CA1294973C - Process for the hydroformylation of sulfur-containing thermally cracked petroleum residua and novel products thereof - Google Patents

Process for the hydroformylation of sulfur-containing thermally cracked petroleum residua and novel products thereof

Info

Publication number
CA1294973C
CA1294973C CA000549993A CA549993A CA1294973C CA 1294973 C CA1294973 C CA 1294973C CA 000549993 A CA000549993 A CA 000549993A CA 549993 A CA549993 A CA 549993A CA 1294973 C CA1294973 C CA 1294973C
Authority
CA
Canada
Prior art keywords
alkyl
aldehyde
alcohol
hydroformylation
aldehydes
Prior art date
Legal status (The legal status is an assumption and is not a legal conclusion. Google has not performed a legal analysis and makes no representation as to the accuracy of the status listed.)
Expired - Lifetime
Application number
CA000549993A
Other languages
French (fr)
Inventor
Alexis A. Oswald
Ram N. Bhatia
Alexander P. Glivicky
Edmund J. Mozeleski
John R. Hooton
Barry G. Brueggemann
Charles M. Smith
Current Assignee (The listed assignees may be inaccurate. Google has not performed a legal analysis and makes no representation or warranty as to the accuracy of the list.)
ExxonMobil Technology and Engineering Co
Original Assignee
Exxon Research and Engineering Co
Priority date (The priority date is an assumption and is not a legal conclusion. Google has not performed a legal analysis and makes no representation as to the accuracy of the date listed.)
Filing date
Publication date
Application filed by Exxon Research and Engineering Co filed Critical Exxon Research and Engineering Co
Application granted granted Critical
Publication of CA1294973C publication Critical patent/CA1294973C/en
Anticipated expiration legal-status Critical
Expired - Lifetime legal-status Critical Current

Links

Landscapes

  • Organic Low-Molecular-Weight Compounds And Preparation Thereof (AREA)

Abstract

ABSTRACT OF THE DISCLOSURE

A catalytic process for the hydroformylation of olefinic, sulfur containing thermally cracked petroleum streams to produce aldehydes and/or alcohol is disclosed. The catalysts are homogeneous transition metal carbonyl complexes. Especially preferred catalysts for low and medium pressure hydroformylation are cobalt and rhodium carbonyl hydride complexes in which some of the carbonyl ligands have been replaced by trivalent phosphorus ligands. In a preferred high pressure hydroformylation. the sulfur-containing naphtha and gas oil distillate feeds are produced from vacuum residua by high temperature thermal cracking. Such feeds contain more than 20% olefins with 1-n-olefins as the single major types. These olefin components are hydroformylation in the presence of a cobalt carbonyl complex to product a novel type of semilinear aldehyde and/or alcohol product containing an average of less than one alkyl branch per molecule.
The alcohols are converted to dialkyl phthalates and other esters having a unique balance of plasticizer properties. They are also useful for producing novel surfactants, particularly ethoxylated derivatives.
For the preparation of produces containing minimal concentrations of sulfur compounds, narrow distillate fractions of thermally cracked residua are preferred. In the C6 to C11 carbon range, single carbon fractions of sharply reduced aromatic hydrocarbon and thiophenic sulfur content can be obtained. These fractions of increased linear olefin content can be advantageously used as hydroformylation feeds in the derivation of low sulfur containing alcohols and related products of increased linearity.

Description

f 3 Thl~ l mon~lon provlde~ ~ c~e~lytlc proc~ for the hydro~or~ylAelon of cerealn oloflnlc, ~lfur coneA~nlng. ch~rM-lly cracked p~erol0u~ dl~elllat-s~ ro~dlly avalle~lo ~t low co~e, eo produco c~rtaln d~lr-bl~ llln~ar nld~hyd~J and ~lcohol~. by r~actlng tha olQfln co~pon~ntb wlth CO ~nd N2. Th~ c~caly~e~ ~rc yrof~r-bly dlssolvod eranJieson ~ctal c~rbonyl co~pl~xcJ. E~p~clally prof~rr~d catAly~e~ aro cob~le ~nd rhodLu~ carbonyl hydrld~ co~ploxc9 In ~hlch ~oo- of tho carbonyl llg-nd~ hav- b~n rcpl~c~d ~y er~v~lone pho~phorw ll~und~. A prcfcrrcd fcod 1~ produc~d by eh~ hlgh t~p8r~euro thor~l cr~cklng of vacuu~ r~slds, partlcul~rly by Fluld-coking and Fl~xlcoklng.
On~ asp~ct of eh~ dlsclo3ura ~ a d~cripelon of ehc types and truceur~3 of eha co~pound~ producod by ch~ ehor~l cr~ckin~ of p~erol~u~
re~lds. The n~phcha and ~ oll dl~tlllaea fracelon~ d~rlvod by the cr~eklng of v~euun reaLda In fluld~zod bad proeo~sos waro lnvo~lgaeed by a eonbSn~eion of hi8h reso1uelon eaplllary ga~ ehromaeography. mass speeero~acry and nuele~r n~gnaele re~on-neo speeeroscopy. The differenc eypa~ of ol-fln ro-etaneJ and cho pocenelal ~ulfur eo~pound inhlbieors vere partieularly an41yzod. Tho dlserlbuclon of the sulfur compound eoQponencs ln eha dlse~ eo foad~ uaJ an~ly2~d by u~lng a sulfur 5p~ciflc deeeeeor.
~ nothsr ~poec of cha dlselo3uro ls th~ eorrelatLon of che ~Crueeuro~ of tho l-n-olein and eha llne~r ineornal olafln reaetanc couponQne~ oi tha foed and tha vArlou~ cypa~ of tran31tlon mncal eomplex c~e~ly~es u~d wlth cha unlqua structura~ of che ss~illnear ~ld~hyda and aleohol produet~. Iho hl&h pras~uro eobalc earbonyl eo~plax eataly~ed hydrofor~yl~tlon of Cs co Cls naphsha and ga5 oll dl~tlllaea fraeclon~ and tha resuleing aldahydo product Qlxtura~ conslsclng ~o~tly of che eorra~pondlng n--ldahyd~s 2-methyl branehod a1dahyde3 and 2-subsclcuced aehyl and hl8h~r n-alkyl aldehydei are p~rrlcularly do~erlbod. Th~ trace sulfur eone~lnlng co~ponHne~ of cho aldchyd~- ~nd thoir alcohol derlvaCives ware al~o studlad by ~ulfur 3paelfle ga ehro~togr~phy.

~k 3~ 3 A further aspect of the disclosure is the description of the reactions of the present aldehyde and alcohol products. The esterification of the alcohols leading to phthalate ester plasticizers having unique properties is particularly discussed. Ethoxylated surfactant derivatives of the alcohols are also described in some detail.

PRIO~ ART VE~SUS THE ~ S~T INV~NTION
~ ydroformylation is a well-known reaction for the conversion of pure olefln streams with CO and H2 to aldahydes but has not been generally suggested for use on dilute olefin streams, such as petroleu~ distillates, which contain high concentrations of sulfur compounds and some nitrogen compounds. Streams containing these sulfu-r and nitrogen containing i~purities have been considered as unsuitable hydroformylation ~eedstocks.
Presen~ olefin feeds for hydro~ormylation are mostly propylene and its oligomers plu5 ethylene oligomers. The C7 to C13 alcohols derived from propylene oligomers and propylene/butenes copolymers are generally highly branched. In contrast, the Cg to Cls alcohols derived from ethylene oligomers are usually highly linear. Both types of higher alcohols are widely used intermediates in the production of plasticizer esters and ethoxylated surfactants. For most applications linear or semilinear alcohol intermediates are preferred. However, ehe ethylene oligomer feeds of linear alcohol production ara much more costly than the branched olefin feeds derived from C3/C4 olafins.
~ s a part of the present invention it was discovered that thermally cracked petroleum distillates, particularly those derived from residual fuel oil by Fluid-coking and Flexicoking, contain une~pectedly ma~or quantitias of linear olefins. These olefins are valued below distillate fuel co~t, because such cracked distillates have high concentrations of sulfur compounds and have to be extensively hydrogenated before thay can be used as distillate fuels. The olefin components are convarted to paraffins during such hydrogenation~.
Furtherm~re, it was found in the present invention, that the sulfur compounds in such thermally cracked petroleum distillates are mostly innocuous aromatic, thiophene type compounds rather than catalyst inhibiting mercaptans. This finding led to the discovery of the present hydroformyla~ion process which comprises reactlng ehe linear and lightly branched olefin components oi thermally cracked patroleum distillates -3 ~29~9~3 containing sulfur compounds with C0 and H2 to produce semilinear aldehydes and alcohols.
When such olefin componenta were reacted with C0/H2 in the presence of cobalt carbonyl complex catalysts at high pressure, the major aldehyde products were n-aldehydes, 2-methyl and 3-methyl substituted aldehydes, 2-ethyl and higher alkyl substituted aldehydes in the order of decreasing concentrations.
As such, the present procsqs produces novel, highly desired, semilinear chemical intermediates at a low cost. Due to the unique olefin composition of the present cracked distillate feeds, such a unique mixture of compounds cannot be produced by known processes.
~ hen using thermally cracked petroleu~ residua of naphtha distillatiGn range, it was found partlcularly advantageous to use narro~
boiling distillates rich in a particular l-n-olefin. Such distillates mostly contain compounds having the same number of carbon atoms per molecule. They were found enriched in 2-methyl-1-olefin and linear internal olefins. They have much lower arom tic hydrocarbon and thiophenic sulfur concentrations compared to the broad naphtha feed. They also provide hydroformylation products of more linear character than broad cut distillate feeds.
The process of tha present invention is particularly advantageous when the cracked petroleum distillate is a high bolling gas oil fraction containing 10 to 20 carbon atoms per molecule. In contrast to higher molecular wei~ht branched olefin~ derived by the oligomerization of C3/C4 olcfins, these gas oils are surprisingly reactive feeds for hydroforrylation without prior treatment.
A group of preferred thermally cracked distillates, not previously considered as a hydroformylation feed, comprises naphtha and gas oil fractions produced in fluidized coking units. Integrated fluidized coking processes such as Fluid-coking and Flexicoking represent a superior refinery me~hod for the conversion of residual fuel oil. The thermal cracking step of Fluid-coking and Flexicoklng is identical. However, Fluid-coking does not utilize the residual coke produced with the coker distillate while Flexicoking employs the coke by-product for the production of low thermal value gas. A discussion of these processes is found in U.S.
Patent Nos. 2,813,916; 2,905,629; 2,9Q5,733; 3,661,543; 3,816,084;
4,055,484 and 4,497,705.

The preferred Fluid-coking and Flexlcoking processes are low severity ther~sl cracking operations. Low severity is usually achieved by keeping the temperature relatively low in ths range of 482 to 538C (900 to 1000 F) while usin~ a long residence, i.e., contact, time of about 20 to seconds. Alternately, low severity can be achieved using high temperatures, in the order of 538 to 705~C tlOOO to 1300~) and contact times of less than 5 secondQ. In a long residenc~ time operation, additional amounts of the desired olefin compononts can be produced by reinjecting the heavy gas oil distillate products into the cracking lLne.
The residual fuel feeds for the above coking processes are usually vacuum residua which remain after most of the crude petroleum is removed b~ refinery distillation processes. As such, these r sidua typically possess boiling polnts above 565C (1050F) and have Conradson carbon contents above 15X. These residua contain ~ost of the undesirable components of the crude, i.e. sulfur and nitrogen compounds and metal complexes. On coking, much of the sulfur ends up in the dLstillate products. As a result of high tempera~ure thermal cracking, major amounts of olefinic components are also formed and become ma~or constituents of such distillates. In spite of their high monoolefin content such distillates generally wera not considered as hydroformylation feeds because of their high sulfur and conJugated diolefin content.
Other residual fuel type feeds for coking are produced from heavy asphaltic oils and tar sands. The hlghly olefinic distillates produced by the thermal cracking of heavy tar oil are suitable feeds for the process for the present invention. A particularly attractive feed is produced by the coking of Cold Lake and Athab~sca tar sands oil residues without prior removal of sulfur co~pounds.
While coker naphtha and gas oil distillates resulting from the high te~perature thermal cracking of residual oil can by hydroformylated as such, their further fractionation i~ preferred. This results in feeds of improved hydroformylation process characteristics which lead to superior aldehyde and alcohol productQ.
The aldehyd2 and alcohol products of the present hydroformylation process contain 20X by weight or more linear, i.e. nor~al, isomer. The preferred products con~ain 20 to 50X normal iso~ers, 3 to 20X 2-methyl branched compounds and 3 to 15X 3-methyl branched compounds. 2-Ethyl and higher 2-alkyl branched co~pounds represent another slgnificant type of constituents. The balancs is composed of monobranched aldehydes or ~?,~73 alcohols with minor amounts of dibranched aldehydes or alcohols. The average number of alkyl branches per product molecule is less than 1. As such the products have a unLquely branched semilinear character. They are considered to be novel products and constitute further embodiments of this invention.
The semilInear alcohol products of the present invention are attractive intermediates for plasticizer and surfactant products. The properties of these products critically depend on the branchiness of the alcohol inter~ediates. The dialkyl phthalate ester plasticizer showed a desirable combination of low temperature properties and heat stability.
The ethoxylated alcohol surfactants had superior wetting properties.
Surfactant and plasticizer derivatives of thes~ se~ilinear alcohols are further embodi~ nts of this invention.
Although sulfur compounds in genaral were regarded as catalyst inhibitors, the production of alcohols or aldehydes via the hydroformylation of the olefinic components of so~e refinery streams has been previously suggested. For instance, U.S. Patent No. 4,454,353 to Oswald et al., issued June 12, 1984, teAches the use of trihydrocarbyl silyl substituted diaryl phosphine transition metal carbonyl hydride complex hydroformylation catalysts with ~refinery streams of olefins, containing paraffin by-products such as Cl to C20 paraffins...n.
Haag and Whitehurst in U.S. patents 4,098,727 and 4,487,972 discloqe the production of aldehydes and alcohols via tha hydroformylation of olefinic streams in the presence of insoluble, polymer anchored complexes of Group VIII metals with nitrogen, sulfur, phosphine and arsine ligands. Exa~ple 32 Yhows the hydroformyla~ion of a cracked gasoline feed containing 230 ppm sulfur in the presence of a rhodium amine complex attached to a styrene-divinylbenzene copolymer.
The process disclosed in U.S. Patent No. 4,417.973 to Angevine et al., is one for ~upgrading~ various straight chain olefin-containing feedstocks, such as shale oil, FCC light cycle oil, and coker liquids, to branched paraffins. The process involves the sequential steps of hydroformylation and hydrotreating/hydrogen reduction, preferably, in ehe presence of heterogeneous supported Co/~o catalyst. The reaction products of the hydr~formylation s~ep were neither separated nor idantified. The final products are branched p~raffins. The sulfur content of the various feedstoc~s are show~ in the Examples to be 0.29 to 1.33 wt. X.

