 221 203 0.22


To further investigate the transport properties of them, isothermal TGA experiments were carried out, in which the temperature of the STA furnace was held at 150 °C for 600 minutes and the rates of the linear mass loss, presumably by evaporation instead of decomposition, were calculated from the TGA curves shown in Fig. 4. The fact that both DSC curves remain constant for these samples tested verifies the assumption that only evaporation contributes to the mass loss reflected by isothermal TGA curves. Compound 1 showed faster mass loss rate of 97.2 μg min−1 (22.9 μg min−1, 2) at 150 °C and purged by 60 sccm Ar, which coincided with its lower T50 and onset temperature and confirmed its better volatility between these two compounds. Significantly, the higher volatility of 1 over the other precursor most likely arises from its lower molecular mass and suggests the future strategy for designing new volatile CVD/ALD precursors.


image file: c5ra09755j-f4.tif
Fig. 4 Isothermal TGA of 1 and 2.

Vapour pressure data are very important for industrial process control and in manufacturing processes. Especially, the film growth rate can be limited by the precursor vapour pressure. If the vapour pressure of precursor compounds is known, then it would be easy to compare their volatilities and find the right evaporation temperature for each precursor so that the precursor flux would be maintained in the CVD reactor at the right level.

Herein, vapour pressure–temperature plots were obtained by thermogravimetry, which is considered a rapid and convenient technique for the determination of vapour pressure curves. The theoretical basis of the TG procedure is the Langmuir and Antoine equation, where benzoic acid was chosen as a standard.23

As seen from the PT curves (Fig. 5) for 1 and 2, compound 1 possesses higher vapor pressure at relatively low temperatures with almost the same range (90 °C). The plot of ln[thin space (1/6-em)]P against 1/T as straight lines presents a more comparable view. Compound 1 is shown to be more volatile at lower temperatures, and to have 1 Torr of vapor pressure at 121 °C. Compound 2 has a slightly lower volatility, and gives 1 Torr of vapor pressure at 156 °C. This also verifies the T50 results are reliable. However, both of these compounds can be expected to give reasonably high vapor pressures at normal bubbler temperatures, and the similarity of the slopes of their evaporation curves suggests that they experience similar intermolecular attraction during evaporation. Moreover, the above temperatures cover the interval important for the CVD process.24


image file: c5ra09755j-f5.tif
Fig. 5 Vapor pressure–temperature curves of 1 and 2.

On the basis of the above results, we consider both of them to be qualified precursor for a robust CVD process. And a similar assumption could be valid under the real CVD conditions.

Potential as CVD precursors

The straightforward way to qualitatively verify the potential of these compounds as CVD precursors is making them undergo a deposition process. Thus compound 1 was chosen as the representative to undergo a single-precursor deposition process at 900 °C. The as-grown film, with a thickness about 400 nm, showed comparable morphology without obvious agglomeration as displayed in ESI.

Besides, Fig. 6 displayed the XP survey spectra acquired from the surface of the as-grown film. The relative peak positions for the Si 2p and Si 2s are 103.7 and 154.08 eV respectively, being consistent with earlier data reported25 and thereby indicating the participation of the precursor in the film deposition. The oxygen observed here was assumed to be largely due to surface oxidation or contamination. Thus, we qualitatively conclude that compounds 1 and 2 with guanidinato ligands have the potential to be CVD precursors. Certainly, further work should be done in regard to the deposition process such as optimization of conditions, quantitative analysis of the film component and mechanism. All these will be presented in the subsequent report.


image file: c5ra09755j-f6.tif
Fig. 6 XP spectra of the as-grown film.

