ebook img

Kinetics of acid-catalyzed aldol condensation reactions PDF

13 Pages·2007·0.28 MB·English
by  
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Kinetics of acid-catalyzed aldol condensation reactions

Author's personal copy ARTICLE IN PRESS AtmosphericEnvironment41(2007)6212–6224 www.elsevier.com/locate/atmosenv Kinetics of acid-catalyzed aldol condensation reactions of aliphatic aldehydes (cid:2) Mia T. Casalea, Aviva R. Richmana, Matthew J. Elroda, , Rebecca M. Garlandb,c,1, Melinda R. Beaverb,c, Margaret A. Tolbertb,c aDepartmentofChemistryandBiochemistry,OberlinCollege,Oberlin,OH44074,USA bDepartmentofChemistryandBiochemistry,UniversityofColorado,Boulder,CO80309,USA cCIRES,UniversityofColorado,Boulder,CO80309,USA Received29September2006;receivedinrevisedform21December2006;accepted3April2007 Abstract Field observations of atmospheric aerosols have established that organic compounds compose a large fraction of the atmospheric aerosol mass. However, the physical/chemical pathway by which organic compounds are incorporated into atmospheric aerosols remains unclear. The potential role of acid-catalyzed reactions of organic compounds on acidic aerosolshas beenexploredasapossible chemical pathwayfortheincorporationoforganic materialintoaerosols. Inthe present study,ultraviolet–visible (UV–vis) spectroscopywas used tomonitor the kineticsofformation ofthe productsof theacid-catalyzedaldolcondensationreactionofarangeofaliphaticaldehydes(C –C ).Theexperimentswerecarriedout 2 8 atvarioussulfuricacidconcentrationsandarangeoftemperaturesinordertoestimatetherateconstantsofsuchreactions on sulfuric acid aerosolsunder troposphericconditions. Therateconstants weregenerally foundtodecrease asthechain lengthofthealiphaticaldehydeincreased(exceptforacetaldehyde,whichhadanunusuallysmallrateconstant),increase asafunctionofsulfuricacidconcentrationaspredictedbyexcessaciditytheory,andshowednormalArrheniusbehavior asafunctionoftemperature.Whilethekineticdataaregenerallyconsistentwithpreviouslaboratoryreportsofaldehyde reactivityinvarioussulfuricacidmedia,thealdolcondensationreactionsinvolvingaliphaticaldehydesdonotappearfast enough to be responsible for significant transfer of organic material into atmospheric aerosols. r 2007 Elsevier Ltd. All rights reserved. Keywords:Troposphere;Organicaerosols;Acid-catalyzedreactions;Kinetics 1. Introduction scattering and absorbing incoming solar radiation, leading to a net cooling at the surface. Indirectly, It is now recognized that tropospheric aerosols aerosols impact climate by serving as cloud con- play an important role in the global climate system. densation nuclei (CCN). The optical and hygro- Tropospheric aerosols directly affect climate by scopic properties, cloud condensation ability and chemical reactivity of aerosols are impacted by heterogeneous reactions, as all of these properties (cid:2) Correspondingauthor.Tel.:+14407756583. depend on composition. Field measurements indi- E-mailaddress:[email protected](M.J.Elrod). 1MaxPlanckInstituteforChemistry,BiogeochemistryDepart- catethatambientaerosolscanbecomposedofupto ment,P.O.Box3060,55020Mainz,Germany. 80% by mass organic compounds (Saxena and 1352-2310/$-seefrontmatterr2007ElsevierLtd.Allrightsreserved. doi:10.1016/j.atmosenv.2007.04.002 Author's personal copy ARTICLE IN PRESS M.T.Casaleetal./AtmosphericEnvironment41(2007)6212–6224 6213 Hildemann,1996;Middlebrooketal.,1998;Molnar elevated, Brock et al. (2003) found that the particle et al., 1999; Turpin et al., 2000). growth exceeded that which could be explained by Organic compounds may be transferred from the SO oxidation alone. However, Zhang and Wexler 2 vapor phase to the aerosol phase by physical (2002) found no significant difference in the organic processes such as dissolution and condensation. compositionofaerosolswhethertheambientsulfate However, it has been difficult to rationalize the high aerosols were acidic or neutralized. organic content of aerosols by invoking only Because of the well-known acid-catalyzed aldol physical mechanisms. For example, it is clear that condensation reaction for aldehydes, there have manyatmosphericallyabundantorganiccompounds been a number of laboratory studies in which the are not sufficiently soluble to partition significantly incorporation of gaseous aldehydes into acidic into aerosols. Recently, chemical mechanisms have solutions as compared to non-acidic solutions have been explored as a means by which the transfer beeninvestigated(IraciandTolbert,1997;Jangand of organic compounds from the gas phase to the Kamens, 2001; Jang et al., 2002; M. Jang, et al., aerosol phase may be enhanced. In particular, 2003; M.S. Jang, et al., 2003; Michelsen et al., 2004; acid-catalyzed reactions of organic compounds on Liggio et al., 2005; Zhao et al., 2005; Surratt et al., tropospheric aerosols have been investigated in 2006). Similar studies have been performed for this context, as these reactions are postulated ketones (M. Jang, et al., 2003; M.S. Jang, et