-6~ 73 Other disclosures discussing the use of cobalt-based homogeneous catalysts are known.
For instance, a series of papers by Marko et al. teach the reaction of dicabalt octacarbonyl, a hydroformylation catalyst precursor, with elemental sulfur and organic sulfur compounds. Various sulfur-containing cobalt complexes were isolated. Reactlons with sulfur led to [Co2StCO)s]n and Co3S(CO)g. See, Chem. Ber., 94, 847-850 (1961);
Chem. Ind., 1491-1492 (1961); Chem. Ber., 96, 955-964 (1963). Hydrogen sulfida is said to react to give the sa~e complexes. Mercaptans and disulfides lead mainly to sulfide derivatives of cobalt trimers and tetramers. Marko et al. states that, under hydrofor~ylation conditlons, all these complexes are converted to catalytically inactive cobalt sulfide [Chem. Ber., ~, 926-933 (1964).] Cobalt thioether complexes are also said to be either inactive or less active in hydroformylation than unsubstituted dicobalt octacarbonyl ~Acta Chim. Sci. Hung., 59, 389-396 (1969)].
Another series of papers by Marko and co-workers describes the hydroformylation/hydrogenation of C6/Cg olefins present in cracked gasoline. The papers describe a process for converting a sulfur-containing C7 ~raction of cracked gasoline using a 1 to 2 ratio of hydrogen ~o carbon monoxide at 200C under 300 atm (4,409 p3i) pressure to produce 85X octyl alcohol, an intermediate for a dioctyl phthalate plasticizer, with lOX
hlgher boiling by-product formation tJ. Berty, E. Oltay and L. Marko, Chem.
Tech., ~Berlln~ 9, 283-286 (1957); M. Freund, L. Marko and J. Laky, Acta Chem. Acad. Sci. Hun~., 31, 77-84 (1962)). Under these reaction condition~, using cyclohe~ene as a model olefin, ethyl mercaptan and diethyl disulfida were found to be strong inhibitors of hydroformylation even in small amounts while diethyl sulfide and thiophene had no effect in molar concentration~ up to tenfold of cobalt [L. Marko, Proc. Symp. Coordn.
Chem. Tihany, Hungary, 271-279 (1964)]. Similar but more pronounced effects were observed on ~he hydrogenation of aldehyde intermediates to alcohols [J Laky, P. Szabo and L. Marko, Acta Chim. Acad. Sci. Hun~., 46, 247-254 tl965)]. Sulfur containing cobalt trimers, e.g., of the formula Co3(CO)gS and Co3(CO)6(S)(SR) were postulated as intermediates in the converslon of active Co2(CO)g into soluble inactlve CoS [L. Marko and M.
Freund, Acta Chi~. Acad. Sci. Hung., 57, 445-451 (1968)].
Russian researchers, particularly Rudkovskii and co-workers, also published a series of articles on the hydrofor~ylation of olefin components in patroleu~ distillstes with dicobalt octacarbonyl catalyst. These 7 ~ 73 distilla~0s were not characterized chemically. One paper describes the production of Cll to C17 alcohols from high boiling distillate fractioos of contact coking. The process entails hydroformylation, preferably at 170C
and 300 atm (4409 psi), followed by hydro~enation in a mixture with unreacted hydrocarbons over a 2NiS.WS2 catalyst [R. A. Alekseeva, D. M.
Rudkovskii, M. I. Riskin and A. G. Trifel, Khim. i Tekhnol. Topliv i Masel 4 (5), 14-18 (1959)]. Another paper describes a similar hydrofor~ylation of lower molecular weight cracked gasoline olefins [D. Rudko~skii, A. G.
Trifel and K. A. ~lekseeva, Khim. i Tekhnol. Topliv i Masel, 3(6), 17-24 (1958)]. Suitable C7 to C8 naphth~ feeds from thermal cracking of a mixture of petroleum fractions, phenol extr~cts and petrol~um were later described [P. K. Z~iewski, T. N. Klyukanova and G. ~. Kusakina, Neft. i Gas Prom., Inform. Nauchn. Tekhn. Sb. (4) 48-49 (1964)].
Anotber journal article, appeared in a Russian ~ournal, Khim.
Tekhnol. Gotyuch. Slantssv i Produktov Lkh Pererabotki, on pages 325 to 332 of the 13th issue of 1964, and was authored by N. I. Zelenin and co-workers. This publication considered the hydroformylation of che olefin components of shale gasoline and diesel fractions to produce plasticizer and surfactant alcohols. It particularly discussed the removal of sulfur compounds which can be hydroformylation inhibitors.
A research report, Forschungsbericht T-84-064, was made to the German Federal Department of Research and Technology in April 1984. The authors, B. Fell, U. Buller, H. Classen, J. Schulz and J. Egenolf disclose the hydroformylation of a Cs to C6 cracked gasoline between 150-175C at 200 atm (2939 psi) in the presence of 0.4-0.2X cobalt to obtain oxo-products with 65X selectivity. The use of a triphenyl phosphine rhodium complex ba~ed catalyst system at this high pressure was reported to result Ln littl~ conversion.
Two monographs on the organic chemistry of carbon monoxide by Falbu and co-workers of Ruhrchemie include ma;or chapters on hydroformylation. The effsct of hydroformylation of cobalt catalys~
poisons, particularly sulfur compounds, is su~marized on pages 18 to 22 of the first monograph [J. Falbe, Carbon Monoxide in Organic Synthesis, Chapt~r I, The Hydro~ormyl&tion Reaction (Oxo Reaction~Roelen Reaction), pages 1 to 75, Springer Verlag, New York tl970)1. The second monograph also reviews the effect of poisons on modified rhodiu~ catalysts and concludes that these catalysts, due to thoir low concentration, are more susceptlbl~ to poisoning [New Syntheqis with Carbon Monoxide, Ed. J. Falbe, Chapter 1 by B. Cornlls, pagQs 1 to 224, particularly page 73, Springer Verlsg, Naw York, 1980].
Higher aldehydeY derived via hydroformylation are known versatile chemical intermedia~eq. Thcy are utilized for the synthesis of primary alcohols, carboxylic acids and amines. The so called oxo-alcohols are the most important products. They are most widely used in the preparation of phthalate ester plasticizers and surfactants. Howe~er, known methods for the preparation of oxo-aldehydas and alcohols have carbon number and/or product linearity llmitations.
Highly line~r oxo-alcohols are the most desired for most applications. However, the~r preparat~on requires completaly linear olefin feeds derived fro~ ethylene which ~re prohibitively expensive for many applications. Highly branched oxo-~lcohols derived via the hydroformylation of propylane oligomers are less costly to produce but their plasticizer derivatives have poorer low temperature properties and their surfactant derivatives are less biode~radable.
More recently, U.S. Patent 4,598,162 by D. Forster, G. F.
Schaefer and G. E. Barker disclosed the preparation of aldehydes and alcohols via the aldolization of axo-aldehyde~ containing little branching in the 2~position. The alcohols derived via this route are more biodegradable than the prior art branched compounds. However, their preparation requires an additional step and leads to products having more than one branch per molecula.
Overall, the prior art taught away from the hydroformylation process of thc present invention rather than suggesting it. In general, the use of cracked petroleum distillates containing high concentrations of sulfur was to be avoided. Soluble transition metal carbonyl complexas containing trivalent phosphoruQ ligands were never used successfully for the hydroformylation of such distillstes. Known low pressure hydroformylation processes have low sulfur limits for the feeds.
Although the high prassure hydroformylation of cracked gasolina of relativoly low sulfur content wa~ axtensl~ely studied by Marko et al. in the presence of added dicobAlt octacarbonyl, the feeds and conditions of the present process were neither used or sug~ested. It was not proposed to utilize coker distillata feeds of high linear olefin and sulfur compound content for tha production of aldehydes and alcohols by hydroformylation.
The high prossura, cobalt catalyzed C7 gasoline hydrofor~ylation/hydro-genation process Marko et al. developed is run at 200C and produces 9 ~ 3 alcohol~ in one step. In contrast, th~ temperature range of the present hi~h pr0ssure cobalt catalyzed process is 110 to 180~C, preferably 120 to 145C and the main products are aldehydes. Pure alcohol products in this process are produced in a separaee stap.
The present cobalt carbonyl complex catalyzed high pressure proce~ employ~ Cg to C20 di~tillatQ fQeds pxoduced by hi~h temperature fluid-coking of vacuum resids. These feeds contain more than 0.1% sulfur and more than 20X olefins of a unique isomer composition. More than 30X of ehe total olsfins present are of Type I. More than 10~ of the olefins are of Typ~ II. The most prevalent Typa III olefin components are 2-methyl-1-olefins.
Due to the spacific linear olsfinlc character of the present feeds, such hydroformylations produce uniqua ald~hyde and alcohol products of a semilinear charac~er having less than one branch per molecule. The ma~or components of the primary aldehyde products are n-aldehydes, 3-methyl-branched aldehydes, and 2-methyl-branched aldehydes. Much of the rest are 2-ethyl or higher 2-n-alkyl-branched ~ldehydes. However, the amount of higher 2-alkyl-branched compounds is much les3 than in prior art compositions prepared via aldolization. On hydrog~nation they provide the corresponding alcohols. Such aldehyde and alcohol compositions cannot be directly prepared by any othar method. Their prepar~tion by blending the appropriate components would be economically prohibitive.
The novel semilinear alcohol product3 of the present inventlon can be converted to ester plasticizers and ethoxylated surfactancs of unique properties. The C6 to C13 alcohols provide the corresponding dialkyl phth~lates having a combination of supsrior low temperature properties and high temperature stability compared to branched alcohol derlvatives. The Cg to C30 alcohols lead to ethoxylated surfactants of hi&h biodegr~dability and superior wetting properties. In both, the case of pla~ticlzers and surfactants, the unique properties are attributable to the unique semilinear structure of the alcohol precursors.
Dialkyl phthalate ester are a well known, large volu~e group of placticizers for polyvinylchlorida. As such they comp~te on the basis of their propertiac and cost. Fro~ the viswpoint of mose of the desired propereies, particularly the low eemperature properties of plastlcizad PVC, phthalate esters derived from linear alcohols are supsrior to derivatives of highly branched primary alcohol~. Howev~r, highly branched alcohols can be producad in a broad carbon rznge at a cost ~ignificantly balow that of ~$4~973 linear alcohols. Thus, there has been a continuing effort to produce low cost, less branched primary alcohols and their mixtures. However, to date, no low cost primary plasticizer alcohol with less than one branch per molecule is availabla.
Ethoxylated higher alcohols are a highly important class of nonionic surfactants. They are dominating the detergent industry where biodegradability is important. They ar0 also widely us~d as sulfate derivatives. Most ethoxylated higher alcohols are derived from costly linear alcohols to enhance their biodegradability. The higher linear alcohols are solids and, as ~uch, difficult to handle. In contrast, the present detergent range semilinear alcohols are low cost liquids of biodegradable character. As such, they combine the advantages of both branched and linear alcohol surfactant intermediates.
None of the references teach either alone or in co~bination the presently described and claimed process and~or products.

D~S~RI~TIO~ OF TECL1~33E~
Figur~ 1 shows the capillary gas chromatogram of a Fluid-coker naphtha feed $n the C4 to C12 range, with an indication of the major l-n-olefin and n-paraffin components by a flame ionization detector and the ma~or thiophenic components by a sulfur specific detector.
None of the references teach either alone or in combination the presently described and claimed proce3s and products.
Figure 2 shows the 400 MHz proton nuclear magnetic resonance spectrum of the olaiinic protons of Fluid~cokar naphtha feed, with an indication of the chemical shift regions of various types of olefins.
Fi~ure 3 shows the capillary gas chromatogra2 of the C6 fraction of a Fluid-coker naphtha feed, with an indication of the ma~or olefin and parafiin corponents.
Figure 4 shows tha capillary gas chromatogra~ of the G8 fraction of Flexicoker naphtha feed with an indication of the ma~or hydrocarbon components.
Figure 5 shows the sulfur specific capillary gas chromatograms of the narrow and broad C8 fractions of Flexicokar naph~ha with an indication of the main methylthiophene and dimeehylthiophene components.
Figure 6 shows the capillary gas chromatogram of che Clo fraction of a Fluid-coker naphtha faed ~ith an indication of the major olefin, paraffin and aro~atic components.

Figure 7 shows the capillary gas chromatogram of the light Fluid-cokar gas oil feed in the Cg to C16 range, with an indlcation of the major l-n-olefln and paraffin components.
Figure 8 shows the 500 MHz proton nuclear magnetic resonance spectrum of light Fluid-coker gas oil feed, with an indication of the olefinic, paraffinic and aromatic components.
Figure 9 shows the capillary gas chromatogram on a highly polar column of a C12 fractLon of ligh~ Fluid-coker gas oil, with separation of various types of aliphatic and aromatic components and sulfur compounds.
Figure 10 shows the capillary gas chromatogram of a Fluid-coker light gas oil mixture after ~rioctyl phosphine cobalt complex catalyzed hydroformylation, with an indication of the ma~or n-paraffin and capped n-alcohol components.
Figure 11 shows the capillary gas chromatogram of a Clo Fluid-coker gas oil after trieehyl phosphine cobalt complex catalyzed hydroformylation, with an indication of ths isomeric Cll alcohol products formed.
Figure 12 shows the capillary gas chromatogram of a Fluid-coker naphtha mixture after cobalt catalyzed hydroformylation, with an indication of the major n-paraffin and n-aldehyde components.
Figure 13 shows the capillary gas chromatogram of ehe aldehyde region of the reaction mixture obtained in the cobalt catalyzed hydroformylation of a C6 Fluid-coker naphtha fraction.
Figure 14 shows the capillary gas chromatograms by flame ionization and sulfur specific detectors of tha reaction mixture produced by the cobalt catalyzed hydroformylation oP a C8 Flexicoker naphtha fraction.
Figure 15 shows the capillary gas chromatogram of a Clo Fluid-coker naphths after cobalt catalyzed hydroformylation, with an indication o the isomeric Cll aldehyde produces formed.
Figure 16 shows the packed column gas chromatogram of a Clo Fluid-coker naphtha after cobalt catalyzed hydroiormylation, with an indication of the Cll aldehyde products and dimer and trimer by-products.
Figure 17 shows the capillary gas chromatogram of a Fluid-coksr light ~as oil mixture after cobalt catalyzed hydroformylation, with an indication of the ma~or n-paraifin and n-aldehyde components.

-12- 1 ~ Ci~ 3 Figure 18 shows the capillary gas chromatogram of the aldehyde region of the reaction mixture obtained in the cobalt catalyzed hydroformylation of C12 Fluid-coker light gas oil fraction.

So~U81LI~ IL_INV~NTION
ThLs invention describes a hydrofornylation process in which the olefin components of a cracked petroleum distillate fractLon containing substantial amounts of l-n-olefins and sulfur bearing compounds are reacted with carbon monoxide and hydrogen in the presence of a homogeneous Group VIII transition metal carbonyl cornplex catalyst. The invention is also concerned with the novel products of the present process. These products are aldehydes and/or alcohols of largely lineat- character and as such pre~erably have less than one alkyl branch per molecule on the average.
The products may be separated by distillation fro~ the unreacted components of the distillate feed.
The preierred catalysts are soluble rhodium or cobalt carbonyl complex catalysts. The complex may be modified by a trivalent phosphoruc~
arsenic, nitrogen and/or sulfur li~and. Triorgano-phosphine ligands are most preferred. Cobalt carbonyl catalysts may also desirably be used without added phosphorus ligands.
The reaction conditions under which the feeds may be hydroformylated cover broad ranges. Temperatures ranging from 50 to 250C
and pressures ranging from essentially atmospheric to 5000 psi (340 atm) may be used. The more preferred condieions depend on the eype of the olefin to be reacted and the type of transition metal catalyst to be used.
When phosphorus ligand rhodium complex based catalysts are employed, low pressures between 50 and 2000 psi (3.4 and 136 atm), preferably 100 to 1500 psi (~.8 to 102 atm), are used. A broad range of temperatureq preferably from 50 to 250C, more preferably from 80 ~o 200C, can be used.
Phosphine cobalt complex catslys~s can be advantageously employed at pressures between 500 and 4500 psi (34 and 305 atm), preferably between aboue 500 and 2500 psi (34 and 170 atm), and at reactLon temperatures between 150 and 200C.
High pressure cobalt catalysts, in the absence of added ligands, require pressures between 2500 and 6000 psi (170 and 408 atm), preferably between 3000 and 4500 psi (204 and 306 at~). They are preferably employed between 100 and 180C, more preferably between 110 and 170C, most -13 ~Z~ 3 prsferably betwesn 120 and 145C. Higher pressures of reactant gas, spacifically C0, allow the use of higher reaction temperatures without catalyst decomposition and/or deactivation.
In summary, the dependence of reaction conditions on the type of catalyst systems employed is shown by the following tabulation:

Group VIII Trivalent _______~eaction ConditiQn~ __ Metal P LigandTemperature _ P~rU~gJQL________ Employed Employed C psi atm e~ . _ __ Rh Yes 50 - 250 50 - 2000 3.4 - 136 Co Yes 150 - 200 500 - 4500 34 - 306 Co No 100 - 180 3000 - 4500 204 - 306 In ehe prssent process, the feed for the high pres~ure cobalt catalyst contains l-n-olefins as the major type of olefins snd is derived from the pstroleum residua by FlexLcoking or an equivalent high temperature thermal cracking process. Starting with this feed, the present process provides aldehydes and/or alcohols of a highly linear character having less than one alkyl branch per molecule on an average. This feed and product are also preferred for the other catalysts.
The preferred thermally cracked distillate feeds have a further increased l-n olefin content and a reduced aromatic hydrocarbon and sulfur content. In the C6 to Clo feed range this is advanta~eously achieved by a process additionally comprising the fractional distillation of cracked naphtha separating narrow faed frac~ions containing mainly linear aliphatic hydrocarbons from fractions containing ma;or amounts of aromatic compounds including thiophene3.
Th~ preferred high pressure cobalt cataly~ed process of the present hytroformylation process i5 partlcularly suitable for the conversion of the olefins of the pressnt fesd to novel se~ilinear aldehydes having one carbon more than the parent olefins. The structure of the aldehyde~ contQining less than one alkyl branch per molecule refleces the unique mixture of the starting olefins. The major components of tha preferred aldehyde compo3itions are n-aldehydes, 2-methyl-aldehydes and 3-mdthyl-aldehydes derived from the major l-n-olefin and l-methyl-l-olefin components of the feed.
The invention is also concerned with tha derivatives of che primary aldehyde products. These aldehydes can be hydrogenated during - 14 ~ 3 and/or after the hydroformylation process to provide the corresponding mixture of semilinear alcohols. Elther the aldehydes or the alcohols can bs converted to ths corresponding amines and quaternary ammonium compounds The novel alcohol compositions can be converted to valuable ester plasticizers and ethoxylated surfactants. Polyvinylchloride plasticized with ~he phthalate esters shows a unique combination of low temperature flexibility, high temperature stability and reduced volatility attributable to present semilinear alcohol intermedia~e Similarly, the sthoxylated and propoxylated surfactant deri~atives of these alcohols show a desirable combination of biodegradability and wetting properties. Such surfactants will generally contain about 1 to 30 mole3 of ethylene oxide or propylenP
oxide per mole of semilinear alcohol.

