Hydrolysis and Condensation Kinetics of Dimeric Sol—Gel Species by

May 5, 1989 - High-resolution 29Si NMR spectroscopy was used to study the hydrolysis and condensation kinetics of monomeric and dimeric species in the...
0 downloads 0 Views 1MB Size
Downloaded by UNIV OF MICHIGAN ANN ARBOR on February 18, 2015 | http://pubs.acs.org Publication Date: May 5, 1989 | doi: 10.1021/ba-1990-0224.ch014

14 Hydrolysis and Condensation Kinetics of Dimeric Sol-Gel Species by 29 Si NMR Spectroscopy Daniel H. Doughty, Roger A. Assink, and Bruce D. Kay Sandia National Laboratories, Albuquerque, NM 87185

High-resolution Si NMR spectroscopy was used to study the hydrolysis and condensation kinetics of monomeric and dimeric species in the silicate sol-gel system. Peak assignments for the kinetics experiments were determined by comparing acid-catalyzed reaction solutions prepared with limited amounts of water with the synthetically prepared dimeric, trimeric, and tetrameric species. Si NMR peaks were assigned for 5 of the 10 possible dimeric species. The temporal evolution of hydrolysis and condensation products has been compared with a kinetic model developed in our laboratory, and rate constants have been determined. The results indicate that the water-producingcondensation of dimeric species is approximately 5 times slower than the water-producing condensation of the monomeric species. The alcohol-producing condensation of dimeric species is comparable with that of monomeric species. 29

29

S O L - G E L P R O C E S S I N G has become an active area of research (I, 2), because it provides synthetic routes to novel materials. To achieve this goal, the manipulation and control of solution variables to tailor the chemical or physical properties of the product are necessary. The processing and structural evolution of sol-gel-derived materials are intimately linked with the solution chemistry used in their formation (3). Although many sol-gel processes have appeared in the literature, detailed kinetics studies are usually lacking. We are exploring the kinetics and mechanism of the constitutive reactions in the

O065-2393/9O/O224-O241$O6.0070 © 1990 American Chemical Society

In Silicon-Based Polymer Science; Zeigler, J., et al.; Advances in Chemistry; American Chemical Society: Washington, DC, 1989.

242

SILICON-BASED POLYMER SCIENCE: A COMPREHENSIVE RESOURCE

silicate sol-gel system to provide the foundation necessary to understand the underlying chemistry. At the simplest level, silicate sol-gel chemistry can be described by three reactions: ^Si-OR + H 2 0

Downloaded by UNIV OF MICHIGAN ANN ARBOR on February 18, 2015 | http://pubs.acs.org Publication Date: May 5, 1989 | doi: 10.1021/ba-1990-0224.ch014

^ S i - O H + =Si-OH

^ S i - O H + ROH

(1)

^Si-O-Si^ + H20

(2)

=Si-OH + = S i - O R - ^ > ^ S i - O - S i ^ + ROH

(3)

In these equations, k , fccw, and fcca are the rate constants for hydrolysis, water-producing condensation, and alcohol-producing condensation, respectively. Equation 1 shows the hydrolysis of a Si-OR group to produce a silanol group. A silanol group can undergo condensation polymerization with another silanol (equation 2) or with an alkoxy group (equation 3) to produce a siloxane linkage plus water or alcohol, respectively. In this chapter, the rate constants discussed for these reactions are intended to be phenomenological rate constants and not meant to imply mechanistic details. Proposed mechanisms for ester hydrolysis (4-6) under acidic conditions involve bimolecular nucleophilic substitution reactions that have pentacoordinate intermediates. The proposed mechanism for water-producing condensation (5) under acidic conditions (pH < 2.5-4.5, depending on the silicon substituents) involves rapid, reversible protonation of a silanol, followed by a bimolecular reaction forming the siloxane bond and eliminating a hydronium ion. By using these observed rate constants for hydrolysis and condensation, a detailed formalism has been developed that treats the various reactions at the functional-group level (7-9). The observed rate constants were reported (8) for the acid-catalyzed hydrolysis and condensation of monomeric species prepared from Si(OCH 3 ) 4 (tetramethoxysilane [TMOS]). In this chapter, we report the observed rate constants for condensation of the dimeric species (CH30)3Si-0-Si(OCH3)3. u