al., to lead to enhanced uptake of organic compounds 2003; Esteve and Noziere, 2005), acrolein and a- by aerosols (Jang et al., 2002; Zhang and Wexler, pinene (Czoschke et al., 2003; Iinuma et al., 2004; 2002). Tolocka et al., 2004) and isoprene (Limbeck et al., Thereareseveraltypesofacid-catalyzedreactions 2003). Many of these studies have concluded that of carbonyl-containing organic compounds that are the occurrence of accretion reactions are necessary well known in bulk solutions (Carey and Sundberg, toexplaintheuptake oforganicmaterialinto acidic 1990;VollhardtandSchore,1994).Forexample,the media (Jang and Kamens, 2001; Jang et al., 2002; acid-catalyzed aldol condensation reaction of alde- Czoschke et al., 2003; M. Jang, et al., 2003; M.S. hydes results in the conversion of two aldehyde Jang, et al., 2003; Kalberer et al., 2004; Tolocka et molecules into a single molecule that has a carbon al., 2004; Jang et al., 2005). However, the actual backbone made up of the sum of the carbon atoms identification of specific molecular products (which in the reactant aldehydes. Barsanti and Pankow are needed to prove which accretion reactions are (2004) have termed any such carbon backbone occurring) has been difficult. Garland et al. (2006) building processes (of which the aldol condensation recently used a suite of analytical approaches reaction is but one) as accretion reactions, since the (aerosol mass spectrometry, FTIR and UV–vis end result is the formation of larger molecules from spectroscopy, 1H NMR and GC/MS) to directly smaller ones. The net result of these types of show that hexanal undergoes aldol condensation reactions is the formation of relatively less volatile reactions on sulfuric acid aerosols under laboratory products (which helps prevent the physical transfer conditions. However, high hexanal concentrations oforganicmaterialbackintothegasphase),andthe were used in the study, and no kinetics data were dissolution of more reactant to maintain the availabletoestimatethelikelihoodofsuchreactions solubility equilibrium (the ‘‘enhanced’’ solubility on tropospheric sulfuric acid aerosols. There have effect). also been several other experiments in which the SomefieldstudieshavenotedthatwhenbothSO heterogenous interaction of gaseous aldehydes with 2 and volatile organic compound (VOC) concentra- sulfuric acid surfaces has been probed. Michelsen tions are high, there appears to be additional et al. (2004) studied the interaction of gaseous particle growth, perhaps indicating that acid-cata- acetaldehyde with a film of sulfuric acid using a lyzed reactions are causing the accretion of organic Knudsen cell apparatus as a function of H SO 2 4 compoundsintoaerosols(Brock etal.,2003;Chu et composition and temperature. They found evidence al., 2004). Brock et al. (2002) found that in power for acetaldehyde reaction in the observation of the plant plumes where SO concentrations are high apparent enhanced solubility of acetaldehyde under 2 and the VOCs concentrations are low, the observed certain conditions. On the other hand, Zhao et al. particle growth can be accounted for by SO (2005) found only reversible uptake of octanal on 2 oxidation alone. However, in a different air parcel sulfuric acid films, which suggests that aldol where both SO and VOC concentrations were condensation reactions for octanal are possibly 2 Author's personal copy ARTICLE IN PRESS 6214 M.T.Casaleetal./AtmosphericEnvironment41(2007)6212–6224 quite slow. Noziere and Esteve (2005,2006) studied irreversibly and an a,b-unsaturated carbonyl com- the UV–vis spectra of H SO solutions exposed to pound is formed. 2 4 several different carbonyl-containing compounds Baigrie et al. (1985) carried out the first extensive (including the aldehydes acetaldehyde, propanal kinetics study of the acid-catalyzed aldol reaction and butanal), and concluded that several sequential for the self-reaction of any aldehyde in their bulk aldol condensation reactions were occurring. Acet- solution study of the acetaldehyde aldol reaction as one has also been found to undergo aldol con- a function of sulfuric acid concentration at room densation on highly acidic H SO surfaces to form temperature. They found that crotonaldehyde 2 4 mesityl oxide, and ultimately, 1,3,5-trimethyl ben- (the a,b-unsaturated carbonyl compound formed zene (Duncan et al., 1999). in reaction 4) became the dominant product of In chemical synthesis situations, aldol condensa- the reaction for H SO compositions equal to 2 4 tion reactions are widely utilized to build molecules 60wt% and higher. Under all conditions studied, with larger carbon backbones. However, base the kinetics were observed to be second order in catalysis is the preferred synthetic route for a acetaldehyde, confirming that reaction 3 is the variety of reasons (Carey and Sundberg, 1990; rate-limiting step in the mechanism. The rate Vollhardt and Schore, 1994), and thus the mechan- constants were found to increase as a function of ism and kinetics of the acid-catalyzed