D~_C~IPTI0~ 0~ ~HE PR~F~R~D ~M~OPIM~NTS
This in~ention describes a hydrofor~ylatlon process for the production of aldehydes and/or alcohols of a largely linear character, i.e., prod~cts stream having preferably less than on0 alkyl branch per mole on the average, from a cracked petroleum distillate feedstock containing substantial amounts of l-n-olefin~ and sulfur co~pounds. The process comprises reacting the distillate with C0/H2 in the presence of a Group VIII transition metal complex catalyst.
As such, the present hydroformylation process comprises reacting with hydrogen and carbon monoxide an olefinic cracked petroleum distillate feed, particularly in the C8 to C3s carbon range, preferably produced fro~
petroleum residua by high temparature thermal cracking, and coneaining l-n-olefins as the ma~or type of olefin components, the percentage of Type I olefins being preferably more than 30X, said feeds also containing organic sulfur compounds in concentrations preferably exceeding O.lZ, more preferably exceeding lX.
The hydrofor~ylation reaction is carried out at temperatures between about 50 and 250~C and pressures in the range of S0 and 6000 psi ~3.4 and 408 atm) dependent on the particular catalyst employed.
The reaction takes place in the presence of affective a~ounts of a Group VIII transition metal carbonyl complex catalyst preferably selected from the ~roup of Fe, Co, Rh, Ru, Ir and 03, more preferably Rh, Co, Ru and ~r, most preferably Co or Rh, a preferred gro~p of complexes being modified by a trivalent phosphorus ligand, preferably triorgano-phosphine or phosphite ester.

-15- ~2~a~3 Such hydroformylations produce aldehydes and/or alcohols, preferably aldehydes of a semilinear character, preferably having an average of less than one alkyl branch per molecule. These products more preferably contain n-aldehydes, 2-methyl and 3-methyl branched aldehydes as the major products, most of the rest being mainly various 2-ethyl or higher 2-n-alkyl branched aldehydes. The reduction of these aldehydes by hydrogen to the corresponding alcohols is preferably carrled out in a separate step in the presence of a sulfur insensitive catalyst, preferably based on Co, Mo, Ni, W in a sulfided form.
Thu~, according to another aspect of the invention, a hydroformylation-hydrogenation process comprises reacting the above described olefinic cracked petroleum distillaee fzed with carbon monoxide and hydrogen under the conditions already defined to produce said aldehyde products and ~han reacting said aldehydes and temperatures bet~een 100 and 220C in the presence of a catalyst in effective a~ou~ts to produce the corresponding alcohols of a semilinear character having an average of less than one alkyl branch per molecule.
According to a further aspect of the invention, the novel aldehyde and alcohol compositions prepared via the present process are dsscribed. These isomeric aldehyde compositionq comprise C7 to C21 mostly saturated aliphatic aldehyde mixtures of a semilinear character having an average of less than one branch per ~olecule. They contain more than 30X
normal alkanal and ma~or amounts of 2-methylalkanals and 3-methylalkanals and minor amounts of 2-ethyl and higher 2-n-alkylalkanals. Similarly, the isomeric alcohol compositions comprise C7 to C21 saturated aliphatic alcohol mi~tures of a semilinear charac~er having an average of less ehan one branch per molecule. These alcohols contain more than ~OX, preferably more th~n 30X, normal alkanols, ma~or amounts of 2-methylalkanols and 3-methylalkanols and minor amounts of 2-ethyl and higher 2-n-alkylalkanols.

Dist~ eo Ye~ts The cracked petroleum distillate feeds of the present hydroformylation procass arz preferably derived via ther~al cracking.
Thermal cracking processes produco hydrocarbons of more linear olefinic character than catalytlc cracking. Thz presence of linear olefin components, partlcularly l-n-olefins, in the cracked dlstillates is important for the production of normal, non-branched aldehydes and mono-branched aldehydas using hydroformylation. For example, the ~l2~ 73 hydroformylation of l-hexene can produce n-heptanal as the maln n-aldehyde product and 2-methylhexanal as the minor iso-aldehyde product. These in turn can be hydrogenated to the corresponding alcohols:

CH3(cH2)3cH~cH2 ~ CO ~ H2 ~ CH3(CH2)sCHO + CH3(CH2)3CHCHO
c~3 ~ H2 CH3(CH2)5CH20H ~ C~l3(cH2)3cHcH2oH

The llnear nor~al aldehyde and alcohol products sre generally more desired than ehe branched iso-compounds as intermediates for the production of high quality plastici~ers and surfactants. Among the iso compounds, the 2-methyl branched products have the least adverss effect on product quality.
The percentage of 1-n-olefin components of thermally cracked peeroleu~ distillates generally increases with the temperature of cracking.
Therefore, the distillate products of high temperature thermal cracking processes such as Fluid-coking and Flexicoking are preferred feeds for the present process. Delayed coking, which is normally operated at a lower temperature, can also produce suitable feeds for the present process when operated at sufflciently high temperature. Other less preferred, milder cracking processea such as the thermal cracking of gas oils and the visbreaking of vacuum residues can also produce distillate feeds for the present process. Sultable distillate feeds can also be prepared in thermal proce~es e~ploying a plurality of cracking zones at different tempera~ures. Such a proc~ss is described in U.S. patents 4,477,334 and 4,487,686. Eacb of these thermal cracking processes can be ad~usted to increase the olefin contents of their distillate products. Higher distillate fractions of steam cracking can also be used as a feed in the present process.
The olefin content of the cracked distillate feeds of the present invention is above 20X, pre~erably above 30X, more preferably abo~e 40X.
The l-n-oleflns are preferably the major type of olefin components.
In the high pressure operation of the present process, using cobalt ~arbonyl co~plexes without any added phosphine ligand, the feeds should be thermally cracked distillates containing l-n-olefins as the major olef$n type. These feedstocks are preferably produced by the FLEXICOKINC
process or FLUID-COKING process and slmilar high temperature coking processes.
Distillate fractions of cracking processes can be hydroformylated without prior purification. However, ehe cracker distillate feeds may be treated to reduce the concentration of certain sulfur and nitrogen compounds prior to the hydroformylation process. These impurities, particularly the mercaptans, can act as inhibitors to the hydroformylation step. The disclosed process is operable in the presence of the impurities but adjustments to the catalyst level and/or to the reactant gas partial pressure (notably the CO pressure) are preferably made to compensate for the inhibition by the sulfur compound~.
One method for the removal of mercaptans, is selective extraction. Most of the extractive processes employ basic solvents.
Examples o~ such processes include the use of aqueous and meehanolic sodium hydroxide, sodium carboxylats (isobutyrate, naphthenate) sodium phenolate (cresolste) and tripotassium phosphate. Sulfuric acid of carefully controlled concentration and temperature can be also used although it is less selective than caustic. For example, a 30 minute treatment with 12X
H2SO4 between 10 and 15 can be used.
The preferred cracked dLstillates of the present feed contain relatlvely high amounts of organic sulfur compounds. The sulfur concentration is preferably greater than O.lX (1000 ppm), more preferably greater than lX (10000 ppm). The prevalent sulfur compounds in these feeds are aromatic, mainly thiophenic. Most preferably the aromatic sulfur compounds reprasant more than 90X of the total. This finding is important for the present process since thiophenes, benzothiophenes and similar aromatic sulfur compounds do not inhibit hydroformylation.
~ or the removal of sulfur, as well as nitrogen compounds, adsorption on columns packed polar solids, such as silica, fuller's earch, bauxite, can also be used. ~reating columns containing such adsorptive solids can bs regenerated, e.g., by steam. Alternatively, ~eolites can be used to enrich the present feed~ in l-n-olefins and n-paraffins.
The aromatic hydrocarbon components of the feed can also be removed together with tha aromatic sulfur co~pounds, preferably by methods ba~ed on the Increased polarity of aromatics compared to ths aliphatic components. Selective solvent extraction methods usin~ a polar solvsn~

-18- 4 ~ ~ ~

such as acetonftrils ~ay be Pmployed for extracting the polar components.
As a feed for extraction, preferably narrow distillate fractions of up to 3 carbon ran~e are used.
Finally, sulfur compounds can also be converted to easily removable hydrogen sulfide by pasqin~ the cracked distillate through a high temperture fixed bed of either bauxite or fuller's earth or clay, preferably between 700 to 750~C. Ona disadvantage of this catalytic desulfurization method is the concurrent isomerization of olefin.
The cracked refinery distillata feed is prsferably separated into various fractions prior to hydroformylation. Fractional distillation is the preferred method of separation. The different distillate fractions contain different ratfos of the various types of olefin reactants and have different inhibitor concentrations. The preferred carbon range of the thermally cracked feeds is Cs to C3s. The C8 to C2s range is more preferred. The most preferred r~nge is Cll to C20. It is desirable to limit the carbon number range of any given distillate feed by efficient fractional distilLation to 5 carbons, preferably three carbons, more preferably one carbon, to allow efficient separation of the products from tha unreacted feedstock.
For example, a cracked distillata feadstock fraction might contain hydrocarbons in the C7 eo Cg ranga. The maln components of such a fraction would be C8 hydrocarbons. Upon hydroformylating the olefinic components of such a fraction, C8 to Clo (mainly Cg) aldehydes and alcohols would be obtained. These oxygenated products all boil higher than the starting C7 to Cg hydrocarbons. The products could therefore be separated by distillation from the unreacted feed fraction.
For the pr~paration of plasticizer alcohols, olefin feeds containing fron 5 to 12 carbon atoms ara prsferred. These can be convarted to C6 to C13 aldehydes and in turn to C6 to C13 alcohols. The more proferred feQda contain C8 to C12 olefins and as such provide Cg to C13 ~lcoholq. Tha mo~t preferred feeds are Clo to C12 olafin-Q. The alcohols may be reacted with phthalic anhydride to produce dialkyl phchalate pla3elcizers of appropriate volatility. The more linear the character of the alcohol employed, tha better are the low tempera~ure properties of the plasticized products, e.g., plasticizad PVC. The praferred feeds of the present invantion are unlquely advantageou-~ in providing low cost olefins for the darivation of high val~e plasticizers.

12~ ,3 For the preparation of surfactants, higher molecular ~eight olsfins are usually preferred. Their carbon numbers per molecule range from Cg to C3s. These feeds can be used for the derivation of Cg to C36 aldehydes, C12 to C20 olefin feeds leading to C13 to C21 surfactant alcohols are more preferred. These aldehydes can be either reduced by hydrog~n to the corresponding alcohols or oxidized by oxygen to the corresponding carboxyl~c acids. The alcohols can then be converted to nonionic surfactants, e.g., by ethoxylation; anionlc surfactants, e.g., by sulfonation and cationic surfactants, e.g., by amination or cyanoethylation followed by hydrogenation.

Ol~fln R~acean~C~ und~
The main olefin reactant componan~s of the pre~snt feed are nonbranched Types I and II or mono-branched Types III and IV as indicated by the follo~ing for~ulas (R - hydrocarbyl, preferably non-branched alkyl):

R-CH~CH2 R-CH-CH-R R-C-CH2 R-C-CH-R
R R
I II III IV
non-branched linearmono-branchedmono-branched termLnal internal terminal internal The concentration of Type I olefins is preferably greater than 30X of the total olefin concentration. The percentage of Type II olefins is greater than 15X. Type V olefins of the formula R2C~CR2 are essentially absent.
The n-alkyl sub~tituted Type I olefins, i.e., l-n-olefins, are generally present at the highest concentration in thermally cracked distillateq a~ong the YariOus olefinic species. The main product of l-n-olsfin hydroformylation is the corresponding n-aldehyde having one carbon mor~ than the reactant. The hydrofor~ylation of Type II linear inte m al olefin~ ant Type III mono-branched terminal olefins provides mono-branched aldehydes and in turn to alcohols:

-20- ~Z~ 3 R-CH-CH-R CO/H2 E~CH2CHCHO H2 ~ RCH2CHCH20H
R R
II

R-C-CH2 CO/H2 ~, RCHCH2CHO H~2 .. RcHcH2cH2oH
R R R
III

The hydroformylation of Type IV mono-branched olef~ns leads to dibranchet products.

R R R - R R

Characteristically, the alkyl branches of the Type III and IV
olefins are mostly mathyl groups. The absence of long alkyl branches is important in determining the properties of the oxo-derivatives of these feed components. Types I, Il, III and IV olefins have a decreasing reactivity in this order. Thus it Ls possible, using the selective catalytic process of the present invention, to convert either to Type I, or the Types I and II, or the Types I to III olefins, selectively to products containing ton an av0rage) less than one branch per molecule. Of course, the most lin0ar products can be derived by hydroformylating only the Type I
olefin.
Type II linear internal olefins can also be converted to non-branchod ald~hyde~ and alcohols via the present process. To achieve this conversion, combined isomerization-hydrofor~ylation may be carried out. This process uses an intern~l-to-terminal olRfin isomeri7ation step followed by a selective hydroformylation of the ~ore reaceive ter~inal olefin isomer. For example, in: ~he case of 3-hexene, the following reactions are involved: `
, CH3CH2CH-CHCH2CH3 ~ CH2CH2CH2CH-CHCH3 ~CH3CH~CH2CH2CH2CU2CHO /~2 CH3cH2cH2cH2cH-cu2 :~ :

:

-21- ~2~ 3 Due to lts much greater reactiYity, the terminal olefLn is selectively hydroformylated even though its equilibrium concentration is smaller than those of the internal olefin isomers. The cobalt-phosphine~complex-based catalyst systems are particularly effective for coupling the isomerization and hydroformylation reactions.