Experimental Procedures Preparation of Dimer. Hexamethoxydisiloxane [(CH 3 0) 3 Si] 2 0, termed "dimer" in this chapter, was synthesized by the method of Klemperer et al. (JO) with modifications. By using Schlenk techniques, hexachlorodisiloxane (57 mL, 0.312 mol) was added dropwise (over 50 min) to a solution of freshly distilled methyl orthoformate (trimethoxymethane) (400 mL, 3.74 mol) in heptane (200 mL, dried over Na). The reaction was carried out under argon, and the temperature of the reaction mixture was maintained at 50 °C for 5 h. The mixture was stored overnight at room tem­ perature under argon. Purification by vacuum distillation was carried out. The fraction boiling at 50 °C (43-53 Pa) was collected. Final purification to remove H C l was

In Silicon-Based Polymer Science; Zeigler, J., et al.; Advances in Chemistry; American Chemical Society: Washington, DC, 1989.

14.

DOUGHTY ET AL.

Dimeric Sol-Gel

Species

243

Downloaded by UNIV OF MICHIGAN ANN ARBOR on February 18, 2015 | http://pubs.acs.org Publication Date: May 5, 1989 | doi: 10.1021/ba-1990-0224.ch014

accomplished by passing the straw-colored liquid through neutral alumina and dis­ carding the first cut. The purity of the colorless liquid was determined by 2 9 Si N M R spectroscopy and by gas chromatography-mass spectroscopy (GCMS). Samples were stored in the dark under argon. NMR Spectroscopy. The 2 9 Si N M R spectra were recorded at 39.6 MHz on a Chemagnetics console interfaced to a General Electric 1280 data station and pulse programmer. The silicon-free probe and 20-mm-diameter poly(trifluorochloroethylene) sample tubes were purchased from Cryomagnetics. The 4.7-T widebore magnet was constructed by Nalorac Cryogenics. From 4 pulses (during the early stages of the reaction) to 64 pulses (during the latter stages of the reaction) were accumulated. The pulse repetition interval was 17.7 s. Ή broad-band decoupling was applied only during data acquisition to suppress any residual negative nuclear Overhauser effect (NOE) (JI). However, no difference was observed between spectra obtained with gated decoupling and those with con­ tinuous decoupling, a result indicating negligible NOE. This result is expected, because the observed relaxation time is at least a factor of 10 less than the dipolar relaxation time. No field lock was necessary, as the line widths of accumulated spectra were typically 0.5 Hz before a 0.3-Hz line broadening was applied. The total integrated intensities of all the silicons in the sample were usually reproducible to within ± 3% over the course of the reaction. The 2 9 Si spin-lattice relaxation time, T was 4.2 ± 1.0 s, with a slight tendency for the more highly condensed species to exhibit times toward the lower limit of the uncertainty range. The pulse repetition period of 17.7 s ensured that the magnetization has recovered to 98 ± 2% of its equilibrium value. h

Kinetics Studies. The sol-gel kinetics experiments were performed on a methanolic solution of 1.12 M dimer and 1.57 Χ 10~2 M chromium acetylacetonate [Cr(acac)3], a spin relaxation agent. Previous studies (12, 13) showed that Cr(acac)3 concentration does not affect the product distribution or reaction rate of TMOSderived sol-gel solutions. The solutions were acid catalyzed (1.64 X 10 3 M HC1), and various amounts of water were added. To compensate for the exothermicity caused by dimer hydrolysis when water is added to the dimer-methanol solution, the alcoholic silicate solution was chilled in a thermostatically controlled bath prior to the addition of water. By adjusting the temperature to the appropriate level, the desired reaction temperature (25 ± 1 °C) could be achieved within 60 s of mixing. At this time, the sample was removed from the thermostatically controlled bath and inserted into the spectrometer probe. 29 Si N M R spectra were taken immediately after the sample was mixed and then as a function of time for four water concentrations to give water/alkoxide ratios (H 2 0/[Si-ORj) of 1/12, 1/6, 1/4, and 1/3. Figure 1 shows representative spectra for H 2 0/[Si-OR] = 1/6 taken immediately after mixing (5 min) and at a later time (4 h). The figure shows that hydrolysis is rapid under these conditions and that con­ densation occurs at a much slower rate, as evidenced by inspection of the Q 1 region (14), which shows peaks assigned to the first (chemical shift [δ] = -84 ppm) and second (δ = -82.6 ppm) hydrolysis products of dimer within 5 min of mixing. The 2 9 Si N M R spectra were used for two types of measurements. For the first type, the amount of dimer and the extent to which it had been hydrolyzed were determined very early in the reaction by integrating the Q 1 resonances in the -81to -86-ppm region of the spectra. For the second type of measurement, the extent of condensation as a function of time was determined by comparing the integrals of the Q 1 and Q 2 species.