aldol con- acidity, and because of the highly non-ideal nature densation reaction has received considerably less ofthesolutions,thekineticsresultswereinterpreted study. However, it has been established that the with the aid of excess acidity theory. In a further mechanism given in Fig. 1 (shown for acetaldehyde, study of the acetaldehyde aldol self-reaction, but easily generalized for any aldehyde) is operative Esteve and Noziere (2005) used a rotating wetted- in the acid-catalyzed aldol condensation of alde- wall reactor to estimate the bulk rate constant for hydes (Noyce and Snyder, 1959). Reactions 1 and 2 acid-catalyzed reaction of acetaldehyde for two are fast, while reaction 3 is slow and known to be different sulfuric acid compositions. Noziere the rate-limiting step in the process. Under high and Esteve (2005); Noziere et al. (2006) also studied acidity conditions, reaction 4 occurs quickly and the kinetics of formation of subsequent products (larger than crotonaldehyde) in the acetaldehyde system as a function of acid strength (75–96wt% H SO ) and temperature (273–314K). Kinetics 2 4 measurements have also been carried out for the acid-catalyzed reactions of the following ketones: acetone (Duncan et al., 1999; Esteve and Noziere 2005), 2-butanone, 2,4-pentanedione (Esteve and Noziere, 2005), and methyl vinyl ketone (Noziere et al., 2006). As discussed above, the kinetics literature for the acid-catalyzed reactions of the aldehydes is primar- ily focused on the studies of the acetaldehyde system. In particular, studies concerning the effect of the molecularity of the aldehydes on the rate constant and the acid and temperature dependence of the rate constant for aldehydes other than acetaldehyde are lacking. In the present study, UV–vis spectroscopy was used to monitor the kinetics of formation of the products in the bulk phase acid-catalyzed aldol condensation reaction of a range of linear aliphatic aldehydes (C –C ). 2 8 Theexperimentswerecarriedoutatvarioussulfuric acid compositions and a range of temperatures in order to estimate the rate constants of such reactions on sulfuric acid aerosols under tropo- Fig. 1. Acid-catalyzed aldol condensation mechanism for acet- aldehyde. spheric conditions. Author's personal copy ARTICLE IN PRESS M.T.Casaleetal./AtmosphericEnvironment41(2007)6212–6224 6215 2. Experimental transient, locally high concentrations of aldehyde) to the sulfuric acid solution and vigorously mixing In concentrated H SO (and in H O) the strong for several minutes to ensure solution homogeneity. 2 4 2 p-p* (e(cid:2)10,000cm(cid:3)1M(cid:3)1) transition arising from For example, a typical acetaldehyde experiment the C ¼ O chromophore for aliphatic aldehydes involved adding 100ml of a 1.5M acetaldehyde occurs at peak wavelengths less than 200nm, while solution to 50ml of 75wt% sulfuric acid, resulting the very weak n-p* (e(cid:2)15cm(cid:3)1M(cid:3)1) transition in a 3(cid:4)10(cid:3)3M acetaldehyde in 75wt% H SO 2 4 occurs around 270nm (Lambert et al., 1987). solution. A small portion of the reaction mixture Therefore, the reactant aldehydes are not easily was then added to either a 1mm or 1cm pathlength monitored using standard UV–vis spectroscopy. quartz cuvette, and the reaction system was However, the dehydrated aldol condensation pro- continuously monitored in the cuvette by UV–vis ducts for the reactions of aliphatic aldehydes spectroscopy. As mentioned above typical initial are a,b-unsaturated aldehydes (the product of aldehyde concentrations were on the order of reaction 4 in Fig. 1) which have strong p-p* 10(cid:3)3M, which provided sufficient UV–vis absor- (e(cid:2)10,000cm(cid:3)1M(cid:3)1) transitions at wavelengths of bance values and convenient initial rates. For the at least 245nm in concentrated sulfuric acid. roomtemperatureexperiments,thetemperaturewas According to the Woodward rules (Lambert et al., 29571K. For the temperature dependence experi- 1987), the peak wavelengths for hydrated aldol ments, the 50ml sulfuric acid solution was either condensation products (the reactant in reaction 4, precooled or preheated to the appropriate tempera- Fig. 1) are not expected to shift significantly from ture before the addition of the aldehyde solution, the parent aldehyde wavelengths. Therefore, the and the cuvette was maintained at the desired experiments described here specifically concern the temperature by the automated temperature control kinetics of the formation of the dehydrated aldol system of the UV–vis spectrometer. The lowest condensation product (as opposed to the overall temperature achieved ((cid:3)241C) was dictated by the loss of the parent aldehyde or the production of limitedcapabilitiesoftherefrigerationsystemofthe hydrated products). For this reason, it was neces- temperature control system. sary to perform experiments at sulfuric acid concentrations of 60wt% and higher to ensure 3. Results and discussion complete dehydration of all aldol reaction products (Baigrie et al., 1985). 