C~2 S~nqu,~g~
As a reactant gas for hydroformylating the olefin components of the pre~ent feed, mixturcs of H2 and CO, preferably in ratios ranging from 1- 2 to 10-1, can be used. Ratios between 1 and 2 are preferred. When reacting higher oiefins, most of the total reactor pressure is that of H2 and CO. High H2/CO pressures, particularly high CO partial pressures, usually stabilize the catalyst system. The CO as a ligand competes with the sulfur compound ligands for coordination to the transition metal to form the metal carbonyl complex catalyst. The partial pressure of carbon monoxide affects the equilibria among catalyst complexes of different stability and selectivity. Thus, it also affects the ratio of linear to branched products (n~i) and the extent of side reactions such as hydrogenation.
High CO partial pressures are particularly important in forming and stabillzing the desired carbonyl complq~ catalysts of high pressur cobalt hydroformylation. They stabilizs the catalyst complex against deactivation by the sulfur compound components of the feed. In a preferred operation, the active catalyst system is produced at a lo~ H2/CO ratio.
Thereafter, the cstalyst is operated at incraasing H2/CO ratios.
The effect of CO partial pressure on the n/i ratio of aldehydes and alcohol products is particularly important in the presence of rhodium complexes of tri~alent phosphorus ligands, particularly phosphines.
Phosphine ligands increase the strength of CO coordination to rhodium.
Th~us, the need for increased CO partial pressures to stabilize the catalyst complex i9 reduced. Increa~ed CO partial pressuras result in increased sub~titution of the pho~phlne ligands by CO; i.e.~ rhodium catalyst complexes leading to reduced n/i rAtios. To produce products of high n/i ratios rhodiu~ complex~s containing only one CO per Rh are preferred. Thus in this case, the partial pre~sure of CO is preferably below 500 psi.

-22- 12~73 C~t~l~st Comple3es and Select~ve Feed Conversion3 Catalysts suitable for use in this hydroformylation process include transition metal carbonyl complexes preferably selected from ehe group of Fe, Co, Rh, Ir and Os. The more preferred transition metals are rhodium, cobalt, ruthenium and iridium. Rhodium and cobalt complexes are most preferred. A preferred group of catalysts consists of transition metal carbonyl hydrides. Some of the carbonyl llgands of these complexes may be replaced by ligands such as trivalent phosphorus, trivalent nitrogen, tr~org~noarsine and divalent Yulfur compounds. Trivalent phosphorus ligands, and particularly triorganophosphines and phosphite ester~ are preferred.
~ he preferred triorganophosphine ligands include substituted and unsubstituted triaryl phosphines, diaryl alkyl phosphines, dialkyl aryl phosphines and trialkyl phosphines. These phosphines may be partially or fully open Chain or cyclic, straight chain or branched. They may have various substituents, such as those disclosed in U.S. Patent 4,668,809 by Oswald et al.
In general, the stable but not directly active catalyst complexes of the pxesent invention are coordinatively saturated transition metal carbonyl hydrides. They include metal carbonyl cluster hydrides. In the case of Co, Rh and Ir they are preferably of the formula LpM(CO)qH
wherein L is a ligand, preferably P, N or As ligand, M is transition metal, p is O to 3 and q is 1 to 4, with the proviso ehat p + q - 4. These complexes lead to catalytically active coordinatively unsaturatea compounds via L and/or CO ligand dissociation ~ (C0)qH ~ ~ (C0))qH ~~ LpN(co)q-lH.
In the presence of the sulfur containing olefinic feeds of the pres~nt invention some of the CO and/or other ligands can be exchanged for appropriate sulfur ligands during hydroformylation.
~ preferred subgenus of complex catalysts consists of penta-coordinate trialkyl phosphine rhodiu~ carbonyl hydrides of the general fornula (R3p)xRh~co)yH
wherein ~ is a Cl to C30 unsubstituted or substituted alkyl; x is 2 or 3 and y is 1 or 2, with the proviso that x + y is 4. The alkyl groups can be the same or different; straight chain or ~yclic, substituted or unsubstituted. The trialkyl phosphine rhod~um carbonyl complex subgenus of ~r ',,~.

~Z~ /3 catalyst co~plexes show~ outstanding thermal stabillty in the presence of exc~ss erialkyl phosphine ligand even at low pressure. Thus, it can be advantageously employed at temperatures between 140-200C under pressures ranging from 100 to 1000 psi. Tri-n-alkyl phosphine complexes of this type can be employed for the selective hydroformylation of Type I olefins.
In general, phosphoru~ ligands of low steric demand, such as tri-n-alkyl phosphinss and n-alkyl diaryl diphenyl phosphines, can lead to high n/i product ratios derived fro~ Type I olsfins in rhodium cataly~ed hydrofor~lation. this requires a hlgh P/Rh ratio in the catalyst system and a low partial pressure of C0.
Trialkyl phosphine complexes having branching on their ~-or/and ~- carbons have increased steric demand. They tend to form catalyst complexe~ of struceures which have lncreased reactivity toward Type II and Type III olefins. For example, the ~- branched tricyclohexyl phosphine and the p- branched tri-i-butyl phosphine ' CH2 - CH2 -CH CH2 and p 1c~2-cH-cH3i are attractive catalyst ligands of thia type. These catalysts, while highly active, do not provide high n/i product ratios.
Another pref~rred type of phosphorus lignnd for rhodium consists of alkyl diaryl phosphines of low steric demand. The tris-phosphine rhodiu~ carbonyl hydride complexes of these ligands show a desired co~bination of operational hydroiormylation catalyst stability and selectivity to producs hi~h n/i product ratios.
In general, tha hydrogenation activity of phosphine rhodium carbonyl complexes i3 relatively low. Thuc~ in the presence of these complexes, aldehyde protuces of hydroformylation can be produced in high s~lectivity without ~uch alcohol and/or paraffin formation, particularly at low temperaturo~.
Another subgenu3 of suitable catalyst complexes is that of p~ntacoordinate trialkyl phosphins cobalt carbonyl hydrides of the formula:
(R3P)u~o(c3~v~
wherein R is preferably a Cl to C3D alkyl as above, u is 1 or 2, v is 2 or 3 with the provi~o that u + v is 4. Tri-n-alkyl phosphine ligands are -24- ~2~ ,3 particularly advantageous in these cobalt phosphine catalysts since they provide high selectivity in the production of normal alcohol products when hydrofor~ylating the l-n-olefin and linear internal olefin components of the present cracked feeds. Tri-n-alkyl phosphine ligands include those wherein the n-alkyl substituents ~re part of a cyclic structure including the phosphorus, ~.g., CH3(CH2)20CH2 ~ P\

Using theqe cat21ysts, it ~s pref~rred to operate at high temperatures.
Thus, the preierable te~perature3 are beeween 160 and 200-C at pressures of 500 to 4500 p9i. The more preferabls pr2ssure range is irom 1000 to 3000 p9i. Low medium pressurss ranging from 1000 to 2000 ps~ are most pr~ferred.
Another subgenus of catalysts is reprssented by cobalt carbonyl complexe~ free from phosphorus ligands. These catalysts include dicobalt octacarbonyl and teeracarbonyl cobalt hydride.
Co2(CO)g and Co(C0)4H
The latter compound i9 assumQd to be an immed~ate precursor of catalytically active specie~. Cobalt carbonyl catalysts are stabillzed by bigh CO/H2 pressures ranging from 2000 to 6000 psi (136 to 408 atm) during hydroformylation. They are preferably used in th0 lO0 to 180~C temperature range. For a sel~ctive conversion of Type I olefins, lower temperatures up to 145C are used.
The above cobalt carbonyl complex can be generated by reacting cobalt or cobalt salt3 with C0 and H2. It is particularly adv~ntageous to e~ploy cobalt carboxylates as reactants for ths generation of cob~lt cArbonyl catalyst precursors.
When the cobalt catalyzed hydrofor~ylation ls completed, the cobalt carbonyl complex is con~ert2d into Co, i.e. metallic cobale or Co2~, e.g. cob~lt formate or acetate. The conv0rsion to cobalc acetate can be advantag~ously carri~d oue with hot aqueous aceeic acid and molecular oxygcn (air~. This allows the recovery of cobalt in the aqueous phase.
The cobalt acstate can then be conver~ed to an oil soluble higher molecular weight carboxylate and recovered. ~or more sxtensive description of the .

various method3 of cobalt recovery and recycle see pages 162 to 165 of the Falbe reference.
In the hi~h pres~ure cobalt ca~lyzed reaction of the present process uslng high sulfur feeds, dicobalt octacarbonyl is converted to partially sulfur ligand sub3tituted components as it is indicated by the following schemes.

C2(C)8 ~ ~ Co4tCO)7(sR)3 + C3(CO)6(S)SR
R2S~ ~
Co2tCO)7SR2 ~ [Co2(CO)sS~ + Co3(CO)gS

These and similar complexes and their hydride derivative~ form equilibria with dicobalt octacarbonyl and tetracarbonyl cobalt hydride. The resulting catalyst system provides active catalyst species with or without sulfur.
The sulfur containing species may also iead to insoluble and thus inactive CoS. The conditions of the present proce g, particularly the CO partial pressure, are set to suppress CoS for~ation.
In general, the transition metal complex hydroformylation catalysts of th~ present invention are employed in effective amounts to achieve the dasired olefin conversion to aldehyes and/or alcohols. The catalyst concentratlon is typically higher in the present process using feeds of high sulfur content than in other similar processes using pure olefin feeds. The transition metal concentration can range from 0.001 to 5X. The more preferred concentrations primarily depend on the ~etal employed. Cobalt concentrations range from 0.01 to 5~, preferably from O.01 to 5X, more pref~rably from O.05 to lX. Rhodium concentrations range fro~ about 0.001 to 0.5X. Other factors determining the optimwm catalyst concentration are the concentration and types of olefin in the feed and the desired olefin conversion. l-n-Olefins are generally t,he most reactive.
For a complete conversion of branched olefins, higher catalyst concentrations sre needed.
The phosphorus, nitrogen and ~rsenlc containing catalyst ligands are employed in excess. High excess llgand concentrations have a stabilizing effect on the catalyst complsx. Particularly in the case of the phosphoru3 llgands, it is preferred to employ a mlnimum of 3 to ligand to transition metal ratio. In the case of the phosphine rhodiu~
complexes, the mLnimu~ P/Rh ratio is preferably greater than 10. P/Rh -26- ~qa~$~

ratio~ can be as high a~ 1000. The sulfur-containing ligands may be provided in the feed.
The use of P-, N- and As-containing ligands, particularly phosphine ligands, leads to increased catalyst stability and selectivity for linear product formation. At the same time activity is usually decreased. Thus, the choice of metal to ligand ratio depends on the desired balance of catalyst stability, selectivity and activity. The S-containin~ ands can improve the aldehyde selectivity of the present process.

~ig~ P~su~e Low_Femp~at ~ Hvdrofo~mYl~tion The high pressure cobalt catalyzed hydroformylation in the absence of stabilizing added ligands such ~s phosphines is preferably carried out at low te~peratures below 180C where the reduct$on of aldehyde products to alcohols and the aldol dimerization of aldehydes during hydroformylation is reduced.
The aldehyde primary products are generally of a semilinear character. The linear n-aldehydes are the largest single aldehyde type pxesent in the products. The linearity of the alcohol products of hydrogenation is of course determlned by that of the parent aldehyde mixture. The linearity of the aldehyde products In turn is mainly dependent on the unique feed of the present process and the catalyst and conditions of the conversion. In the following the aldehyde product mixtures are further characterized particularly for the cobalt catalyzed hydroformylation.
Ihe maJor types of aldehydes are the n-aldehydes, the 2-~ethyl branched aldehydes and 3-methyl branched aldehydes. Much of the rest of the aldehyda~ are 2-ethyl or higher n-alkyl branched aldehydes. In general, the normal, the 2-methyl and 3-methyl branched products preferably represent mora than 40Z of the totai.
At the lowar temperatures, between 100 and 145C, the Type olefins, ma~or conponents of the present feeds, are not effectively Isomerized to the internal, Type II olefins of lesser reactivity. Thus a high concentration of the most reactive, terminal, Type I olefins is maintained. In addition, the low temperatures favor a higher n/i ratio of the hydroformylation products of Type I olefiDs:

RCH-CH2 . ! 2~ RCH2CH2CHO + RCHCHO

R-C3 to C33 alkyl n- i-(2-methyl) Thus, the use of low temperatures maximized the selectivity of the present process to the desired n-aldehyda and the 2-methyl substituted i-aldehyde products. For tha Type II, linear $nternal olafins, 2-methyl, 2-ethyl, 2-propyl, etc. substituted aldehyde~ are formed in decreasing concentrations as indicated by the following scheme (R - Cl to C31 alkyl):

RCH2CH2CH'CH2 ~ RCH2CH-CHCH3 , ~ RCH-CHCH2CH3 .1 \~ ~ \ 1~
R(CH2)4CHO R(CH2)2CHCH0 RCH2CHCHO RCHCHO

It was established by combined GC/MS studies that this product distribution of normal and 2-alkyl substituted i-aldehydes i9 a feature of the present process.
The 3-methyl substituted aldehydes are derived from 2-methyl-1-olefins which constitute most of the Type III olefin components of the feed. Some of the 2-methyl-1-olefins are isomeri2ed to internal, methyl-branched Type IV olefins and lead to other isomeric methyI branched aldehydes, a.g.

R'CH2CH2C-CH2 CO/H2 ~ R'CH2CH2CHCH2CHO
I R-Cl to C31 alkyl R'CH2CH-C-CH3 CO/H2 R'CH2-CH-CH(CH3)2 R'CH-CHCH-CH3 CO/H2 R'CH-CH2CH(CH3)2 C~3 C~O

~29~i3 The low temperature cobalt catalyzed process results in high selectivity to aldehydes having one carbon more than their olefin precursors. Little aldol addition of the aldehyde products occurs during such hydroformylations. Thus, the so-called dimer by-products, consisting mainly of aldol condensaticn products are minimal. Similarly, the amounts of trimars, largely consisting of acetals and products of the Tischenko reaction of aldol adducts, Is reduced.
A potential disadvantags of tha low temperature operation is the relatively low reactivity of the Types II and III and particularly the Type III olefins. This can be overcome in a staged operation which involves the hydroformylation of Type I olefins Ln the low t0mperature regime and the hydrofo~ylation of Type III olefinq in the high temperature regime, between 145 and 180C.
The low temperature operation can be effectively used for the selective conversion of Type I olefins to highly linear aldehydes. At low te~peratures, tha hydrogenation of the pri~ary aldehyde products to ~he corresponding secondary alcohol products is insignificant. Thus, the aldehydes can be separated and utilized as versatile chemical intermediates in various reactions.
Under the conditions of the present process, the desired hydrofor~ylation of the olefinic components of the feed occurs selectively witbout any significant conversion of the thiophenic aromatic sulfur compounds. ~le aliphatic sulfur compounds, particularly the thiol and disulfide components undergo a series of conversions, presu~ably via hydrogen sulfide. It was shown by sulfur specific gas chro~atography (S
GC) of the reaction mixturc using a nonpolar capillary GC column that most of the trace sulfur compound~ for~ed wera beyond the aldehyde product boiling range. It was found by GC/MS that these sulfur compounds were thiol esters and alkyl sulfides. Their alkyl groups had one carbon more than the olefin reactants. This indicated that they were probably derived from the aldehyde products via the following reactions with thiol and H~S
respectively.

2 RCHO ~ RSH ~ RCOSR + RCH20H
2 RCHO + H2S ~ 2H2 - (R~H2)2S

The hydroformylation reaction mixtures did undergo fur~her reactions on prolonged standlng. This resulted in the formation of si~nificant amounts oi` higher boiling sulfur compounds, Lncluding some boiling in the aldehyde range. To obtain aldehydes and derivatives oi low sulfur content, it is preferred to distill the reaction mixture without much delay after cobalt removal.