In Silicon-Based Polymer Science; Zeigler, J., et al.; Advances in Chemistry; American Chemical Society: Washington, DC, 1989.

SILICON-BASED POLYMER SCIENCE: A COMPREHENSIVE RESOURCE

244

Q1 regie>n

Q2 region ,

,

'

V

Downloaded by UNIV OF MICHIGAN ANN ARBOR on February 18, 2015 | http://pubs.acs.org Publication Date: May 5, 1989 | doi: 10.1021/ba-1990-0224.ch014

time = 5 min

time = 4 hr

< •""1 - 7 5

- 8 0

,I

- 8 5

. .

i

J

τ — , — τ — ι — ι — ι — ι — ι — ι — ι — r — ι — 1 - 9 0

9 5

1 - 1 0 0

P P M

Figure 1. Si NMR spectra showing formation of hydrolysis and condensation products. 29

The extent of condensation that occurred during the experiment was simply equal to the number of Q 2 silicons (no Q 3 silicons are observed at early times), because all silicons were initially Q 1 species. No Q° species were observed during the course of these experiments. Cyclic trimers consisting of Q 2 silicon atoms are expected to resonate in the Q 1 region of the spectrum (12). However, because low H 2 0/[Si-OR] ratios were used and only the early stages of the reaction were ex­ amined, no evidence of cyclic trimers was observed. Moreover, cyclic trimers could not form without siloxane bond cleavage and, consequently, would not be expected be formed in these experiments (12). Cyclic tetramers have resonances in the Q 2 region (12) and were observed in these experiments.

Results and Discussion 29 Si NMR Assignments. 2 9 Si NMR spectroscopy has been used ex­ tensively to probe silicate sol-gel reactions (8,10,12-23). Peak assignments for silicon nuclei depend on the next-to-nearest neighbor of the silicon atom being observed. With this approach, the degree of hydrolysis and conden­ sation of a given silicon nucleus can be determined. We have confirmed these assignments for Q 1 , Q 2 , and a Q 3 species by recording the 2 9 Si NMR spectra of the unhydrolyzed, pure compounds of dimer, linear trimer (CH30)8Si302, and a mixture of linear and branched tetramer (CH 3 O) 1 0 Si 4 O 3 (Table I). The values are in excellent agreement with the results of Marsmann et al. (15) and previous work in our laboratory (12). We have used these assignments for the kinetics calculations discussed in the next section.

In Silicon-Based Polymer Science; Zeigler, J., et al.; Advances in Chemistry; American Chemical Society: Washington, DC, 1989.

14.

DOUGHTY ET AL.

Dimeric Sol-Gel Species

245

Table I. NMR Assignments for Unhydrolyzed Silicates Silicon Environment This Work Réf. 15 Q 1 species Si-O-Si -85.81 -85.82 Si-O-Si-O-Si -85.99 -86.00 Si-O-Si-O-Si-O-Si -85.98 -86.00 O-Si Si-O-Si -86.19 -86.22 y

Downloaded by UNIV OF MICHIGAN ANN ARBOR on February 18, 2015 | http://pubs.acs.org Publication Date: May 5, 1989 | doi: 10.1021/ba-1990-0224.ch014

N

O-SÎ

Q 2 species Si-O-Si-O-Si -93.69 -93.64 Si-O-Si-O-Si-O-Si -93.90 -93.86 Q 3 species /)-Si -102.00 Si-O-Si -102.10 N 0-Si NOTE: For clarity, -OR groups are not shown. All data are chemical shifts in parts per million downfield from tetramethylsilane.