3.1. Identification of reaction products The Beer’s law molar extinction coefficients—to allowforconversionoftheUV–visabsorbancedata The aldol condensation reaction products of into absolute production concentrations—were di- acetaldehyde (crotonaldehyde) and butanal (2-ethyl- rectly determined in 75wt% H SO for crotonalde- 2-hexenal)arecommerciallyavailable.Thespectraof 2 4 hyde (the acetaldehyde aldol condensation product) thesetwo products wereinvestigatedinconcentrated and 2-ethyl-2-hexenal (the butanal aldol condensa- sulfuric acid solutions in order to identify their tion product) by preparing dilute solutions in characteristic peak wavelengths and to determine 75wt% H SO and measuring the absorbance at their Beer’s law molar extinction coefficients. In 2 4 the appropriate peak wavelength. The spectra were 75wt% H SO , crotonaldehyde has a peak absor- 2 4 collected quickly, to avoid any possible acid- bance at 245nm and e¼ 9280cm(cid:3)1M(cid:3)1. Similarly, catalyzed reaction. in 75wt% H SO , 2-ethyl-2-hexenal has a peak 2 4 Aqueous solutions of the aldehydes were pre- absorbance at 266nm and e¼ 11960cm(cid:3)1M(cid:3)1. pared by adding pure aldehyde to deionized water While the peak wavelengths shifted as a function of and vigorously mixing for several hours to ensure acid strength, the Beer’s law molar extinction complete dissolution. For the larger aldehydes, the coefficients varied by less than 10% over the acid concentrations of these aqueous solutions were and temperature range of the experiments. determined by their water solubility. The sulfuric Fig. 2 shows several spectra collected as a function acid solutions were prepared by dilution of con- oftimeforanacetaldehydereactionsystemin75wt% centrated (96wt%) H SO with deionized water. H SO at 295K at a relatively high initial acetalde- 2 4 2 4 The kinetics experiments were typically performed hyde concentration (6.8(cid:4)10(cid:3)3M) carried out in a by adding the aqueous solution of the aldehyde 1mm cuvette. At early times, it is readily apparent (diluted aldehyde solutions were used to avoid that crotonaldehyde (with a peak wavelength of Author's personal copy ARTICLE IN PRESS 6216 M.T.Casaleetal./AtmosphericEnvironment41(2007)6212–6224 creasedalkylchainlengthwillcausearedshiftfrom crotonaldehyde (the acetaldehyde aldol product) to 1.2 hr 2.6 2-methyl-2-pentenal (the propanal aldol product). 2.4 However, the additional lengthening of the alkyl 2.2 chain in the parent aldehydes will not lead to an 2.0 additional red shift in the dehydrated aldol con- 1.8 densation products (Lambert et al., 1987), as the ce 1.6 0.8 hr nature of the alkyl group has a much smaller effect ban 1.4 onthepeakwavelength(becauseitisnotpartofthe or 1.2 s chromophore electronic system) than does the size Ab 1.0 0.4 hr of the conjugated system (which is part of the 0.8 19 hr chromophore electronic system). 0.6 16 hr 0.4 10 hr 0.2 3.2. Integrated rate law analysis and reaction 0.0 efficiency 200 250 300 350 400 450 500 Wavelength (nm) Data were collected at the peak absorption Fig.2. UV–visabsorptionspectraatdifferentreactionstimesfor wavelength for each system as a function of time, the acetaldehyde reaction system at [acetaldehyde] ¼ 0 and the absorption values were converted to 6.8(cid:4)10(cid:3)3M,75wt%H SO and295K. 2 4 absolute aldol product concentration via Beer’s Law. Fig. 3 shows such a plot for an acetaldehyde 245nm) is a major product of the reaction. At later (3(cid:4)10(cid:3)3M) kinetics run in 75wt% H SO and times, the crotonaldehyde peak goes off scale, but a 2 4 295K carried out in a 1mm cuvette. At the lower newpeakat364nmisobservedtogrowin.According initial acetaldehyde concentration used in this run to the Woodward rules for a,b-unsaturated carbonyl (as compared to the data shown in Fig. 2), the compounds (Lambert et al., 1987) the later time 245nmabsorptionpeakstaysonscale(A ¼ 1.2), product is likely to have either three additional units max and the crotonaldehyde concentration can be of conjugation in an aliphatic structure, or two calculated according to Beer’s law at all times. As additional units of conjugation as part of a ring mentioned above, acetaldehyde does have a weak structure. The latter possibility has been suggested as n-p* transition in this range that could potentially a possible product of crotonaldehyde self reaction via complicate the conversion of absorbance data to a Michael addition mechanism (McIntosh et al., crotonaldehyde concentration. However, using the 1980). In our work to determine Beer’s law molar initial acetaldehyde concentration of 3(cid:4)10(cid:3)3 and extinction coefficients for crotonaldehyde in H SO 2 4 solutions, we noted that these solutions eventually showed small absorptions at 364nm. Noziere and 0.0014 Esteve have also noted this absorption in their study 0.0012 ofacetaldehydereactionsinsulfuricacid(Noziereand Esteve,2005;Noziereetal.,2006),buthaveattributed 0.0010 M) it to additional aldol condensation reactions of e] ( 0.0008 d products