~vdro~or~ylation-Acetalizatio~ in $ho i~#uYq__~of Cobalt Low temperature, high pressure, cobalt catalyzed hydroformylation can be advantageously carried out in the presence of added Cl to C6 monoalcohols, diols or triols such as methanol, ethanol, 1,6-hexanediol, glycerol. In the presence of these lower alcohols, preferably employed in excess, ths aldehyde products of hydroformylation undergo diacetal formation catalyzed by cobalt carbonyl complexes. Using higher molecular weight alcohols, higher boiling acetals are formed. After the removal of the cobalt catalyst, these are readily separated from the reacted components of the cracked distillate feed by fractional distillation.
Thereafter, the acetals are hydrogenated in the presence of added water to produce the corresponding alcohols as indicated by the general reaction scheme:
RCHO R _H RCH(OR')2 H2 ~ R'CH20H ~ 2R'OH

The added lower alcohols form water soluble cobalt complexes and thus also facilitate the removal of the cobalt catalyst after such combined hydroformylation acetalization reactions.
In an alternate sequence of operation, the hydroformylaeion can be carried out in the absence of added alcohol or in the presence of less than stoichiometric amounts to minimize reactor volume. Additional amounts of alcohol are then added to the reaction mixture after hydroformylation eo compl~te the acetalization.
The use of added alcohols increases the stabllity of the catalyst syste~ and the reaction raee. Due to rapid acetal formatlon, other secondary reactions of the aldehyde reaction products such as aldoli2ation are suppressed. Another ma~or advanta~e of producin~ the acetal derivatives is their ease of separation. In contrast to ehe aldehydes, which aldolize on heating durin~ distillation, the acetals of the present invent~on can be separat~d without any significant yield loss by fractional distillation.

2~73 The hydroformylation-acetalization process of the present invention comprises reacting the previously described feed at first with C0 and H2 under hydroformylation conditions as described. The aldehyde products are then reacted with a Cl to C6 alcohol at temperatures between 15 and 250C, pressure~ between 15 and 250C and pressures between 0 and 5000 psig durin~ and/or after 3aid hydroformylation. If the acetalization is carried out or completed after hydroformylation, the conditions are milder, preferably ranging from ambient temperature to 100C at atmospheric pressure.

Hyd~f~L~LL~ion~- HYt~o~na~lo~
The aldehyde and aldehyde plus alcohol products of hydroformyl~tion are usually reduced to alcohols substantLally free from aldehydes by hydrogenation. The hydrogenation catalysts are preferably sulfur resistant heterogeneous composltlons based on Group VIII metals, particularly cobalt, molybdenum, nickel and tungsten. Cobalt sulfide and molybdenum sulfide are specifically preferred. They are preferably employed in the liquid phase at temperatures between about 50 and 250C, preferably, 120 to 220C, and pressures in the range of 50 and 6000 psi (3.4 and 408 atm), preferably 300 and 4000 p9i t204 and 272 atm).
The hydrogenation of the aldehyde mixtures of the present invention can be advantageously carrled out at variable temperatures, wherein the n-aldehydos are reduced to alcohols at first at lower temperatures thsn those needed for i-aldehydes. The n-aldehyde components are highly reactlve and sub~ect to conversion at high temperature to low value n-paraffin by-product~ and aldol condensation-hydrogenatlon products.
The 2-alkyl branched aldehydes require higher temperatures for their reduction to the de3ired aldehydes but have less tendency for paraffin and aldol by-product formation. Thus, a preferred selecti~e hydrogenation process for the present aldehydes in the presence of a CoS/MoS2 based catalyst comprises hydrogenation of mo3t of the n-aldehyde components in the temperature range of 130 to 190C followed by the hydrogenation of the r~st of the aldehydes between 170 and 2205C. The temperatures employed of course will also depend on ths catalysts used and the reaction time. Since most hydrogenations are carried out in a contlnuous manner, liquid hourly space velocities are another important factor in hydrogenation.
To increase the yield of the deqired alcohol products, the hydrogenations are carried out in the presence of mlnor a~ounts of water, preferably 1 to lOX based on the aldehyde reactant. The upper level of the water i3 limited by the sensitlvity of the catalyst. The water suppresses the formation of the aldehyde dimers during hydrogenation and facilitates the conversion of the dimer, trimer and formate by-products of hydroformylation to alcohols.
The hydrogenation of the present aldehydic feeds is preferably carried out under conditions not affecting the aromatic sulfur compounds, thiophenes and benzothiophenes. In a preferred operation, the cobalt hydrofornylation catalyst is removed and the cobalt free hydroformylation mixture is distilled to separate the unreacted hydrocarbons and aromatic sulfur compounds. The resulting aldehyde distillate or aldehyde distillation residue is then hydrogenated.
It was surprisingly found by sulfur specific GC analyses of the reaction mixtures that most of the sulfur compound components of the aldehyde boiling range are converted during hydrogenation to less volatile derivatives of the aldehyde dimar derivative range. Thus, essentially sulfur free alcohols could be obtained by a subsequent fractional distillation.
Dependent on the sulfur content of the aldehyde products, catalysts of varying sulfur sensitivity can be used. Such catalyst compositions include CuO and ZnO reduced by H2 or CO. For the reduction of the low carbon number Cs to Clo aldehydes, a vapor phase rather than liquid phase hydrogenation process can be used.

Contl~uous H~drofor~ylation The preferred mode of operating the present process is obviously continuoui rnthqr than batchwise. The reaction conditions of continuous and batchwise operation are nevertheless slmilar. Continuous hydroformylation can be carried out in a single reactor or in a saries of reactors using various method~ of ~eparating the cataly3t from the products and unreacted feed components. S~irred, pncked and plug flow reactors can be employed. Re~c~ants are continuously introduced. ~
When added stabilizing ligands (such as non-volatile phospbines) are used, ths products and unreacted feed may be separated from the catalyst qystem by flash distillatIon. In low pressure hydroformylation, direct product flash-off from the reactlon vessel can be employed. At increased pressures, a recirculation flash-off mode of operation is preferred. This latter ~athod would includ~ a continuous removal of liquid reaction mixture from the reactor. This liquid is then depressurized and flash distilled at atmospheric pressure or in vacuo. The residual solution of the catalyst may then be continuously returned to the reactor.
Stabilizing ligands of hydrophilic character may also be employed to make the transition metal complex water, rather than hydrocarbon, soluble. This allows b~phase ¢atalysis in a stirred water-hydrocarbon feed mixture and a subsequent separation and return of the aqueous catalyst solution to the reaction mixture.
In the absence of stabilizing ligands, the reaction mixture may be continuously withdrawn from the reactor and the transition metal carbonyl complex catalyst chemically converted to a water soluble, usually inactive form. After separa~ion of the aqueous solution, the transition metal compound is reconverted to the precur~o~ of the active catalyst which i5 then recycled to the reactor.
A vsriety of reactor schem0s can be used for the optimum conversion of ehe olefIn reactants in a continuous reactor. For instance, interconnected reactors may employ different catalyst systems. The first reactor may employ a phosphine-rhodium complex catalyst which selectively converts l-n-olefins and employs direct product flash-off. This might be connected to a second reactor containin~ a phosphine cobalt complex catalyst which converts the linear internal olefins via isomerization-hydroformylation. Aleernatively, cobalt alone may be used in the first reactor followed by a phosphine cobale complex.

~x~o~o~ at~n-~ tion A further variation of the present process is the aldolization of the product ald~hydes. A hydroformylation plus aldolization step in the pres~nce of a base followed by A hydrogenation step converts a Cn+2 olefin to C2n+6 aldehydss and alcohols. This i9 indicated in the following g~n-ral~ch-u- by th- exa~pl-s of ~yp- I o~-f~=-.

.

2 CnH2nllcH-cH2 / 2 ~ 2 CnH2n+lCH2cH2-cHO
n-al 1 Base CnH2n+lCH2CH2CH-C-cHO ~ CnHzn~lCH2CH2CH2CH-CH-CHO
CH2CnH2n~l OH CH2CnH2n+1 n,n-enal n,n-hydroxyanal ¦ H2 CnH2n+lcH2cH2cH2-cH-cHo 2 CnH2ncH2cH2cH2-cH-cH2-oH
CH2cnH2n+l CH2CnH2n~1 n,n-anal n,n-anol wherein the slmpla n-aldehyde product of hydroformylation is "n-al", the therm~lly unstable primary product of aldolization is "n-hydroxyanal", the unsaturated aldehyde resulting from aldolization is "n,n-enal", the selectively hydrogenated saturated alcohol is "n,n-anal~ and the final hydrogenated saturated alcohol is "n,n-anoln. The n,n-prefixes indicate that both segm~nts of the aldol compounds are derived from the terminal, i.e., normal, product of hydrogenation.
The hydrogenated saturatsd alcohol products of hydroformylation can be also derived by ths Guerbet reaction of tho alcohols produced from the primary aldehyde products of hydroformylation, ~.g.

CnH2n+lCH2CHO ~ CnH2n~lCH2CH20H

CnH2n+lcH2cN2cHcH20H
nH2n+l ~ I The Guerbet reaction is also a base and metal cataly~ed conversion It is carried out at elevated temperatures concurrent with the removal of the water condensation product.
~ ~ Minor iso-aldehyde components of the aldehyde product mixture can also be converted in a so-called crosc-aldolization reaction with che normal aldahyde:

.

CnH2n+lCHCH0 + CnH2n+lcH2cH2cHo i-al n-al r Base CnH2n~lcHcH-c-cHO
CH3 C~2cnH2n+
i, n-enal H2~
CnH2n+lcHcH2-cH-cHO
CH3 CH2cnH2n*
i,n-anal The rate of the above cross-aldolization process is slower than that of the simple aldoliza~ion. However, the relative rate of cross-aldolization increases with increasing temperature and decreasing n/i aldehyde ratios.
The latter can be achieved by the addition of extra i-aldehyde to the reaction mlxture.
The aldolization step can be carried out separately by condensing tha aldehyde product intsrmediates in the presence of a base catalyst.
Hydroformylation and aldolization plus hydrogenation r.an be combined by carrying out the hydroformylation in the presence of ~he above-described transition metal complex based catalysts plus a base aldolization catalyst.
A preferred mode of combined hydroformylation-aldolization is carried out in the presence of a trialkyl phosphine rhodium carbonyl hydride plus excess trialkyl phosphine hydroformylation catalyst system plus a base aldalization catalyst such as potassium hydroxide.
To carry out tho present combined hydroformylation-aldolization proca s in the preferred homogeneous, liquid phase, solvent selection is important. The preferred solvent will dissolYe all the widsly different components of the reaction system. Solvency for the nonpolar olefin reactant and polar caustic catalyst and water by-product is therefore a compromiSe. Alcohols, partlcularly hydrecarbyIoxyethyl alcohols are excellent choices. They may be of the formula J(oc~2c~{2)~oH

~2~4~ ~3 whsrein J - Cl to C4 alkyl, preferably primary alkyl, most preferably methyl, C6 to Clo substituted or unsubstituted phenyl, preferably phenyl, j is 1 to 8, preferably 3 to 8. Desirable solvents include methoxytriglycol, CH3(0CH2CH2)30H, and phenoxyethanol, PhOC~2H20H. In general, the weight proportion of the relatively nonpolar hydrocarbyl segment J to that of the highly polar oligo (-oxyethyl) alcohol segment determines the relative solvent power for the nonpolar versus polar co~ponents of the reaction mixture. As such, this type of a solvent can be readily optimized for any special application of the present proces3.
In ~ cont~nuous combined hydroformylation-aldolization process, product flash-off is more difficult to real~ze because of the high boiling points of the aldol condensation products. Therefore, direct produce flash-off is not generally feasible. Rec~rculation flash-off, aqueous catalyst separation and chemical catalyst recovery are preferred. Due to the high boiling point of the aldol condeDsation products, separation from the unreacted components of the distillate feed by fractional distillation is facilitated. Thus, broader carbon range distillate feeds can provide reaction mixtures suitable for aldol aldehyde or aldol alcohol separation by fractional distillation.
Since high aldolization rates can be readily achieved in the combined process, the reaction parameters can be readily adJusted to provide either ths unsaturated or saturated aldehydes as the major products. Short reaction times, and low olefin conversions, preferably below 50%, plus high base concentration, favor the unsaturated aldehyde.
However, mostly the saturated aldol condensation product is desired. This is, of course, the favored high conversion product.
Due to the improved thermal stabililty of the present trialkyl phosphine rhodlu~ complex hydroformylation catalyst, the aldol condensation products can be flashed off or distllled without affecting the catalyst.
However, strong bases have an adverse effect on the thermal stability of the system. These can be either removed before distillation or replacsd with weaker base aldolization cat lystq such as amines and Schiff bases.
For example, basic ion exchange resins can be filtered off. For known, applicable aldolization catalysts, reference is made to Volume 16, Chapter 1 of the monograph ~Organ~c Reactions~, edited by A. C. Cope et al., published by J. Wiley ~ Sons, Inc., New York, N.Y., 1968.

-36- ~ 73 The preierred concentration of the strong organlc base, i.e , alkali hydroxide, aldolization catalyst is low, between about 0.~1 and lZ, preferably between 0.05 and 0.5~. Of course, small caustic concentrations have less adverse effect on the stability of the reaction system.

Aldehxde p~od~ an~ Per$v~ti~es The present hydroformylation process, particularly the high pressure cobalt catalyzed reaction, leads to unique semilinear mixtures of aldehydes. Due to the specific mixture of 012fins found in the hydroformylation feed, it is now possible to obtain a mixture of aldehydes which cannot be economically produced in any other way. The aldehyde products of the present invention are versat~ile chemical intermediates.
They can be readily converted to alcohols, acetals, carboxylic acids and amines. The properties of thess compounds and of their ester plasticizer and ethoxylated surfactant derivatives are distinct and desired. They reflect the semilinaar character of their aldehyde precursors.
The semilinear aldehyde compositions have less than one branch per molecule. They have preferably Cs to C21, more preferably C7 to C21, most preferably Cg to Clg carbon atoms per molecule. They comprise 15 to 50X by weight of normal aldehyde which is preferably their major constituent. Other significant components are 3 to 20X of 3-methyl branched aldehyde and 3 to 20X of 2-methyl branched aldehydes. These components constitute preferably more than 40X, more preferably more than 50X of the eotal. The higher semilinear C7 to C21 aldehydes preferably also contain ~ to 20X of 2-ethyl and higher n-alkyl branched components.
The mixtures of semilinear Cs to Cls aldehydes possess alkyl moieties which make them suitable intermediates for the preparation of ester plasticizers having advantageous low temperature` properties.
Similarly mixtures of the semilinear Clo to C21 aldehydes have alkyl moietiea which makes ehem sultable intermediates for surfactants having appropriate blodegradability.
The reactions leading eo the formatlon of the present aldehyde mixtures were previously described. The struceural for~ulas and perc~ntag~s of ths key aldehyde constituent~ are shown by ~he following tabulation:

CH3(cH2)ncHo CH3(CH2)mCHCHO CH3(CH2)pCHC112CHO

15 to 50X3 to 20X 3 to 20X
n ~ 3-19m ~ 1-17 p - 0-16 CH3(CH2~pCHCH0 and CH3(CH2)qCHCH0 C2H5 ( CH2 ) rCH3 p - 0-16 q + r - 6 - 21 q ~ 2 ; r ~ 2 3 to 20%

An exemplary aldehyde mixture is a semilinear isomeric C
aldehyde having less than one branch per molecule and comprising 15 to 50X
of normal undecanal, 3 to 20X of 3-methylundecanal and 3 to 20Z of 2-methylundecanal, said Cll aldehydes togother constituting 40Z or more of the total. Another exemplary composition iq a semilinear isomeric C13 aldehyde having less than one branch per molecule and comprising 15 to 50X
of normal tridecanal, 3 to 20Z of 3-methyl-dodecanal, and 3 to 20Z of 2-methyldodecanal, said C13 aldehydes together constituting 40X or more of the total. Percentages are by weight.
In spite of the high sulfur content of their olefinic feed precursors, the present aldehyde mixtures are preferably of low sulfur content. They have less than 1000 pp~, more preferably less than 200 ppm sulfur. ~istilled aldehyde mixtures of narrow boiling range, containing mostly isomeric aldehydes of the same carbon nu~ber are preferred low sulfur compositions.
A preferret type of derivatives of the present aldehyde mixtures are the correspond~ng primary alcohol mixtures. They comprise semilinear Cs to C21 alcohol mixtures having less than one branch per molecule and comprising 15 ~o 50X of normal alcohol, 3 to 20X of 3-methyl branched alcohol and 3 to 20X of 2-methyl branched alcohol. The C7 to C21 alcohols preferably also contain 3 to 20X 2 eehyl and higher 2-alkyl branched alcohols. Thesa alcohol constituents and thsir percentages by weight are defined by formulas oi the following tabulation:

-38- ~2~ 73 CH3(CH2)nCH20H CH3(CH~)mCHCH20H CH3(CH2)pCHCH2CH20H

15 to 50~ 3 to 20X3 to 20X
n - 3-19 m ~ 1-17p - 0-16 CH3(CH2)pCHcH2~H and CH3(CH2)qcHcH20~
C2H5 ( CH2 ) rCH3 p - 0-16 q ~ r 6 ~ 21 q 2 2 ; r 2 2 v 3 to 20X

A The preferred subgroups of these alcohol mixture~ are the same as those of their aldahyde precursors. The above 3 eypes of components prefsrably constitute more than 40X, preferably more than 50X of the total.
The semilinear Cs to Cls primary alcohol mlxtures provide ester plasticizers with advantageous low temperatur0 properties. Similarly, the Clo to C21 alcohols are intermsdiates for biodegradable surfactants.
An exemplary alcohol mixture is an isomeric primary Cg alcohol having less than one branch per molecule comprising 15 to 60% of normal nonanol, 3 to 20X of 3-msthyloctanol, and 3 to 20X 2-methyloctanol said Cg alcohols constituting 40X or more of the total alkyl groups. Similarly, a mixture of isomeric primary C7 alcohols has less than one alkyl branch per ~oleculs and comprises 15 to 60Z of normal heptanol, 3 to 20~ of 3-methylhexanol and 3 to 20X 2-methylhexanol. Said C7 alcohols also constituting 40X or mora of the total.
The plasticizer e~ters based on the present alcohols are neutral alkyl esters of mono-, di- and tribasic carboxylic acids and phosphorus acid~ such as phosphoric, phosphorus and phosphonic acids. On an avsrage their alkyl groups have less than one alkyl branch and compxise 15 to 50Z
of normal alkyl, 3 to 20Z 3-methyl branched alkyl and 3 to 20X 2-methyl branched alkyl group~ and together they pref~rably repreqent more than 40X
of the total.
Exemplary and preferred types of the prssent plastici~er compositions are alkyl b~nzoates, dlal~yl phchalates, dialkyl adipates, trialkyl trimellitates, trialkyl phosphates, trialkyl phosphites, dialkyl benzenephosphonates.