As we analyzed spectra taken at very high resolution for these experiments, we realized that additional assignments could be made. Other workers (22) have reported fine structure within a region attributed to one silicon environment (e.g., six peaks were reported for Si(OCH3)2(OH)(OSi), which resonates at approximately -84 ppm). The authors conjectured that the various peaks were due to different degrees of hydrolysis of the adjacent silicon. Because of the conditions of our experiments, notably the lower water concentrations and the use of a spectrometer with higher resolution, we are able to assign some of these peaks. Assignment is aided by symmetrical relationships that exist between adjacent silicon atoms. For example, the integrated intensity of each silicon environment in the dimer (CH 3 0)3Si-0-Si(OH)(OCH 3 )2 must be equal at any given time and must have identical time dependence. A representative spectrum is given in Figure 2. Our assignments of chemical shifts for five of the possible 10 dimeric species are given in Table II.

Analysis of Condensation Kinetics. An earlier publication (8) explained how, under certain limiting conditions, the observed rate constants for water- and alcohol-producing condensation reactions in the TMOS sol-gel reactions can be calculated. The limiting conditions are (1) hydrolysis occurs on a time scale short compared with condensation and (2) the concentration of Si-O-Si bonds formed by condensation reactions is small compared with the initial concentration of the methoxy functional group. If these conditions

In Silicon-Based Polymer Science; Zeigler, J., et al.; Advances in Chemistry; American Chemical Society: Washington, DC, 1989.

In Silicon-Based Polymer Science; Zeigler, J., et al.; Advances in Chemistry; American Chemical Society: Washington, DC, 1989. 2

2

3

3

PPM

'Si(OR) -0-Si(OR)

1—ι—ι—|—ι—ι—ι—ι—|—ι—ι—ι—ι -85 -86 -87

2

29

1

Figure 2. Si NMR spectrum of Q region. Dimer hydrolysis products are readily distinguished; 5 of the 10 possible dimeric species are shown.

1—ι—ι—τ—τ—J—ι—ι—ι—ι—J—ι -82 -83 -84

2

Si (OR)(OH) -0- Si (OR) (OH)

2

^Si (OR)(OH) -0- S M O R ) ^

3

Si (OR) (OH)-0-Si(OR) (OH)

2

Si (OR) (OH)-0- Si (0R)

Downloaded by UNIV OF MICHIGAN ANN ARBOR on February 18, 2015 | http://pubs.acs.org Publication Date: May 5, 1989 | doi: 10.1021/ba-1990-0224.ch014

14.

DOUGHTY ET AL.

247

Dimeric Sol-Gel Species

Table II. Chemical Shifts of Dimeric Species δ (ppm from TMS)

Downloaded by UNIV OF MICHIGAN ANN ARBOR on February 18, 2015 | http://pubs.acs.org Publication Date: May 5, 1989 | doi: 10.1021/ba-1990-0224.ch014

Si Environment

Sî(OCH 3 ) 3 -0-Si(OCH 3 ) 3 Si(OCH 3 ) 3 -0-Si(OCH 3 ) 2 OH Sê(OCH 3 ) 3 -0-Si(OCH 3 )(OH) 2 Si(OCH 3 ) 2 (OH)-0-Si(OCH 3 ) 3 Si(OCH 3 ) 2 (OH)-0-Si(OCH 3 ) 2 OH Sz(OCH3)2 (OH)-0-Si(OCH 3)(0H)2 Si(OCH 3 )(OH) 2 -0-Si(OCH 3 ) 3 Si(OCH3)(OH)2 -0-Si(OCH 3) 2 O H

-85.81 -85.76 -85.72 -83.99 -84.02 -84.04 -82.62 -82.69

NOTE: The table shows the effect of different degrees of hydrolysis of the adjacent Si atom on the chemical shift of the observed Si.

are met, a good approximation of the exact kinetics expression for this system is

^£om>

d[

=

(fccw

" *-> + U S i O C l U

(4)

in which d[(SiO)Si]/dt is the initial rate of formation of siloxane (Si-O-Si) bonds, is the average value of the silanol functional group concentration over the measurement time window, and [SiOCH 3 ] 0 is the initial concentration of methoxy functional groups. Equation 4 suggests that by plotting the initial condensation rate divided by versus , the value of fcca can be determined from the intercept, and the value of the difference between fccw and k can be determined from the slope. C3L