with acetaldehyde. A detailed study of this y h e crotonaldehydeself-reactioniscurrentlyunderwayin ald 0.0006 n our labs. oto 0.0004 Cr All other aldehyde reaction systems in this study [ 0.0002 showed similar product behavior. The peak wave- length for the aldol products for all other systems 0.0000 was 266nm (the Beer’s law molar extinction -0.0002 coefficient was assumed to be the same as for 0 50000 100000 150000 200000 250000 2-ethyl-2-hexenal for the purposes of the kinetics Time (s) analysis) and a later time product was also always Fig. 3. Experimental data (circles) and second-order kinetics fit observed at 364nm in 75wt% H SO . Again, the 2 4 (line) of the aldol condensation reaction of acetaldehyde at Woodward rules accurately predict that the in- [acetaldehyde] ¼3(cid:4)10(cid:3)3M,75wt%H SO and295K. 0 2 4 Author's personal copy ARTICLE IN PRESS M.T.Casaleetal./AtmosphericEnvironment41(2007)6212–6224 6217 Beer’s law molar extinction coefficient for this Inthiscase,theinitialrateofthealdolproduction transition of e(cid:2)15cm(cid:3)1M(cid:3)1, the maximum absor- is measured as a function of the initial aldehyde bance due to acetaldehyde under these conditions is concentration. Fig. 4 shows a set of initial rates only 0.004, which is insignificant compared to the experiments for the butanal system at 75wt% absorption values (A ¼ 1.2) observed in the H SO and 295K. To linearize the determination max 2 4 experiments. of the rate order n, logarithms are taken of both AswillbeconfirmedinSection3.3,thereactionis sides of Eq. (3): second order in the reactant aldehyde: (cid:2) (cid:3) d½aldol(cid:5) 1d½aldehyde(cid:5) d½aldol(cid:5) log ¼ nlog½aldehyde(cid:5)þlog k. (4) (cid:3) ¼ ¼ k½aldehyde(cid:5)2. (1) dt 2 dt dt The kinetics curve can be directly fit by the Fig.5showstherateorderanalysisforthedatain integrated second-order product rate law arising Fig. 4. From these experiments, a rate order of 1.95 from Eq. (1) above (the solid line in Fig. 3): was determined, thus confirming the expected rate (cid:3)1 ½aldehyde(cid:5) order of 2 as given in Eq. (1). Rate constants were ½aldol(cid:5) ¼ þ 0. (2) t ð2=½aldehyde(cid:5) Þþ4kt 2 calculated via Eq. (3) from measurements of the 0 initial rate and [aldehyde] . Typically, at least three 0 ThefitisshownasthesolidlineinFig.3andisin kinetics runs (with different initial aldehyde con- excellent agreement with the data, resulting in a centrations to confirm the rate order of the process) bimolecular rate constant of 1.2(cid:4)10(cid:3)3M(cid:3)1s(cid:3)1. were carried out for each combination of parent However, this method for determining rate con- aldehyde, H SO concentration and temperature. 2 4 stants was extremely time-consuming (data were The statistical error (1s) for the entire data set of collected for 3 days to construct Fig. 3). Section 3.3 rate constants was 14%. The sources of systematic describesanalternativeinitialratesmethodthatwas error in the determination of the rate constants are used to more efficiently determine rate constants. expected to arise from linearly dependent uncer- However, the data in Fig. 3 provide unique tainties in Beer’s law molar extinction coefficients information on the reaction. As the reaction has (10%, as discussed above) used to calculate the been followed virtually to completion in Fig. 3, it is initial rates, and in the initial concentration of the possible to estimate the final concentration of the aldehydes (10%, as estimated from the volumetric aldol product crotonaldehyde. Therefore, using the precision used in the preparation of the solutions) stoichiometry of the reaction and the initial con- used to calculate the bimolecular rate constants centration of acetaldehyde, it is possible to deter- themselves. mine the aldol yield of the reaction. The crotonaldehyde concentration at the longest time is about 0.0013M, and the initial acetaldehyde con- 0.00012 centration was 0.003M. According to the stoichio- metry of the reaction (two acetaldehyde molecules 0.00010 react for every crotonaldehyde molecule formed), a M) rceraoctotinoanldyeiheyldde.ofTh1e0r0e%fores,hothueldolbesaedrvetod 0cr.0o0to15naMl- nal] ( 0.00008 e dehyde yield of the reaction is greater than 85%. ex 0.00006 h 2- yl- 3.3. Initial rates and aldehyde rate order analysis eth 0.00004 2- [ In order to expedite the determination of rate 0.00002 constants, an initial rates method was used. In the initial rates method, the kinetics of the reaction are 0.00000 0 200 400 600 800 1000 1200 monitored before the reaction has proceeded Time (s) significantly such that the following approximation holds: Fig. 4. Initial rates measurements for the aldol condensation reaction of butanal at 75wt% H SO and 295K. Circles: d½aldol(cid:5) 2 4 ¼ k½aldehyde(cid:5)n ¼ k½aldehyde(cid:5)n. (3) [butanal]0¼1.0(cid:4)10(cid:3)3M, squares: [butanal]0¼5.0(cid:4)10(cid:3)4M, dt 0 triangles:[butanal] ¼2.5(cid:4)10(cid:3)4M. 0 Author's personal copy ARTICLE IN PRESS 6218 M.T.Casaleetal./AtmosphericEnvironment41(2007)6212–6224 