39 ~z~5~3 Th9 moqt preferred plasticizer ester derivatives of the present alcohols are the dialkyl phthalate ester~. They are prepared by reacting the Cs to Cls alcohol mixtures with phthalic anhydride according to known methods. The two alkyl groups of these esters each have an average of less than one alkyl branch and comprise 15 to 50X normal alkyl, 3 to 20Z
3-methyl branched alkyl, 3 to 20X 2-methyl branched alkyl moieties.
Tog~ther they preferably represent 40X or more of the total.
A preferred exemplary phthalate ester of the present invention is ditridecyl phthalate having tridacyl groups with an averags of less than one alkyl branch and co~prlsing 15 to 50X normal tridecyl 3 to 20X
3-methyldodecyl and 3 to 20X 2-mathyldodecyl groups, said tridecyl groups together rspresenting 40X or more of ~he total.
The plast~cizer esters of the semilinear alcohols of this invention may be employed to plasticize thermoplastic resins, sspecially the vinyl resins. Suitable resins includ~ PVC resins derived from vinyl chloride monomer as well as copolymer~ of vinyl chloride and other mono-and di- olefinically unsaturated mono~ers copolymerizable therewith. The plastici~ers may also be used in con~unction with other polymers or mixtures thereof including, for example, polyvinyl alcohol, polyvinyl acetate, polyvinyl butyral, polyvinylidene chloride, polyethyl acrylate, polymethyl acrylate and poly~ethyl methacrylate. Preferred are vinyl halides such as polyvinyl chloride and copolymers of vinyl halides such as those containing at least 70 wt. ~ vinyl halide, e.g., vinyl chloride. The plastici~ers are employed in effective pla~ticizing amounts and generally from about 1 to 200 pArts of plastici2er per hundred pArts oi resin by weight (phr) and preferably 10 to 100 phr. Plasticized resins containing the esters of this invention exhibit excellent low temperature flexibility, high te2p~rature stability and reduced volatility.
Some of the esters of monocarboxylic acids, especially the acetic acid acid esters of the present semilinear C6 to C12 alcohols are also useful as solvents. The alkyl groups of these ester3 also possess less than one branch per molecule and co~prise 15 to 50X nor~l alkyl, 3 to 20 3-methyl branched alkyl and 3 to 20X 2-~athyl branched alkyl groups.
The semilinear C8 to C21 primary alcohols of the present invention are attract~ve intsr~ed$ates for e~hoxylated and/or propoxylated nonionic ~uriactants. Sulfated or sulfonated surfa tants derived from either the present alcohols or from thelr ethoxylated and/or propoxylated derivatives are of an anionic character. The preferred cationic surfactant ?73 derivatives af these alcohols, are primary, secondary and tsrtiary amines, ethoxylated and/or propoxylated tertiary amines and their quaternary ammonium deriva~ives, especially in their ammonium salt form. The semilinear alkyl moiety of the alcohol pr~cursors advantageously affects the biodegradability of all three classes of surfactants. Besides the hydrophilic-lipophilic balance, the properties of nonionic, anionic and cationic surfactant mixtures of the present invention depend on the presence of semilinear Clo to C21 isomeric primary alkyl groups derived from the present alcohols The nonionic, anionic and cationic surfartant derivatives of the present semilinear alcohols are derived via known methods, Their derivation is exemplified by the following reaction schemes wherein the symbol of the C8 to C21 alcohol reactan~s is RCH2OH.

RCH20CH2CH2S03Na _ RCH20il `~ - RCH20S03Na RCH20H ~ n CH2-CH2 - ~ RCH2(OCH2CH2)nOH n - 1-30 ~t/ ~
RCH2(0CH2CH2)nS03Na RCH2(0CH2CH2)nOS03Na RCH20H + NH3 ~ ~D RCH2NH2 + (RcH2)2N~l + (RCH2)3N

x+y CHz-CH2 ,(cH2cH2o)xH
RC~2N
(cH2cH2o)yH

RCH20H ~ CH2-CHCN -~ RCH20CH2CH2CH2NH2 x + y - 2-30¦~ x + y CH2-CH2 ,(CH2CH20)XH
RCH20CH2CH2CH2N ~
(CH2CH20)yH

As indicatad by the product formulas oi` the above scheme, the .preferred semilinear surfacta~ts are selected from the group of nonionic suriactants consisting of ethoxylated andjor propoxylated alcohols; the ~2~4~73 group of anionic surfPctants consisting of alkyl sulfates, ethoxylated and/or propoxylated alkyl sulfates or alkanesulfonates; the group of cationic surfactants consisting of alkylamines, ethoxylated and/or propoxylated alkylamines, ethoxylated and/or propoxylated alkyloxypropyl amines and quaternary salts of said alkylamines and alkyloxypropyl amines, wherein the isomeric Ca to C21, alkyl groups of said surfactants each have on an average less than one branch and comprise 15 to 50X normal alkyl, 3 to 20~ 3-methylalkyl, 3 to 20X 2-methylalkyl and 3 to 20Z 2-ethyl and highar n-alkyl groups together representing more than 50Z of the total.
These compounds preferably do not contain any completely substituted, i e.
quaternary carbon.
A preferred subclass of the present surfactants is that of the ethoxylated higher C8 to C21, preferably higher C12 to C16 alcohols wherein the alkyl groups are semilinear and defined as above and the ethoxylated moiety contains from 1 to 30 ethoxy units. These ethoxylated semilinear alcohols compare well with the corresponding ethoxylated branched and linear alcohols. They are better wettin~ agents than the linear derivatives. From the practical point of view, their biodegradability is of the same order as that of the more expensive linear compounds.
As specifically preferred, nonionic surfactant is a semilinear, isomeric ethoxylated tridecyl alcohol containing from 1 to 30 ethoxy units wherein the isomeric tridecyl groups are defined as above.
The semilinear C8 to C21 aldehydes of the present invention can be also advantageously used for the preparation of surfactants. Carboxylic acid surfactants of the anionic type can be produced by the oxidation of these aldehydes or their aldol aldehyde derivatives by molecular oxygen in the presence of a base. For exampla, with the normal aldehyde components, the following conversions are carried out:
.

2 ¦ NaOH2 ~NaoH
RCH2C02NaRCH2CH-CC02Na R

Cationic surfactants can be also derived from the semilinear aldehydes via reductive amination.

RCH0 NH3~ RCH2NH2 ~ (RcH2)2NH

Amines can be also produced directly from the thermally cracked definic stre~ms via hydroamination in the pre~nce of rhodium complex catalysts, e.g., RCH--CH2 ~ RCH2CH2CH2NR 2 HNR'2 Wherein R' is Cl to C8 alkyl and substituted alkyl such as 2-hydroxyethyl.

EW~
In the following, examples are provided to illustrate the claimed hydroformylation process, but not to limit the invention. Prior to the examples the cracked distillate feedstocks are described. The description of the feedstocks details the structural types and amounts of reactive olefins present, this information being a key component of the invention.
Thereafter, the low and high pre3sure hydroformylation procedures used and the product workup are outlined. Thsn the examples of the actual hydroformylation experiments are given in groups according to the feeds and catalysts employed. The summarized reqults of these experi~ents are also provided in tables.
The cobalt catalyzed high pressure hydroformylation of cracked distillate fractions is described in particular detail. The semilinear aldehyde product~ of varying carbon number are characterized. Their hydrogenation to the corresponding alcohols is also outlined. Finally, the conversion of the alcohols to phthalate ester plasticizers and ethoxylate surfactants ic discussed. Some comparative data on plasticizer and surfactant properties are also provided.

Feedstock~
The feedstocks used in the following examples were fractions of liquid distillates produced by Fluid-coking and Flexi-coklng in the temperature range of 482 to 538C (90Q to 1000F). As hi8h temperature 43 ~2~73 thermal cracking processes, Fluid-coking and Flexi-coking produce distillate llquids and residual coke from vacuum residua. In Fluid-coking only the distillate products are utilized. The vacuum residue feeds and the thermal cracking step of Fluid-coking and Flexicoking are identical.
However, the Flexicoking process is further integrated into the refinery by virtue of using the coke to manufaceure low thermal value gas. Flexicoking is disclosed in U.S. Patents 2,905,629; ~,905,733 and 2,813,916 which were previously discussed. Flexicokin~ ls described ln U. S. Patents 3,661,543;
3,816,084; 4,055,484 and 4,497,705.
The key factor in producing the present highly olefinic feed is the high temperature ther~al cracking. However, ~nother important factor i9 the origin and prior treatment of the petroleum residua to be cracked.
Th~ presence is desired, major l-n-olefin components of the present feed depend on the presence of n-alkyl groups in the feed. These olefins are formed by the cracking and dehydroganation of n-alkyl aromatics and paraffins. In the past the molecular structure of higher boiling coker distillates was not k~ow~. Thus the desired feeds of the present invention were not recognized.
The Fluid-coker distillate feeds were derived from a Northwesc American crude. The Flexicoker distilla~es were produced from mixed crudes of Southwest American and Mideastern origin. Their compositions and those of o~her cracked distillates of different origins were remarkably similar.
An important step of the present inventlon was the structural analysis and recognition of the preferred distillate feeds. Since these feeds are extraortinar~ly complex, several analytical techniques were employed. The feed~ were analyzed using pac~ed column and capillary gas chromatographs (GC). The capillary GC was equipped with 50m or 30m fused silica columns coaeed with methyl silicones to determine the individual components. The sulfur compound components ware also analyzed by capillary GC, using a dual detection system. The colu~n effluent was equally divided and directed to a flame ionization detector (FID) and sulfur specific detector. Sulfur was detected either by a HallTM Electrolytic Conductivity Detector ~iving a linear response to sulfur or a Hewlett-Packard Flame Photometric Detector with a close to square dependenca on sulfur concentration.
A high resolution, 40Q MHz, proton resonance spectrometer (NMR) was used to estimate the various types of hydrocarbons, particularly olefins.

;~

37~

The structures of key feed components and products were determined by combined gas chromatography/msss spectrometry, GC/MS A
Finnigan TSQ-46B triple stage quadrupole GC/MS/MS was used in a single stage mode. Both electron impact ionlzation (EI) and chemical ionization (CI) were used for the identification of the components. EI provided informatlon on the structura of the molecular frag~ents. It was particularly successful ln determining the structure of the 2-alkyl branched aldehydes bassd on the fragments resulting from the McLafferty rearrangement. CI, using a~monia and deuterated ammonia as reagent gases, was used in determining the molecu~ar weight and compound class of components.
Th2 sulfur containing ions ~ere recognized on the basis of the appearance of assoc~ated isotopic peak3. The natural abundance of the 34S
isotope is about 4X of the 32S isotope. Therefore, besides the peak for the 32S fragmene, an appropriate weaker peak having a higher m/z value by 2 is exhibited for the isotopic 34S moiety.
Elemental and group analysis techniques were used to determine total sulfur, mercaptan sulfur and total nitrogen contents.

Co~er ~E~L~h~
The compo3ition of several coker naphtha distillates was analyzed by capillary GC, using temperature programmed 30 and 50m columns. They key componants of the mixture were identified by GC/MS with the help of standards as required.
The capillary gas chromatograms of Fig~Q_l were obtained using a 30m coluwn with FID and S detectors to show the distribution of hydrocarbon and sulfur compounds in a Flexicoker naphtha.
The GC of they hydrocarbons ~and organic compounds in general) in the bottom of the figure shows that the largest s~ngle types of components in the C6 to Clo range are the l-n-olefins (Cn) followed by the n-paraffins (Cn). This ratio is about 1.3. This ratio is very sensitive to the cracking conditions, particularly temperature. Among the aro~atic compounds, toluene, xylene and trimethyl-b~n7enes are ehe main components in ehis carbon range.
The upper sulfur specific chromatogram shows that the major sulfur compounds prssent were aromaeic: thiophene, mono- di- and trimethylth~ophsnes. The minor suliur compounds wero aliphatic thiols.

1~ 73 Figure 1 indicates that the GC retention times and the boiling point~ of the thiophenic sulfur compounds and those of the aromatic hydrocarbon co~ponents largely coincide. Both differ from the boiling range of the ma~or olefins present.
Thus, it is possible to separate highly olefinic C6, C7 and C8 distillate fractions essentially free from-aromatic sulfur compounds as it is shown by the shaded portions of the figure. The minor thiol components of these fractions can be removed by caustic wash or by converting them by oxidati~e methods to higher boiling compounds which can be then readily separated by distillation.
The hydrocarbon composition of the Fluid-coker naphtha was analyzed with a capillary GC equipped with a 50~ column which provided a higher resolution of the co~ponents. The l-n-olefins and n-olefins were again the main types of components in that order. The complete chromatogra~ is shown by Figure 1 of the parent application.
The corresponding l-n-olefin to n-paraffin ratios of the Fluid-coker naphtha are shown by Table I. In the C6 to Cl2 range chese ratios range from about 1.1 to 2.1. In general, the l-n-olefin to paraffin ratio increases with increasing carbon nu~ber.

l-n-Olefin Versus n-Paraffin Components of Fluid-Coker Naphtha Com~onent. GC~
Ratlo, Carbon l-n n- Olefia_ No. Ql~ Paraffin Paraffin 3 0.120 0.169 0.7101 --4 0.193 0.307 0.6287 0.418 0.523 0.7992 6 1.298 0.924 1.4048 7 1.807 1.496 1.2079 8 2.223 1.960 1.1342 9 2.164 1.651 1.3107 2.215 1.483 1.4936 11 1.534 0.989 1.5Sll 12 0.623 0.299~ 2.0836 3-12 12.295 9,801 1.2545 As summarized by Table I, in the C3 to C12 range, the naphtha contalned 12.3X l-n-olefins and 9.8X n-paraffins. Thus, the overall l-n-olefin to n-paraffin ratio was 1.25.
The ratio of l-n-olefins to n-paraffins is a main factor indicating whethsr or not a given thermally cracked distillate is suitable feed in the present process, particularly in the case of the cobalt based catalysts. The ratio should be above 1, praferably above 1.2.
Lower cracking temperatures result in decreased olefin/paraffin ratios. For example, delayed coking which is carried out at a lower temperature than Fluid-coking give~ distillates of lower ratios. An analysis of a naphtha fraction irom a delayed coker gave an average of 0.3 l-n-olefin/n-paraffin ratio as it is shown by Tab~e I~.