The results of this analysis applied to the condensation of dimer are shown in Figure 3. The observed rate constants for alcohol-producing condensation are approximately the same (k = 0.001 and 0.0007 L/mol-min for monomer and dimer, respectively), whereas the rate constants for waterproducing condensation are significantly different (k = 0.006 and 0.0011 L/mol-min, for monomer and dimer, respectively). The lower value offccwfor dimers can be explained by simple arguments. The steric requirements of the bulky siloxane group should decrease the reaction rate constant of dimers compared with monomers. Inductive effects should decrease the rate as well. This latter conclusion is based on the mechanism proposed by Pohl and Osterholtz (5), which postulates that in acidic solutions (pH < 4.5), an equilibrium concentration of protonated silanol is rapidly established. The protonated silanol reacts with a neutral silicate species in the rate-determining step, and then, a hydronium ion is eliminated. The more electron-withdrawing siloxane group (compared with Si-OR and Si-OH) would decrease the stability of the protonated species and effectively shift the equilibrium toward the unprotonated form. Therefore, the observed reaction rate constant should decrease. ca

cw

In Silicon-Based Polymer Science; Zeigler, J., et al.; Advances in Chemistry; American Chemical Society: Washington, DC, 1989.

Downloaded by UNIV OF MICHIGAN ANN ARBOR on February 18, 2015 | http://pubs.acs.org Publication Date: May 5, 1989 | doi: 10.1021/ba-1990-0224.ch014

248

SILICON-BASED POLYMER SCIENCE: A COMPREHENSIVE RESOURCE

1.0

Avg. [SiOH], mol/l Figure 3. Graph of condensation rate divided by average [SiOH] average [SiOH],

versus

The similarity of feca for monomers and dimers is unexpected. If the reaction mechanism involves a protonated silanol, as in the water-producing condensation, subsequent reaction rate constants must be proportionately larger to give an observed of the same magnitude. However, the likely leaving group from a protonated silanol species is water, not an alcohol. An alternative hypothesis is that a protonated alkoxide moiety (=Si-0(H)R) + is formed in a small steady-state concentration. Subsequently, this protonated species could react with a neutral silanol species and eliminate R - O H . The electron-donating effect of the R group would predominate over the electron-withdrawing effect of the -OSi group on silicon. These effects may explain why the measured values of k for dimers are very similar to those of monomers. ca

Conclusion Further work is needed to clarify the nature of the differences between the water- and alcohol-producing condensations of silicate species. But one conclusion is clear: The two condensation reactions proceed by different reaction pathways (or at least have different transition states) and do not necessarily follow the same reactivity trends.

In Silicon-Based Polymer Science; Zeigler, J., et al.; Advances in Chemistry; American Chemical Society: Washington, DC, 1989.

Downloaded by UNIV OF MICHIGAN ANN ARBOR on February 18, 2015 | http://pubs.acs.org Publication Date: May 5, 1989 | doi: 10.1021/ba-1990-0224.ch014

14.

DOUGHTY ET AL.

Dimeric Sol-Gel Species

249

A second important outcome of these data is a result of the earlier observation (8) that, depending on reaction conditions, either water- or alcohol-producing condensation reactions may be the major pathway for polymerization. If the sol-gel reaction is carried out with small amounts of added water, (i.e., [SiOH] is small and [SiOCH 3 ] is large), the alcoholproducing condensation will predominate. Conversely, at high water con­ densation, [SiOH] is large, and the water-producing condensation is favored. The crossover point for dimeric species will be shifted to higher water con­ centrations, compared with that of TMOS condensation (8). This conclusion has practical consequences for controlling the polymerization rate of silicates in sol-gel processing.

Acknowledgments This work was supported by the U.S. Department of Energy under contract number DE-AC04-76DP00789. We gratefully acknowledge helpful discus­ sions with Jeff Brinker, Walter Klemperer, and Vera Mainz. Walter Klemperer and Vera Mainz (Department of Chemistry, University of Illinois) provided samples of trimer and linear and branched tetramer under Sandia contract 57-3178. We also thank Sheri Martinez for synthesizing hexamethoxyldisiloxane and for other technical assistance. Likewise, the tech­ nical assistance of Duane Schneider is gratefully acknowledged.