the aldehyde rate order but that is instead applied -7.0 here to find the effective H+ rate order (the slope, m, in Eq. (5)) for the acid-catalyzed process. The -7.2 intercept, b, in Eq. (5) represents the value of log k ate) -7.4 at infinite dilution (X ¼ 0). Values for CH+ and X nitial r -7.6 (fCoroxvaarnioduYs aHte2sS,O1497s8o)l,utainodnsahpaavreambeeetenrizceodmfpoilremd og (i -7.8 forawhasbeendeterminedforH2SO4solutions(Shi l et al., 2001). For example, in the previous study of theacetaldehydesystem,whenthedatawereplotted -8.0 according to Eq. (5), a slope of 0.95 was obtained, -8.2 affirming the expectation that the acid catalysis is first order in [H+] for the aldol condensation -3.6 -3.5 -3.4 -3.3 -3.2 -3.1 -3.0 -2.9 mechanism (Baigrie et al., 1985). log [butanal] 0 Fig.6showstheexcessacidityplot(carriedoutas describedabove)fortheaciddependenceoftherate Fig. 5. Rate order determination for the aldol condensation reactionofbutanalfordatashowninFig.4. constant for aldol condensation for the butanal system at 295K, from 62.5–88wt% H SO . The 2 4 3.4. Excess acidity analysis data are obviously fit well by the excess acidity formalism, and a slope of 0.99770.015 was The acid dependence of the aldol formation rate determined, again confirming that the acid catalysis constants for the acetaldehyde system has been is first order in [H+]. Excess acidity analyses were previously investigated and analyzed using the carried out for the 295K kinetic data for the other excess acidity formalism (Baigrie et al., 1985). The aldehydesystems(foracetaldehyde,thepresentdata excess acidity method accounts for the non-ideal were combined with the data from the previous behavior of highly acidic solutions and is necessary investigation (Baigrie et al., 1985)) and the results for the interpretation of acid-catalyzed kinetics are given in Table 1. The self-consistency of the results under such conditions. The method has been analysis method is apparent in that all systems are described in detail previously (Cox and Yates, characterized by a slope of unity (within the 95% 1979), and is related to the Hammett acidity confidence interval). As will be subsequently dis- function, H (Hammett and Deyrup, 1932). The cussed, the parameters from the excess acidity 0 excess acidity formalism is an attempt to define a analysis can be used to predict rate constants for linear relationship between the observed rate con- stant and some acidity function for an acid- catalyzed process, much like a pseudo first-order rate constant can be linearly related to the 2 concentration of the excess reactant. The excess acidity parameter, X, is best thought of as a 1 quantity analogous to pH for concentrated acidic solutions. In their previous study of the acid- 0 catalyzed aldol condensation reaction of acetalde- hyde, Baigrie et al. found a linear relationship between (logk(cid:3)logC (cid:3)loga ) and X, where k is -1 H+ w the bimolecular rate constant, C is the proton H+ concentration, and aw is the activity parameter for -2 H O,suchthatthedatacouldbefittothefollowing 2 expression: -3 3 4 5 6 7 8 ðlog k(cid:3)log C (cid:3)log a Þ ¼ mX þb. (5) Hþ w As X is analogous to log[H+], this analysis is Fig.6. Excessacidityplotforthealdolcondensationreactionof essentially a more sophisticated version of the butanal at 295K. Compositions ranging from 62.5 to 88wt% ‘‘log–log’’ plotting exercise described above to find H SO arerepresentedonthisplot. 2 4 Author's personal copy ARTICLE IN PRESS M.T.Casaleetal./AtmosphericEnvironment41(2007)6212–6224 6219 Table1 0.18 60 wt% Excessacidityparametersat295K 0.16 65 wt% 70 wt% Aldehyde m7s(fromEq.(5)) b7s(fromEq.(5)) 0.14 75 wt% 80 wt% Acetaldehyde 1.00370.049 (cid:3)7.3170.27 0.12 85 wt% Propanal 0.95070.031 (cid:3)5.4770.17 1) Butanal 0.99770.015 (cid:3)5.69170.084 1- s 0.10 PHeenxtaannaall 00..9967167700..002302 (cid:3)(cid:3)56..93307700..1118 -k (M 0.08 0.06 Heptanal 0.9970.15 (cid:3)6.3170.71 Octanal 1.00470.089 (cid:3)6.7270.42 0.04 0.02 0.00 the lower H SO concentrations that were not 2 4 1 2 3 4 5 6 7 8 9 accessible in this study (due to the requirement that Number of Carbons Atoms in Aldehyde thedehydratedaldolproductbethedominantform) but that are of interest in assessing the role of aldol Fig. 7. Dependence of the aldol condensation reaction rate condensation reactions in tropospheric aerosols. constant on the reactant aldehyde and H SO concentration at 2 4 Because previous kinetic work (Noyce and Snyder, 295K. 1959) suggests that the dehydration step (Eq. (4) in Fig. 1) is not rate-limiting, but rather that the rate limiting step is the aldol formation (Eq. (3) in it is often used in atmospheric studies as the Fig. 1), it should be possible to extrapolate the ‘‘prototype’’ aldehyde. From the point of view of present data to lower H SO concentrations, where aldol reactivity, acetaldehyde is clearly not a repre- 2 4 the dehydrated aldol product is not formed. sentative aldehyde. 