Conclusions

In this study, we have synthetized and characterized two silicon(IV) compounds containing guanidinato ligands of the type (Me3Si)2N–C([double bond, length as m-dash]N′R)(–N′′RSiMe3) (with R = iPr or Cy) as CVD precursors. They were obtained by reaction of lithium intermediates, which originated from the reaction of corresponding carbodiimides and LiN(SiMe3)2, with equivalent SiMe3Cl according to a general procedure in high yields. Thermal stability, volatility, transport behavior and vapor pressure were evaluated by simultaneous thermal analyses (STA) to confirm that they are suitable for CVD process. Deposition which was accomplished in a simple hot wall CVD reactor system, qualitatively demonstrated the ability of them as CVD precursors for fabricating silicon containing films.

Moreover, further work with respect to the deposition process such as optimization of conditions, quantitative analysis of the film component, mechanism and deposition along with other auxiliary source (e.g. NH3) to form different kinds of films will be presented in the subsequent report elsewhere.

Experiment section

All manipulations were carried out using standard Schlenk techniques or in a glove box under a nitrogen atmosphere. Prior to use, diethyl ether and n-hexane was freshly distilled from Na. Other chemicals were purchased from Aldrich and used as received. X-ray data was collected with a Bruker SMART APEX II CCD area detector using monochromated Mo-Kα radiation (λ = 0.71073 Å). NMR spectrum was collected on a Bruker ACF-400 spectrometer tetramethylsilane as internal standard (0.03% content as purchased). The C, H, N analyses were performed with a Carlo Erba model EA 1108 microanalyzer. LiN(SiMe3)2 was prepared according to the reported procedure.26 The TG curve was obtained with an STA 449 F3 analyzer in argon at a heating rate of 10 °C min−1 from 30 to 800 °C. The surface morphology and X-ray photoelectron spectroscopy (XPS) analysis were characterized by a Hitachi S-4800 scanning electron microscopy (SEM) and a Thermo ESCALAB 250Xi equipment with a residual pressure of 10−7 Pa.

Synthesis of 1 and 2

Due to the similarities of preparing these two compounds, only a detailed synthetic procedure for 1 is described here. To a 100 mL Schlenk flask charged with LiN(SiMe3)2 (1.673 g, 10 mmol) in 40 mL diethyl ether, diisopropylcarbodiimide (1.262 g, 10.0 mmol) was added. After a stir for 24 h, SiMe3Cl (1.086 g, 10.0 mmol) was added at −78 °C. A white precipitate appeared after the addition started. The reaction mixture was warmed to room temperature and stirred overnight. After removing diethyl ether by vacuum, the residue was extracted by n-hexane (30 mL × 3). Then the filtrate was concentrated and recrystallized out of n-hexane at −30 °C for a few hours to yield colourless crystal 1 (3.274 g, 91% yield, mp 48–50 °C): 1H NMR (400 MHz, CDCl3) δ 3.92–3.82 (m, 2H, –CH(CH3)2), 1.12–1.10 (d, J = 8 Hz, 12H, –CH(CH3)2), 0.24 (s, 9H, –Si(CH3)3), 0.19 (s, 18H, –N[Si(CH3)3]2); 13C NMR (101 MHz, CDCl3) δ 153.68, 45.98, 24.73, 5.06, 2.20; 29Si NMR (79 MHz, CDCl3, 25 °C) δ 3.54, −7.40; 29Si CP/MAS NMR δ −0.489; anal. calcd for C16H41N3Si3: C, 53.41; H, 11.49; N, 11.68; found: C, 53.39; H, 11.58; N, 11.65.