Table II
l-n-Olefin versu~ n-Paraffin Co~ponents of Delayed Coker Napheha Com~onent GC~ ~
Ratio, Carbon l-n n- Olefi~
No~ Olefin Paraffi~Par~in 6 1.9565.008 0.3850 7 2.3447.352 0.3188 8 1.8796.707 0.2802 9 1.4924.148 0.3596 0.3740.994 0.3763 6-10 8.04524.209 0.3323 A comparison of the olefin/paraffin ratios of Table I and Table II indicates that Fluid-coking provides an about 4 times greater olefin/parafiin ratio than delayed coking.
~ any of the other components of ~he naphtha were also identified.
Some of the illustrative detalls will be given in a discussion of certain distillate fractions.
The broad C13 to C12 coker naphth~ fraction was fractionally distilled, using a column equivalene to 15 theoretical plates wieh reflux rativ o f 10, to produce distillates rich in olefins and p~raffins of a particular carbon number. The bolling ranges and amounts of the distillate fractions obtained ondistilling th~ naphtha are shown by rab~es ~I and IV.
The l n-olefin and n-paraffin components and a few kay aromatic 4~ 3 cn tG
C`~ S o ~ In ~ N O C~
L S ~ O O O O O ~ O
C~ Z

C o o I r~ o o o - aJ ~
al O O
1: ~ ~ O ~ ~ _ r~ W ~ I~
~ o ~ O ~

~ o o Q G C In oo 01 l~ ~ C~l C.~ _ o _ ~ ~ ~ r~ ~ ~
c E
LL '- C~ ~oO~ r~ O
O ~ ~ ~ 0_~00 ~ ~ ~ o n C`J~
'P' ~ ~- ~ UO~ C~ ~
L~_ ~ ~ o~ ~ ¦ . . .
-- LL --I --' O
--~ t l -N~ , e ,~ O O T O~ O O L~) O .D ~ ~ ~
I_ ~t o ,s~ , o _ cn u) C~J E
o o O Z
(n .~ _ o O c~ I _ o c~ O
~ O C~ O O ~ ~ ~ O ~C
o E 00 ~0 ol _ O _ ~ O _ ~
~ S ~ I --I O-- --= 00 0 E

~ a~ ~ ~, ~ o~ O I ~ O. ~ . ~D o ~I 3 ~ U~ O c~
C~
~: _ I
.~ U~ I I I .--O r~
~ ~ o o U~
C~ ~ O ~ ~t _ ~ r~ O
__ C~l__ c .. ~.. , _ ~
z o~ o~ ~ E~ O
:n o ~D _ D~r~
Cl~ ~_ ~ ~ X _ r-~

'J

" 1 o ~ o N
L u I_~ ~I N ~ N CO O D
~ U

~ ~ 1~ 5~ ~D ~ N C~) ~ CJ
., _ O O . .. . . . L .
L~ O ~ . C~l O ~ ~ ~
O V~ Oæ
~3J t~ o ~ o a~ C O
._ U O~ ~ ~ ~ . . .. . . . _ 1~ ~ _ N _ t'~ O O CO 0~ O O O
O ~ ~ C

LL ' '`'~ c ~

2~ Q r~ C 1~ ol CO r~ ~ E ~ o m ' 0 3~
Cil O ~ 0 ~o u ~O ~ I~ - C .~, r~ _~ O ~1~0 I~J_ ',~ U~
a~
~e e U ~ ~ o l ~ o ~ ~
~3 ~Q ~11 r~ I o (13 C 'C o CV3 , el~ $ ~ O Oo~ 1~ N ~ _ 3 ~ L~ I C
I~ aJ-O
0 u~ oI ~ ~ t ~ ,u ,~, O _ N~ ~ C0 l~o N O v) ~_ 3 C _ I I IV 2 Q G
3J ~ ~/7 C U
O C ~1 I= o E IU ~ I" _ ~c~v C ~ llCo C 1~ 0 0 ~ _I ~ ~ CU
o . C_~ o ~ o C C~ ~
E a _ c_~ x _ o ~-- _ v ~ _ ~ z _ v -` ~2~4~ ~3 hydrocarbons present are also shown. The results indicate that in the Cs to Clo range, distillates containing about 15.1 to 29.6X of individual l-n-olefins could be produced. In the case of the higher boiling fractions, separation was more difficult and thus the maxlmum l-n-olefin percentage in the case of l-dodecene was 12.7X. The separa~ion of Clo, Cll and C12 fractions was adversely affected by the presence of water in the distillation vessel. This effect could be eliminated by removing the water in vacuo.
The C4 to C12 naphtha and selected distillate fractions thereof were also studied by proton NMR using a JEOL GX 400 ~Hz spectrometer.
Figure 2 shows the NMR spectrum of the olefinic region of the naphtha with an indication of the chemical shift regions assigned to the vinylic protons of various eypes of olefins. A quantitative dete~mination of the olefinic protons of the various types of olefins w8s used to estimate olefin linearity. The relative mole percentage~ of olefins of varying carbon number were calculated on the basis of amounts of the different types of olefinic protons. The results of these calculations are shown in Table V.
The data of Table V show that the Type I olefins, i.e., monosubstituted ethylenes, are the maJor type of olefins in all the distillate fractions as well as in the starting C4 to C12 n~phtha. The percentage of Type I olefins in the distillation residue is, however, reduced to less thsn half of the original. It is assumed that this result is due to l n-olefin conversion during the high temperature distillation.
Minor variations, between 32 and 50Z, are also observed in Type I olefin content of distlllate cuts. Ths reasons for thi.~ variation are unknown.
The only Type I olafins indic~ted in the C8 and higher carbon fractions are l-n-olefins.
The second largest olefin type present in the naphtha and its dlstillate consists of 1,2-diqubstituted ethylenes. The percentags oE
these Type II olefins varies between 18 and 26~. Most, if not all, of these olefins are linear internal olefins.
Type III olefins, i.e., l,l-disubstituted ethylenes were found to be present in amounts ranging from 12 to 17X. The ma~or olefins of this type were 2-methyl substituted terminal olefins. On the basis of MS
studies of aldehydes deri~ed from these olefins, it appears that their branching occurs mostly at the vinylic carbon.
Type IV olefins, i.e. trisubstituted ethylenes, were the smallest monoolefin components of these distillates. Their relative ~olar ~Z~ 7;~

~ o ._ ~ ~ ~

l~- .
C~J o ~
. ~ ~ ~ I ~ N
_~ ~ ol ~1 ~ ~ ~ ~D
~1 . O 1~) Cr ~ N ~
~ ~ o ~ n ~C~

_ e ~ ~ o ~ o N
_ ~- ~
:~O ~ O I C~ ~ D ~I~
~ O ~ ~_ ~ ~ J O

I_ ~ e O ~ ~ ~ ¦ ~ N
o ~ V O _ N ¦ e~ N . __ aJ ~ o o l o co u- o c~ Q
4 ~ c~

l --I~ ~'1 N _ e l ~ r~
~ a~

a!: OT ~ ~ O O
,~ ~~ ~ ~
C_~ ~-- ............ ~ .~
~__ _. ~ ~ ~ ~ C
,c C ~ ,C ~ 3 ~0~
S ~ ~ ~cV
Il~ O _ o 2 a~ O

37~ 1 concentration is in the 6 to 12X range. Interestlngly, the C8 fractions contained the least of these olefins among the fractions examined.
Type V olefins, i.e., tetrasubstituted ethylenes, could not be deter~ined by proton N~R. They are of little interest in the present invention since they are apparently unreactive in hydroformylation.
Finally, Table V also lists small but significant quantltiies (8 to 16Z) of con~ugated diolePins. The amounts listed for these olefins are approximate because conjugated olefins may have a different number of vinylic hydrogens per molecule dependent OD the site of conJugation and the presence of branching at vinylic sites.
The NMR spectra of naphtha fractions w~re also analyzed in the area of aro~atic and paraffinic protons to esti~te tho amounts of olefins.
Ta~le VI summarized the results. It shows the percentage distribution of various types of hydrogens. From this distribution and the elemental analyses of these fractions, th~ weight percentage of various types of compounds was ssti~ated.
The Type I olefins, mostly l-n olefins were estimated to be present in these fractions in the range of 18.7 to 28.3X. These percenta~es depend on both the carbon nu~ber and the particular usually narrow boilin~ range of the olefinic fractions studied. In the C6 to Clo range these values for the Type I olefins approximately correspond to the values obtained for l-n-olefin by GC.
The total olefin content of these fractions is in the 47 to 62X
range as deter~ined by NMR. It is noted that the con~ugated diolefins are included in this p~rcentage since they ~re converted to monoolefins under hydroforDylation conditlons or by a prior mild hydrogenation. The amounts of paraffins are g~nerally decreasing with increasing csrbon numbers while the amounts of the aromatics are generally increasing.
To illustrate the detailed composition of the presen~ naphtha feeds, more detalled data are pro~idsd on the C8 and Clo fractions on che basis of GC and GG~S analyses.
The compositlon of a heart cut C6 Fluid-coker distillate fraction is shown by Table VI~. This frac~ion was obtained by the redistillation using a 15 pla~ column and a lO to 1 rePlux ra~io (15/10) of a broad C6 cut It distlll~d between 56 and 65C (133 - 149F). Table VII shows the compo~tion of the broad cut ~eed and the heart cut product of distillation.

~$~3 C ~
, ._ ~o ~ I_ o o~ ~ _ o ~ ~ ~t ~ ~ ~ U~
l- - o e ~1 _ ~ ~ ~O ~ ~O c~.
O_ _ ~ ~ ~ r_ _ _ C~
I ~C I 00 r~
~ L 4--c~J ~J 'D ~ ~
e CL ~ ~ ."
a~
U~ r~ ~ _ ~ U7 _ 2 U7 0 I-J C ~11 ~r ~O
~IJ
.C I_ ' U ~0 ~ CO _ ~ ~
a e ~ ~ ~ o co ~? ~
U ~ ~ ~ o ~Z ' ~
~` ~

a~ d' O 1~ ~ r~
8-U ~ Ct) tr~ ~ I` t~ O C~J 3 :2 aL V O _ C~i ~ ~ ~ ~0 11 el~
-~o 2 ~ L ~ ~ . ~ r` O O~ O n _ e ~ c a~ _ ~ ~ ,~ r~ O ~ ~
o 3~ o ~ o, ~n " ~ O. _.
o~? ~
~ ~ .0 er ~ 00 CO ~ ~ e~O
3, ~ , ~ r~ _ ~ ~ _ ~~ ~
'- ~ C l _ o o o o o _ o o o aJ u~ C~l ~ ~ O G
,, ~ aJ--' I ~ C~J t~> O ~` C~J ~ C~l ,~ ~ -1 - -' - -C CO ~ I~ L~
~ ~ ~ ~ - - ~ - ~
I ~ a~
I ~O 1~ O

,c ~ ~ ~t '` ~ a~
_ ~:n ~ D r _ _ _ o ~q V ~O O~
a~ ~ o ~ o o~
0 ~ ,_ _ _ _ ~ ~ ~
O C L
~L ' ~O 1~ O = C~ J +
~ ~ et --~L ~ .

Table VII
CO~ ITS 0~ THE C6 DISTILI~TE CUT OF
A FLUID-COKRR N~P~THA BE~ORE AND AFT~ R~DISTILL~TION

Seq. Componçnt Boiling Point Componen~ GC Re-No, ~ erature)GC, Z tention Abbre- -F C Feed Heart Time _ Namç _ vint~,o~ a~ Cut Min 1 l-Pentene 112 44,2422.30 - 6.79 2 n-Pentane 97 36.0653.06 - 7.17 3 2-Methylbutene-2 10138.5683.30 0.10 7.94 4 Cyclopentene CP~ 11244.2422.33 0.37 9.40 4-Methylpentene-1 4MeP- 12953.8652.45 2.44 9.65 6 3-Methylpentene-1 3MeP~ 13054.1781.84 1.60 9.71 7 Cyclopentane CP 12149.2621.91 0.81 9.97 8 2,3-Dimathylbutane 2,3-Di~eB 13657.9880.17 0.27 10.15 9 2,3-Dimethylbutene-1 2,3-DiMeB~ 132 55.6160.68 0.78 10.20 cis-4-Methylp~ntene-2 c-4MeP~2 13356.3870.34 0.44 10.33 11 2-Methylpentane 2MeP 14160.2713.28 5.53 10.40 12 trans-4-Methylpentene-2 t-4MeP~2 13958.612 1.71 2.37 10.48 13 3-Methylpentane 3MeP 14663.2821.76 3.58 11.22 14a 2-Methylpentsn2-l 2MeP~ 14462;113 l-n-Hexçne n-H~ 146.5 63.48521.13 42.00 11.65 16 3-Methylcyclopentene 3MçCP-149 64.910.24 0.50 12.06 17 n-Hexane n-H 15668.73610.6113.71 12.38 18 cis-3-Hçxene c-3H- 15266.450l.S0 2.45 12.52 19 trans-3-Hexene t-3H- 15367.0883.15 4.25 12.69 trans-2-Hexene t-2H- 15467,8845.18 8,38 12.86 21 trans-3-Methylpentene-2 t-3MeP~2 lS970.438 1.12 1.46 13.04 22 4-Methylcyclopentene 4MeCP~ 150 65.67 1,OS 1.93 13.11 23 cis-2-Hexene c-2H- 15668.8911.64 1.64 13.31 24 2,3-Dimethylbutadiene 2,3-DiMeB - 154 67.78 0,16 0.20 13.46 cis-3-Methylpentene-2 c-2MeP~2 15467.7072.16 1.12 13.76 26 Methylcyclopentane MCP 16171.8124.26 2.04 14.14 27b l,l,l-Trichloroethane C13CCH3 165 74.10 - 0,05 15.02 28 Methylcyclopentatiene MeCP~- O.S9 0.12 15.02 29 l-Methylcyclop~ntena l-MeCP~ 168 75.49 0.54 15.79 Benzene Bz 17680.1008.38 1.26 15.92 31 Thiophene S 18484.16 0.10 0.06 16.22 Total Identified 86.94C99.46 a) 2-Methylpentene-l was not separat~d from l-n-hexene on the column used, I~ amounted to about 2X. b) Trichloroethane was present as an impurity as a result of using it as a solvçnt for cleaning the distillation apparatus, c) Thç feed cont~ined significant amounts of higher boiling components, ~Z~4~