3.5. Aldehyde dependence 3.6. Temperature dependence Thisstudyisthefirstsystematicexaminationofthe Temperature dependence measurements were dependence of the aldol reaction rate constant for a made for the acetaldehyde, butanal, and hexanal numberofdifferentaldehydes.Fig.7showsthe295K systems at 75wt% H SO to assess whether the 2 4 rate constants for each aldehyde system at a number temperature dependence of the rate constant varies of different H SO compositions. As discussed in among the different aldehydes studied. Similar 2 4 Section3.4,theaciddependenceoftherateconstants measurements were made for the butanal system can be precisely analyzed through the excess acidity at 65 and 85wt% H SO to assess whether the 2 4 formalism. However, Fig. 7 shows the interesting temperature dependence of the rate constant varies dependence of the rate constant on the alkyl chain among the different H SO compositions used. 2 4 length of the parent aldehyde. In particular, the Fig. 8 shows an Arrhenius analysis of the results 75wt% H SO 295K rate constant dramatically of the 75wt% H SO experiments. It is clear that 2 4 2 4 increases from acetaldehyde (1.61(cid:4)10(cid:3)3M(cid:3)1s(cid:3)1) normal Arrhenius behavior is observed over the to butanal (84.9(cid:4)10(cid:3)3M(cid:3)1s(cid:3)1), and then gradually range of temperatures studied ((cid:3)24 to 421C). decreasesasthealkylchainlengthensfurther.Therise Table 2 contains the Arrhenius parameters deter- in the reactivity from the C to the C aldehyde is mined for all temperature dependence experiments. 2 4 probably related to the fact that the b-carbon in Withinthe95%confidenceinterval,nodifferencein acetaldehyde is a primary carbon atom (–CH ), the temperaturedependence of the rate constant for 3 whereas the b-carbon for the other systems is a thedifferentH SO compositionswasobtained,but 2 4 secondary carbonatom (–CH –). Presumably,oneof a statistically meaningful decrease in the activation 2 the intermediates in the aldol condensation mechan- energy was determined as the alkyl chain of the ism is destabilized if the reaction occurs at a primary parent aldehyde lengthened. Assuming Arrhenius b-carbon.ThedecreaseinreactivityfromtheC toC behavior holds, the parameterization of the tem- 4 8 aldehydeisnotaseasilyrationalized.Inanycase,itis perature dependence of the rate constant deter- important that the present study indicates that minedhereallowsthepredictionofrateconstantsat acetaldehyde is an unusually unreactive aldehyde, as temperatures below the lowest directly used in the Author's personal copy ARTICLE IN PRESS 6220 M.T.Casaleetal./AtmosphericEnvironment41(2007)6212–6224 -1 -2 -3 -4 -5 k n l -6 -7 -8 -9 -10 0.0031 0.0032 0.0033 0.0034 0.0035 0.0036 0.0037 0.0038 0.0039 1/T (1/K) Fig.8. Arrheniusanalysisofthetemperaturedependenceofthealdolcondensationreactionrateconstant(circles:acetaldehyde;squares: hexanal;triangles:butanal)at75wt%H SO . 2 4 Table2 X ¼ 1:07(cid:4)10(cid:3)3(cid:4)ðwt%H SO Þ2(cid:3)1:34(cid:4)10(cid:3)2 2 4 Arrheniusparameters (cid:4)ðwt%H SO Þþ3:34(cid:4)10(cid:3)1. ð7Þ 2 4 Aldehyde wt%H SO (cid:3)E /R7s(K) lnA7s 2 4 a The 298K water activity (a ) has been previously w Acetaldehyde 75 (cid:3)68807200 16.6870.69 parameterized as a function of water mole fraction Butanal 65 (cid:3)55807650 14.572.2 (Y ) (Shi et al., 2001): 75 (cid:3)62507270 18.4170.93 w 85 (cid:3)64707940 19.773.2 ðwt%H SO Þ Hexanal 75 (cid:3)51307250 13.0770.86 Yw ¼ hðwt%H SO Þþ½100(cid:3)2ðwt%4H2SO4Þ(cid:5)(cid:6)MH2SO4i, (8) 2 4 MH2O present experiments ((cid:3)241C); such temperatures where MH2SO4 and MH2O are the molar mass of are often relevant in the upper troposphere. sulfuric acid and water, respectively. (cid:4) log a ¼ log exp½0:00305(cid:4)ð(cid:3)69:775(cid:4)Y w w 3.7. Parameterization of kinetic data (cid:3)18253(cid:4)Y2 þ31072:2(cid:4)Y3 w w In order to improve the utility of the present (cid:3)25668:8(cid:4)Y4(cid:5)(cid:6)(cid:7). ð9Þ w kinetics data, several parameterizations were deter- From the excess acidity analysis for each alde- mined to allow the calculation of aldol condensation hyde (the slope, m, and intercept, b, values given in rateconstantsforanyaldehyde,H SO composition, 2 4 Table 1 obtained by analysis according to Eq. (5)) and temperature. The excess acidity parameters, log and the values of X, logC and loga (para- C and X were taken from Cox and Yates (1978) H+ w H+ meterized above), the rate constant for each aldol and wereseparately fitto a second-orderpolynomial condensationreactionat295Kforany wt%H SO expression as a function ofwt% H SO : 2 4 2 4 can be calculated: log C ¼ (cid:3)2:62(cid:4)10(cid:3)4 (cid:4)ðwt%H SO Þ2 Hþ 2 4 kð295KÞ ¼ 10ðmXþbþlog CHþþlog awÞ. (10) þ3:79(cid:4)10(cid:3)2 (cid:4)ðwt%H SO Þ 2 4 Rate constants at other temperatures can be (cid:3)2:58(cid:4)10(cid:3)1, ð6Þ calculatedbyusingthe295Krateconstantcalculated Author's personal copy ARTICLE IN PRESS M.T.Casaleetal./AtmosphericEnvironment41(2007)6212–6224 6221 above(anychangesintheapparentrateconstantdue the finding of Garland et al. (2006) that H SO 2 