2 (colourless crystal, 83% yield, mp 150–153 °C): 1H NMR (400 MHz, CDCl3) δ 3.44–3.37 (m, 2H, N–CH), 1.74–1.58 (m, 8H, Cy), 1.51–1.41 (m, 4H, Cy), 1.28–1.05 (m, 9H, Cy), 0.24 (s, 9H, –Si(CH3)3), 0.20 (s, 18H, –N[Si(CH3)3]2); 13C NMR (101 MHz, CDCl3) δ 153.92, 55.41, 34.64, 26.17, 25.86, 5.33, 2.34; 29Si NMR (79 MHz, CDCl3, 25 °C) δ 3.77, −7.36; 29Si CP/MAS NMR δ −0.543; 29Si NMR (79 MHz, CD2Cl2, 25 °C) δ 7.41, 3.81, −7.61; 29Si NMR (79 MHz, d8-toluene, 25 °C) δ 3.59, −7.03; 29Si NMR (79 MHz, d8-toluene, 40 °C) δ 7.04, 3.65, −6.99; 29Si NMR (79 MHz, d8-toluene, −60 °C) δ 3.43, −6.67; anal. calcd for C22H49N3Si3: C, 60.07; H, 11.23; N, 9.55; found: C, 60.11; H, 10.98; N, 9.71.

Thermal analysis and CVD deposition

The TG curve was obtained with an STA 449 F3 analyzer in argon at a heating rate of 10 °C min−1 from 30 to 800 °C. Subsequently, chemical vapor deposition was accomplished in a simple system shown schematically in ESI, to further demonstrate the ability of them as CVD precursors, where 1 was chosen as the representative.

The furnace consists of a hot-wall tubular quartz reactor with a large (≈60 cm) isothermal (±5 °C) zone. The substrates used were 1 cm × 1 cm N-type Si (100) & (110) & (110) wafers doped with P with a resistivity of 0.01–0.02 Ω cm. Prior to the deposition, the wafers were treated by a H2SO4 + H2O2 (v/v = 7/3) mixture followed by HF etching, thoroughly rinsed with ultrapure water and then dried with N2. The growth parameters used during the process were: total pressure: 2.1 × 103 Pa, 10% H2/N2 flow: 120 mL min−1, N2 flow: 120 mL min−1, deposition time: 180 min and temperature 900 °C respectively. A 150 °C heating temperature of 1 was kept during the whole deposition. After deposition, samples were cooled to room temperature at a rate of 5 °C min−1 in a N2 (flow: 120 mL min−1) atmosphere.

The surface morphology and X-ray photoelectron spectroscopy (XPS) analysis were characterized by a Hitachi S-4800 scanning electron microscopy (SEM) and a Thermo ESCALAB 250Xi equipment with a residual pressure of 10−7 Pa.

Acknowledgements

We gratefully acknowledge financial support of this work by the National Natural Science Foundation of China (No. 21371080) and “333 Talent Project” of Jiangsu Province (BRA2012165).