References 1. Better Ceramics through Chemistry; Brinker, C. J.; Clark, D. E . ; Ulrich, D. R., Eds.; Elsevier-North Holland: New York, 1984. 2. Better Ceramics through Chemistry II; Brinker, C. J.; Clark, D. E . ; Ulrich, D. R., Eds.; Materials Research Society: Pittsburgh, PA, 1986. 3. Brinker, C. J. J. Non-Cryst. Solids 1988, 100, 31-50. 4. McNeil, K. J.; DiCaprio, J. Α.; Walsh, D. Α.; Pratt, R. F. J. Am. Chem. Soc. 1980, 102, 1859. 5. Pohl, E. R.; Osterholtz, F. D. In Molecular Characterization of Composite Interfaces; Ishida, H . , Kuma, G., Eds.; Plenum: New York, 1985; p 157. 6. Pope, E. J. Α.; Mackenzie, J. D. J. Non-Cryst. Solids 1986, 87, 185. 7. Kay, B. D.; Assink, R. A. In Better Ceramics through Chemistry II; Brinker, C. J.; Clark, D. E.; Ulrich, D. R., Eds.; Materials Research Society: Pittsburgh, PA, 1986;p157. 8. Assink, R. Α.; Kay, B. D. J. Non-Cryst. Solids 1988, 99, 359. 9. Kay, B. D.; Assink, R. A. J. Non-Cryst. Solids 1988, 104, 112. 10. Klemperer, W. G.; Mainz, V. V.; Millar, D. M . In Better Ceramics through Chemistry II; Brinker, C. J.; Clark, D . E . ; Ulrich, D. R., Eds.; Materials Research Society: Pittsburgh, PA, 1986;p3. 11. Horn, H.; Marsmann, H . C. Makromol. Chem. 1972, 162, 255. 12. Balfe, C. Α.; Martinez, S. L. In Better Ceramics through Chemistry II; Brinker, C. J.; Clark, D. E.; Ulrich, D. R., Eds.; Materials Research Society: Pittsburgh, PA, 1986;p27. 13. Kelts, L. W.; Effinger, N. J.; Melpolder, S. M . J. Non-Cryst. Solids 1986, 83, 353. In Silicon-Based Polymer Science; Zeigler, J., et al.; Advances in Chemistry; American Chemical Society: Washington, DC, 1989.

Downloaded by UNIV OF MICHIGAN ANN ARBOR on February 18, 2015 | http://pubs.acs.org Publication Date: May 5, 1989 | doi: 10.1021/ba-1990-0224.ch014

250

SILICON-BASED POLYMER SCIENCE: A COMPREHENSIVE RESOURCE

14. Q n nomenclature is taken from that of Englehardt, G. et al., Ζ. Chem. 1974, 14, 109, where the superscript " n " refers to the number of O-Si groups bonded to the silicon of interest. 15. Marsmann, H . C.; Meyer, E.; Vongehr, M . ; Weber, E. F. Macromol. Chem. 1983, 184, 1817. 16. Orcel, G.; Hench, L. J. Non-Cryst. Solids 1986, 79, 177. 17. Englehardt, G.; Altenburg, W.; Hoebbel, D.; Wieker, W. Z. Anorg. Allg. Chem. 1977, 428, 43. 18. Hoebbel, D.; Garzo, G.; Englehardt, G.; Till, A. Z. Anorg. Allg. Chem. 1979, 450, 5. 19. Harris, R. K.; O'Connor, M . J. J. Magn. Reson. 1984, 57, 115.

20. Harris, R. K.; Knight, C. T. G. J. Chem. Soc. Faraday, Trans. 2 1983, 79, 1525. 21. Harris, R. K.; Knight, C. T. G. J. Chem. Soc. Faraday, Trans. 2 1983, 79, 1539.

22. Artaki, I.; Bradley, M . ; Zerda, T. W.; Jonas, J. J. Phys. Chem. 1985, 89, 4399. 23. Zerda, T. W.; Artaki, I.; Jonas, J. J. Non-Cryst. Solids 1986, 81, 365. RECEIVED for review May 27, 1988. ACCEPTED revised manuscript May 5, 1989.

In Silicon-Based Polymer Science; Zeigler, J., et al.; Advances in Chemistry; American Chemical Society: Washington, DC, 1989.