4 to the temperature dependence of C and a are aerosols exposed to hexanal under such conditions H+ w subsumed in the Arrhenius analysis), and the (cid:3)E /R are composed of as much as 88% organic by mass. a values (given in Table 2) from the appropriate JangandKamens(2005)studiedtheheterogeneous Arrhenius analysis: interaction of octanal vapor with various seed aerosols at 298K with the H SO composition (cid:8)(cid:3)E (cid:2)1 1 (cid:3)(cid:9) 2 4 kðTÞ¼ kð295KÞ(cid:4)exp a (cid:4) (cid:3) . controlled by the relative humidity of the smog R T 295K chamber. Using octanal concentrations of about (11) 260ppb, the aerosols were observed to increase in volume by as much as factor of 2 over the 1h 3.8. Comparison to previous work timescale of the experiment. Zhao et al. (2005) have measured the Henry’s law constant for octanal in Thepresentfindingthatdehydratedaldolproducts 67wt% H SO at 296K and determined a value of 2 4 are formed efficiently for a variety of saturated 2300Matm(cid:3)1. Assuming a H SO composition of 2 4 aliphatic aldehydes in concentrated H SO solutions 67wt% for the highest acidity experiments by Jang 2 4 is in agreement with earlier studies on acetaldehyde and Kamens, an octanal aerosol concentration byBaigrieetal.(1985)andNoziereandEsteve(2005) of 6.0(cid:4)10(cid:3)4M is calculated for these conditions. and a study of the interaction of hexanal with actual Using the present kinetics data (and the parameter- H SO aerosolsbyGarlandetal.(2006).Asdiscussed izationsgivenabove),thebimolecularrateconstantis 2 4 above,anextensivekineticdatasetforacid-catalyzed about 2.9(cid:4)10(cid:3)3M(cid:3)1s(cid:3)1, and the reaction rate can aldol condensation self-reactions exists only for be calculated: 2.9(cid:4)10(cid:3)3M(cid:3)1s(cid:3)1(cid:4)[6.0(cid:4)10(cid:3)4M]2, acetaldehyde(Baigrieetal.,1985;EsteveandNoziere, which is equal to 1.0(cid:4)10(cid:3)9Ms(cid:3)1. Therefore, for a 2005).Thepresentrateconstantsfortheacetaldehyde 1hexperiment,thepresentkineticsresultspredictthat self-reaction are about 50% higher than the rate the reaction has proceeded about only 1% toward constantsdeterminedat65,75,and85wt%inBaigrie completion, and does not appear to provide support et al., but somewhat lowerthan thevaluedetermined for aldol condensation reactions as the explanation by Esteve and Noziere for 85wt% H SO . for the large growth in organic mass observed in the 2 4 Garland et al. (2006) studied the heterogeneous Jang and Kamens experiments. Similarly, Esteve and interaction of hexanal vapor with H SO aerosols Noziere (2005) have used their kinetics data for 2 4 with compositions between 30 and 96wt% and at acetone to show that the acid-catalyzed reactions of 298K. They report evidence for reaction at all acetone are too slow to be responsible for increased H SO compositions, with increasing reaction evi- organic mass observed in aerosols in smog chamber 2 4 dent at the highest acidities. Typical reaction experiments. Esteve and Noziere also asserted that conditions were as follows: 7Torr hexanal, 75wt% since atmospheric conditions are much less favorable H SO and a reaction time of 13min. In order to for such acid-catalyzed reactions than smog chamber 2 4 calculate the acid phase hexanal concentration, the conditions, the reactions of ketones with sulfuric acid appropriate Henry’s law coefficient is required. aerosolsareprobablynotresponsibleforhighorganic Unfortunately, there have been no measurements content of tropospheric aerosols. of this quantity for hexanal in H SO solutions. 2 4 However, if a conservative value of 100Matm(cid:3)1 3.9. Roleofaldehydealdolcondensationreactionson (roughly the value determined by Esteve and atmospheric sulfuric acid aerosols Noziere (2005) for acetaldehyde in 70wt % H SO and at 298K) is assumed, a hexanal While it is clear that acid-catalyzed accretion 2 4 concentration of 1.0M is calculated. Using the reactions involving aldehydes have been observed in present kinetics data (and the parameterizations many laboratory experiments, these experiments given above), the bimolecular rate constant is about typically use higher aldehyde concentrations (most 1.8(cid:4)10(cid:3)2M(cid:3)1s(cid:3)1, and the reaction rate can be often, much higher) than those observed in the calculated: 1.8(cid:4)10(cid:3)2M(cid:3)1s(cid:3)1(cid:4)[1.0M]2, which is atmosphere in order to drive the reactions to occur equal to 1.8(cid:4)10(cid:3)2Ms(cid:3)1. Therefore, for a 780s on a reasonable time scale. Obviously, kinetic data experiment, the present kinetics results predict that are needed to extrapolate from such laboratory thereactionwillgotocompletionseveraltimesover conditionstothoseoftheatmosphere,particularlyas undertheseconditions.Thisresultisconsistentwith a function of H SO composition and temperature. 2 4

Description:
Atmospheric Environment 41 (2007) 6212–6224. Kinetics of acid-catalyzed aldol condensation reactions of aliphatic aldehydes. Mia T. Casale a.
See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.