Notes and references

  1. F. T. Edelmann, in Advances in the Coordination Chemistry of Amidinate, Guanidinate Ligands, ed. A. F. Hill and M. J. Fink, Elsevier Academic Press Inc, San Diego, CA, 2008, vol. 57, pp. 183–352 Search PubMed.
  2. (a) E. Smolensky and M. S. Eisen, Dalton Trans., 2007, 5623 RSC; (b) B. D. Ward, H. Risler, K. Weitershaus, S. Bellemin-Laponnaz, H. Wadepohl and L. H. Gade, Inorg. Chem., 2006, 45, 7777 CrossRef CAS PubMed; (c) S. Ge, A. Meetsma and B. Hessen, Organometallics, 2008, 27, 3131 CrossRef CAS; (d) W. Zhang and L. R. Sita, J. Am. Chem. Soc., 2008, 130, 442 CrossRef CAS PubMed; (e) S. Collins, Coord. Chem. Rev., 2011, 255, 118 CrossRef CAS PubMed; (f) F. T. Edelmann, Chem. Soc. Rev., 2009, 38, 2253 RSC; (g) M. P. Coles, Dalton Trans., 2006, 985 RSC; (h) F. T. Edelmann, Chem. Soc. Rev., 2012, 41, 7657 RSC; (i) C. Alonso-Moreno, A. Antiñolo, F. Carrillo-Hermosilla and A. Oterob, Chem. Soc. Rev., 2014, 43, 3406 RSC; (j) P. J. Bailey and S. Pace, Coord. Chem. Rev., 2001, 214, 91 CrossRef CAS.
  3. S. J. Kim, K. Xu, H. Parala, R. Beranek, M. Bledowski, K. Sliozberg, H. W. Becker, D. Rogalla, D. Barreca, C. Maccato, C. Sada, W. Schuhmann, R. A. Fischer and A. Devi, Chem. Vap. Deposition, 2013, 19, 45 CrossRef CAS PubMed.
  4. T. Chen, C. Xu, T. H. Baum, G. T. Stauf, J. F. Roeder, A. G. DiPasquale and A. L. Rheingold, Chem. Mater., 2010, 22, 27 CrossRef CAS.
  5. V. Krisyuk, L. Aloui, N. Prud'homme, S. Sysoev, F. Senocq, D. Samélor and C. Vahlas, Electrochem. Solid-State Lett., 2011, 14, D26 CrossRef CAS PubMed.
  6. (a) K. Junold, J. A. Baus, C. Burschka, D. Auerhammer and R. Tacke, Chem.–Eur. J., 2012, 18, 16288 CrossRef CAS PubMed; (b) K. Junold, C. Burschka, R. Bertermann and R. Tacke, Dalton Trans., 2010, 9401 RSC; (c) K. Junold, J. A. Baus, C. Burschka, T. Vent-Schmidt, S. Riedel and R. Tacke, Inorg. Chem., 2013, 52, 11593 CrossRef CAS PubMed; (d) F. M. Mück, K. Junold, J. A. Baus, C. Burschka and R. Tacke, Eur. J. Inorg. Chem., 2013, 5821 CrossRef PubMed; (e) K. Junold, C. Burschka, R. Bertermann and R. Tacke, Dalton Trans., 2011, 9844 RSC; (f) K. Junold, J. A. Baus, C. Burschka and R. Tacke, Angew. Chem., Int. Ed., 2012, 51, 7020 CrossRef CAS PubMed.
  7. (a) F. L. Riley, J. Am. Ceram. Soc., 2000, 83, 245 CrossRef CAS PubMed; (b) H. Klemm, J. Am. Ceram. Soc., 2010, 93, 1501 CrossRef CAS PubMed; (c) Y. Q. Chen, X. N. Zhang, Q. Zhao, L. He, C. K. Huang and Z. P. Xie, Chem. Commun., 2011, 47, 6398 RSC.
  8. (a) H. Sirringhaus, Adv. Mater., 2014, 26, 1319 CrossRef CAS PubMed; (b) B. Rech and H. Wagner, Appl. Phys. A, 1999, 69, 155 CrossRef CAS.
  9. (a) N. G. Wright, A. B. Horsfall and K. Vassilevski, Mater. Today, 2008, 11, 17 CrossRef; (b) H. P. Phan, D. V. Dao, P. Tanner, J. S. Han, N. T. Nguyen, S. Dimitrijev, G. Walker, L. Wang and Y. Zhu, J. Mater. Chem. C, 2014, 2, 7176 RSC.
  10. (a) J. K. Rath, Sol. Energy Mater. Sol. Cells, 2003, 76, 43 CrossRef; (b) A. F. B. Braga, S. P. Moreira, P. R. Zampieri, J. M. G. Bacchin and P. R. Mei, Sol. Energy Mater. Sol. Cells, 2002, 82, 418 Search PubMed; (c) S. D. Brotherton, Semicond. Sci. Technol., 1995, 10, 721 CrossRef CAS.
  11. (a) O. Flores, R. K. Bordia, D. Nestler, W. Krenkel and G. Motz, Adv. Eng. Mater., 2014, 1 CrossRef PubMed; (b) A. Badzian, T. Badzian, R. Roy and W. Drawl, Thin Solid Films, 1999, 354, 148 CrossRef CAS; (c) A. Badzian, T. Badzian, W. D. Drawl and R. Roy, Diamond Relat. Mater., 1998, 7, 1519 CrossRef CAS.
  12. (a) I. Golecki, F. Reidingger and J. Marti, Appl. Phys. Lett., 1992, 60, 1703 CrossRef CAS PubMed; (b) A. M. Wrobel, A. Walkiewicz-Pietrzykowska, M. Ahola, I. J. Vayrynen, F. J. Ferrer-Fernandez and A. R. Gonzalez-Elip, Chem. Vap. Deposition, 2009, 15, 39 CrossRef CAS PubMed; (c) A. M. Wrobel, A. Walkiewicz-Pietrzykowska, M. Stasiak, T. Aoki, Y. Hatanaka and J. Szumilewicz, J. Electrochem. Soc., 1998, 145, 1060 CrossRef CAS PubMed; (d) A. M. Wrobel and A. Walkiewicz-Pietrzykowska, Chem. Vap. Deposition, 1998, 4, 133 CAS; (e) N. Fainer, A. Golubenko, Y. M. Rumyantsev and E. Maximovskii, Glass Phys. Chem., 2009, 35, 274 CrossRef CAS; (f) A. M. Wrobel, I. Blaszczyk-Lezak, P. Uznanski and B. Glebocki, Chem. Vap. Deposition, 2010, 16, 211 CrossRef CAS PubMed; (g) R. D. Mundo, M. Ricci, R. d'Agostino, F. Fracassi and F. Palumbo, Plasma Processes Polym., 2007, 4, S21 CrossRef PubMed; (h) G. Suchaneck, V. Norkus and G. Gerlach, Surf. Coat. Technol., 2001, 142, 808 CrossRef; (i) A. M. Wrobel and I. Blaszczyk-Lezak, Chem. Vap. Deposition, 2007, 13, 595 CrossRef CAS PubMed; (j) N. Fainer, A. Golubenko, Y. M. Rumyantsev and E. Maximovskii, Glass Phys. Chem., 2009, 35, 274 CrossRef CAS.
  13. (a) H. Gleskova, S. Wagner, V. Gašparık and P. Kováč, Appl. Surf. Sci., 2001, 175, 12 CrossRef; (b) A. Kshirsagar, P. Nyaupane, D. Bodas, S. Duttagupta and S. Gangal, Appl. Surf. Sci., 2011, 257(11), 5052 CrossRef CAS PubMed; (c) H. Sato, A. Izumi, A. Masuda and H. Matsumura, Thin Solid Films, 2001, 395(1), 280 CrossRef CAS; (d) F. V. Assche, W. Kessels, R. Vangheluwe, W. Mischke, M. Evers and M. van de Sanden, Thin Solid Films, 2005, 484(1), 46 CrossRef PubMed; (e) A. Santoni, J. Lancok, S. Loreti, I. Menicucci, C. Minarini, F. Fabbri and D. Della Sala, J. Cryst. Growth, 2003, 258, 272 CrossRef CAS; (f) J. Song, G. S. Lee and P. K. Ajmera, Thin Solid Films, 1995, 270, 512 CrossRef CAS.
  14. M. Asay, C. Jones and M. Driess, Chem. Rev., 2011, 111, 354 CrossRef CAS PubMed.
  15. (a) P. Chadha, D. J. H. Emslie and H. A. Jenkins, Organometallics, 2014, 33, 1467 CrossRef CAS; (b) I. Giebelhaus, R. Müller, W. Tyrra, I. Pantenburg, T. Fischer and S. Mathur, Inorg. Chim. Acta, 2011, 372, 340 CrossRef CAS PubMed; (c) H. C. Aspinall, J. F. Bickley, J. M. Gaskell, A. C. Jones and G. Labat, Inorg. Chem., 2007, 46, 5852 CrossRef CAS PubMed; (d) R. Anwander, F. C. Munck, T. Priermeier and W. Scherer, Inorg. Chem., 1997, 36, 3545 CrossRef CAS PubMed.
  16. S. T. Barry, Coord. Chem. Rev., 2013, 257, 3192 CrossRef CAS PubMed.
  17. (a) Y. V. Zefirov, Russ. J. Inorg. Chem., 2000, 45, 1552 Search PubMed; (b) S. S. Batsanov, Russ. J. Inorg. Chem., 1991, 36, 1694 Search PubMed; (c) S. S. Batsanov, Inorg. Mater., 2001, 37, 871 CrossRef CAS; (d) P. M. Zorkii and A. A. Stukalin, Crystallogr. Rep., 2005, 50, 522 CrossRef CAS; (e) A. Bondi, J. Phys. Chem., 1964, 68, 441 CrossRef CAS.
  18. U. H. Berlekamp, A. Mix, B. Neumann, H. G. Stammler and P. Jutzi, J. Organomet. Chem., 2003, 667, 167 CrossRef CAS.
  19. B. Gostevskii, K. Adear, A. Sivaramakrishna, G. Silbert, D. Stalke, N. Kocher, I. Kalikhman and D. Kost, Chem. Commun., 2004, 1644 RSC.
  20. (a) J. Wagler, U. Bohme, E. Brendler and G. Roewer, Organometallics, 2005, 24, 1348 CrossRef CAS; (b) T. Cheisson and A. Auffrant, Dalton Trans., 2014, 13399 RSC; (c) B. M. Kraft and W. W. Brennessel, Organometallics, 2014, 33, 158 CrossRef CAS; (d) M. Sohail, R. Panisch, A. Bowden, A. R. Bassindale, P. G. Taylor, A. A. Korlyukov, D. E. Arkhipov, L. Male, S. Callear, S. J. Coles, M. B. Hursthouse, R. W. Harringtond and W. Clegg, Dalton Trans., 2013, 10971 RSC.
  21. (a) M. F. Richardson and R. E. Sievers, Inorg. Chem., 1971, 10, 498 CrossRef CAS; (b) Z. Li, A. Rahtu and R. G. Gordon, J. Electrochem. Soc., 2006, 11, C787 CrossRef PubMed.
  22. E. G. Gillan, S. G. Bott and A. R. Barron, Chem. Mater., 1997, 9, 796 CrossRef CAS.
  23. S. F. Wright, D. Dollimore, J. G. Dunn and K. Alexander, Thermochim. Acta, 2004, 421, 25 CrossRef CAS PubMed.
  24. (a) E. G. Gillan, S. G. Bott and A. R. Barron, Chem. Mater., 1997, 9, 796 CrossRef CAS; (b) T. Chen, C. Xu, T. H. Baum, G. T. Stauf, J. F. Roeder, A. G. DiPasquale and A. L. Rheingold, Chem. Mater., 2010, 22, 271 Search PubMed; (c) W. Hunks, P. S. Chen, T. Chen, M. Stender, G. T. Stauf, L. Maylott, C. Xu and J. F. Roeder, Mater. Res. Soc. Symp. Proc., 2008, 1071, F09 CrossRef.
  25. (a) J. H. Boo, S. B. Lee, K. S. Yu, M. M. Sung and Y. Kim, Surf. Coat. Technol., 2000, 131, 147 CrossRef CAS; (b) M. B. J. Wijesundara, G. Valente, W. R. Ashurst, R. T. Howe, A. P. Pisano, C. Carraro and R. Maboudiana, J. Electrochem. Soc., 2004, 151, C210 CrossRef CAS PubMed.
  26. I. V. Magedov, Y. I. Smushkevich, D. N. Plutitskii and N. N. Suvorov, Zh. Obshch. Khim., 1988, 58, 1934 CAS.

Footnote

Electronic supplementary information (ESI) available: 1H NMR, 13C NMR, 29Si NMR spectra for 1 and 2; STA curves; SEM images; deposition furnace. CCDC 1400899 and 1033427. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c5ra09755j

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.