Table VII shows that the largest components of both the feed and the heart cut were l-n-hexene (n-H~) and n-hexane (n-H). There were also significant amounts of linear internal hexenes (16.7Z) and methylpentenes ~9.4X plu8 2-methylpentene-1) in the heart cut. In addition, l.9Z of 4-methylcyclopentene and 0.5X of 3-methylcyclopentene wsre identified.
Thus the amount of linear internal and monobranched olefins is about 28.5X.
Only 0.8Z of a dibranched olefin, 2, 3-dimethylbutene was found.
The composieion of the heart cut is illustrated by the gas chromatogra~ oX Fi~u~e 3. The co~ponents are id0ntified by the sy~bols explained in Table VII. Figure 3 also shows ths chromatogram of unconverted hydrocarbons remaining a~ter tha hydroformylation of the heart cut. The m2~or unconverted components were n-hexane, methylpentanes and benzene as expected. The comparison of th~ chromatogra~s of the hydrocarbon components of the hydroformylation feed and the final reaction mixture greatly helped in the identificstion of ehe feed components.
Tab~ VI~ shows the composition of two C8 fractions of a Fluid-coker naphtha. It is apparent that beside the major l-octene component, there are significant quantities of all the linaar internal octene isomers. The trans-isomers of octene-2,-3, and 4 were identified.
2-Methylheptene-l was also identified as the largest single branched octene. Toluene, ethylbenzene and xylenes were al90 present.
One fraction is richer in l-n~octene, the other ln n-octane. The sum of identified olefins in the3e fra¢tion~ is 33.lX and 20.1%, respectivoly. Some of the octene isomers wera not identified. The first fraction richer in olefins was used as the feed in the C8 naphtha hydrofor~ylation exp~ri~nts.
Ths composition of C8 Flexicoker naphtha fractions was also studied in some detail. At first a broad C8 cut was obtained by a 15/10 fractional distillation between 110 and 135~C (230 - 275~C). Part of this broat cut was then redistillad with a 36 plate~ column using a reflux ratio 20 ~2h/20). Fractions of the 36/20 distillation boiling betwesn 117 and 124C (243 - 255F) were combined to provide a narrow cut in about 42X
yi~
~ a~ shows the composition oi th0 above broad and narrow Cg Flexicoker naphtha fractions. A comparison of the capillary GO data of Tables VIII and IX indicates that the composition of these Fluid-coker and Flexicoker naphthas is similar in spite of their different crude sources.
The narrow cut Flexicoker naphtha fe0d con~ains hi~her amountc of linear T~ble VIII
~Jor Olefln, Par~ffin and Arom~tic ~ydroc-rbon Components of Dlstill~t2 Fr~ctlons o~ Fluld Cokor N~phth- in th~ C8 ~ange Wei~ht % Co~QsLtion bv GC
Designation of Fraction l-Octene R~ch n-Oct~ne Ri~h Fraction No. 11 12 Quantity, g 2072 1034 ~oiling Point Range, ~F 245-254 254-262 OthersOl~flns Olefins Others X ~ X X

Toluene 4.3 1.3 2-Methylheptene-1 6.3 3.2 Oceene-l 18.5 10.3 trans-Octena-4 1.0 0.6 trans-Octen~-3 2.1 1.3 n-Octana19.9 16.3 trans-Octene-2 3.6 2.8 cis-Octene-2 1.6 1.8 Ethylbsnz~ne 0.6 6.1 m-Xylene 0.1 5.1 p-Xylene 1.8 o-Xylene 0.8 Nonena - 1 Su~ of Identi~ied Compounds 24.9 33.1 20.1 31.4 -56- ~ 2~73 Table IX

Re~ Co~pon~nts oi the CB Bayto~n Flc~icoker Naphtha Feeds Concentration of Components, %
Broad Narrow __Component Identificationa _ _ Bp. Bp.
Boi~in~ Po~lt 110-135C 117-124C
_ _Name C _~E~ 2~Q-27$F 243-255F
l-n-Heptene 94 201 0.14 n-Heptane 98 208 0.32 Methylcyclohexane 101 214 0.57 3-Methylcyclohexeneb~C 104 219 0.3~ 0.15 Toluened 111 232 6.39 O.lS
4-Methyl-l-Heptene 113 235 2.47 0.49 2-Methylheptane 117 243 2.94 3.67 6-Methyl-l-Hepteneb 1.38 1.53 1,3-cis-Dimethylcyclohexaneb 120 248 2.02 3.31 2-Methyl-l-Heptenee 118 244 4.08 7.ô2 l-n-Octene 121 250 11.07 28.12 4-Octene 0.97 2.62 3-Octene 123 253 0.98 3.09 n-Octane 126 259 9.98 20.01 trans-2-Octene 125 257 1.82 2.86 Dimethylhexadieneb 1.78 4.28 cis-2-Octene 126 259 1.25 1.76 Dimethylcyclohexeneb 1.72 1.22 Ethylbenzene 136 277 3.2.2 0.09 2,6-Dimethyl-l-Hepteneb 2.18 m,p-Xylenes 138 280 6.03 o-Xylane 144 291 1.02 l-n-Nonene 0.57 n-Nonane 151 304 0.23 , a)Identification based on GC, GC~MS and boLling polnt correlations.
b)The identification ls tentative.
C)l-Methyl-cyclohexene is also indicated.
d)2,4-Dimethyl-l-hexene is also indicated.
e)3 Methyl-l-heptene is also indicated.

octenes than the broad fraction (36.45 versus 16.09). Most significantly, the percentage of l-n-octene in the narrow cut i5 28.12X while it is only 11.07~ in the broad cut.
The broad cut naphtha is richer in branched olefins including C7 and Cg compounds of an open chain and branched character. In consrast to the narrow fraction, the broad cut had significant amounts of aromatic compounds: 6.39Z toluene, 3.22X ethylbenzene, 7.05X xylenes.
In the broad cut naphtha the pre~ence in small aMounts of a high number of monobranched olefins was indicated. The largest of these 2-methyl-l heptene is present in both the broad and narrow cuts in concentrations of 4.08 and 7.82X, respectively. There are also other methyl branched, mainly terminal, msthylheptenes present. However, the exact structures of these compounds are not known with certainty. In addition, there are branched, cyclic ole~ins present, particularly methylcyclohexene and dimethycyclohexene.
Fi~ure 4 illustrates the composition of the narro~ cut C8 Flexicoker naphtha. It is noted that most of the olefin components are linear or monobranched compounds. The cyclic olefins are largely excluded from this fractLon.
The sulfur content of the broad C8 fraction is lX while that of the narrow C8 fraction is 0.2X. The concentration of the main sulfur containing compounds, i.e. methylthiophenes and dimethylthiophenes is drastically cut in the narrow fraction. The distribution of the sulfur compounds in the two fractions is indicated by the sulfur specific gas chromato~rams of Flgu~e S. Althou~h the sulfur response of the detector is close to quadratic rather than being linear, the figure shows that the thiophenic sulfur was largely removed by fractionation from the narrow fraction.
~ xtraction of the narrow cut with 30X KOH solution in methanol containing 2X water reculted in a fuxther reduction of the sulfur content.
It was specifically shown by sulfur GC that the pentanethiol co~ponent was completely remo~ed.
Fi~u~e 6 illustrates the composition of the Clo naphtha fraction.
As it is indicated, besides the main l-n-decene component several of the linear decenes and 2-methyl nonene-l were identified. It was also shown that indene, a reactive, aromatic cycloolefin, is also present in this fraction. The main aromatic hydrocarbon components are trimethylbenzenes and indane.

~l~9'~ 3 The naphtha and its distillate fractions were also analyzed for sulfur and nitrogen compounds. Table X shows the carbon, hydrogen mercaptan and total sulfur plus total nitrogen contents.
The ~ercaptan content of the C8 and higher fractions is surprisingly low compared to the hlgh total sulfur content when determined by mercaptan tieration by silver nitrate. It is believed that this is in part due to the facile cooxidation of mercaptans and activated olefins.
The total sulfur content generally increased with the carbon number of the di tillates from the C6 fraction upward. Assuming the sulfur compounds of the various fractions had two fewer carbons per molecule than the corresponding hydrocarbon compounds, it was calculated that in the Cs to C12 range the approximate percentage of sulfur compounds has increased from 0.4 to 7~. In contrast to sulfur, the total nitrogen content of the C4 to C12 fractions was generally less than 160 ppm.
The mercaptan content of the two combined C8 fractions ~shown in Table X) was also determined by difference. At first, the tatal sulfur was determined by sulfur specific GC. Thsn the mercaptans ~ere removed by prscipitating them as silver mercaptldes. Based on such an analysis, the following ppm concentrations were obtained for the various sulfur compounds in the order of their retention times: 2 methyl- and 3-methyl thiophenes, 962 and 612; n-pentane and n-hexanethiols, 105 and 78; C6 branched thioether, 200; l-hexanethiol, 384; 2,5-, 2,4-, 2,3 , 3,4-dimethylthio-phenes, 1245, 945, 728, 289 unknown sulfur compounds, 11. Thus, this analysis provided a total suifur content of 5560 ppm and a mercaptan content of 568. The main group of sulfur compounds were thiophenes in a concentration of 3781 ppm.

Cok~ Ga~ Oi~
Similar characterizations were performed on a light coker gas oil produced by the same Fluid-coking unit from which the coker naphtha was taken.
Fi~ure 7 shows the capillAry GC of the light gas oil in the Cg to C16 range. About 90X of the components are in the Clo to Cls carbon range.
The Cll eO C13 components are pareicularly large. Obviously, there is some overlap between this composition and that of the broad cut naphtha.
As it is indicated by the symbols of the figure, the ~ain component~ are the l-n-olefins and the n-paraffins. In general, the ~2~ f3 ~ l 1~ 0 N O

'`'~` I"' ~ o ~ u o .e ~' ~ Q~ o 0~

I~ 'o ~ - ~ ~
o ~oo I ~ ~o .; ~ U C

~
~o I ~ o ~ oo ~ ~ o ~u U o ~o U ~ ~ D v ~

~ CO ~ O O O V ~

~ ~ ~ 70 C~ ~ ~U
_ _ e ~

_ e I~ E ~ v concentrations of the l-n-olefLns are greater than those of the corresponding paraffins. The l-n-olefins to n-paraffin ratio is apparently maintained with increasing carbon numbers.
The light gas oil fraction was fractionally distilled to produce narrow cut distillates of a particular carbon number. The fractions obtained were then an~lyzed by GC. The data are summarized in Tables XI
and XII. The tables show the a~ounts of the individual cuts, the percentage concentration of the main paraffln and olefin components and separately list tha heart cuts of particularly high content of a l-n-olefin of a certain carbon number. These heart cuts were utllized in subsequent hydroformylation experiments.
The data of the tables show that 54X ~44,439 g) of the distillates were in the C12 to Cls olefln range. It is noted that the percentage values for the l-n-olefin and n-paraffin components are relative. Absolute values could not be deter~ined. Wieh the increasing molecular weight of these fractions, the number of isomers is sharply increasing. Thus, the GC resolution is decreased and absolute accuracy decreased. Nevertheless, it appears at least in a qualitative sense that the l-n-olefin concentrations are maintained.
The Cg to C16 gas oil and selected distillate fractions were also studied by proton NMR. The results are illustrated by the spectrum of Figure 8 which shows the aromatic, olafinic and paxaffinic hydrogens. A
quantitative analysis of the spectru~ showed that this gas oLl is highly olefinic with a strong aliphatic character in that 88.2X of the hydrogens in the ~ixture are on satursted carbons, 6.2X on olefinically unsaturated carbons and only 5.6Z on aromatic rings. Overall, the gas oil has a significantly hlgher percentage of linear olefins than does the coker naphtha ag is shown by the following tabulation:

, ~2 0 cl c~ v 0~ O ~ CO V`
L C C _ _ _ _ -r C L
8~ E co O~ _ --= l_~Z
E c ,~ ", ~ o >~
~ ~o O
~ ~ CO r` , ,_ _ _ ~7 o o ~ C
cn C~

c --~1 '`
c _ ~
C ~ C 0 ~ _ ~ o X _ ~ 0 o o o o o o o o o oO oO ,o o ,_ ~1 ~~ ~ c~
U C

g c cl~
c ~ ~ _~ c~ _ o c, ~n ~ ~ ~ ~o _ _ O 11 ~1 0 ~ I ~ ~I 'I
~ I
1~~ O ~ O ~0 ~ ~ ~ r~
~ C _ ~ _ ~
~_ o o -- -- ~ O C~ _ ~ ~ _ ~ ~ ~.~

o E ~,¦ o ~ ~
~ 0 1 ~ o ~ oo ~1: _ O O O O O O O C~
C ~ V ~ ~ .~7 ~ U- ,~ ,~ ~ ,~
_ C_ ~0 ~ ~ ~ ~ ~ ~ ~ ~ ~ I
~ ~ ~ 1~ ~ ~ 0 ~7 ~ er U7 ~O r~ ~ ~`7 E 0 u7 0 c~ r~
c ~ ~ ~ ~ 2 ~
o u 7 ~ a. ~ ~ X ~ X X X x ,. .

~2Q~$`7 ;3 ¦ e~ N 00 N
L _ _ _ N
C
V~ ~
C ~ ~ O~ O' _ _ _ N N
I O
I E
~ 1 Z N ~1 ~ ~s) O o C~ ~, O
I L C N
., , ~ , ~--¦ , e, o ~o o c 0 ~ ~ ~ ~ ~
X O ~ ~ r _ ~ _ ~J ~ C ~ ~ _ (J N N ~ C~ ~n N 1~ N ~ V:~ ~) N ~ I~
~0 ~ _ ~ O '~:t O ~O N D ~ N N ~ N ~ ~) N N
1~ 0 ~ ~OO O o o o ~ o o o o o o o o o o o o o o o 1~ E _ _ _ ~el c ~ ~ Il ~ ~ co o ~1 o _ N N 1~ O ~0 Ot Ln ~ O N O
._ C ~ U~ N t~ IrJ O ~ 1~1 ~ ) N ~) O t~ _ ~ O U- ~> ~
._ ~. C ~ _ N N _ _ N ~ ~ _ N ~ N _ Vl I ~J N N N N N N ~) ~ ) ~ ~ er ~ G ~r ~ ~ 3 C~ __ U~
_I C ~ O o ~0 ~ ~o ~D o r~ ~ ~D ~r O ~ ~ _ O~ Co O = _ ~ ~
~ V ~ ¦ ~ _ ~ æ _ r~ _ ~ N ~) N ~ N N N r~ _ N ~
3 q_ _ ~1 0 N ~ _ CO N 1~ _ N O ~ ) ~ O r~
O 1~ ~ ~ 1~ ~1 ~D N C~ O CO ~D 1~ a~ O-- 1~ t~ N ~D C0 1~ Ol 0') ~O
~ ~1 ~O~_~N~~tO~Iq_~7~)NO~_ CO

L._ E ~ o ~ N N O ~ ~ N O _ ~ O 14 ) r~ ~ r~ U-o V~l N_~N~er,_NN~~ererln~err~~
: ~ ~ C ~ ~ N N N N ~ N N N N r~J N N N ~ N ~.
c O ~---- N N ~'I ~ ~ ~1 . ~ ~ N N ~ ~ ~ er N Næ N er N _ ~ V~ ~ ~
_ ,_ _ ~ In r~ ~ N ~ N ~ a~ n ~ N Ln ~ ~ L~ n N
3 v~ V _ _ N ~ r r C er ~ ~ u~ ~ ~ r~ ~ o7 _ E ~3 u N ~ n ~n N n N r> a~ ~ in ~ ~ In . In ~ U~ r~
~ o o _ _ N N r ~ ~r ~r er U~ e e`r ~` ct) ~ ~ u7 ~ u~ -- ~
c v a LL~ ~ ~ X X X ~ = X X X --X X X X _ ~ X X
O r~ ~ x x ~e X x X X X X x x x 3 xx t73 Mole X Unsaturation VinylicGas OilNaphtha*
Type SegmentC10-C15 C4-C12 _ II -CH~CH- 22 20 Con~. Diolefin -C-C-C-C- 14 14 ~From Table IV.

Type I olefins represent about 42X of the total olefin content in the gas oil and about 37X in the naphtha. Most of thç Type L olefins are l-n-olefins which do not have branching anywhere on their hydrocarbon chain. Tha mass spectrometry data indicated that branching is mostly by methyl groups on the vinylic doubla bonds.
Selected distillate cuts of the light gas oil were also analyzed by NMR in a similar manner. The distribution of their vinylic hydrogens was particularly studied to determine the relative amounts of the various types of olefins present. The results are summarized in T~ç XIII.
The data of Table XIII show that the relatlve olefin percentages of the distillete cuts vary. However, the percentages of tha Type olefins, including the dssired l-n-olefins, is generally more than a third of the total. The Type I and II olefins combined, which includes all che linear olefins represent more than 55X of the total. The vinylically branched olefins are present in le-~s than 35X~amounts. The percentages of the con~ugated diolefins are included in ehe table since they are converted to ~onoolefins during hydroformylation. However, the diene structures are uncertain and as such of approximata ~alues.
~ Table XIII also shows the distribueion of olefin types in the case~of four narrow cut Cl2 distilIate fractions. As expected, varying a~ounts of the diffcrent types of olefins of different boiling points were iound to be present. Thus, the proportion of the Type I olefins chan~ed from 45.5 to 33.8X.

~2~ 73 4', C~J N
U) ~ ~:t ~) N r C~ U ~ O O
~1 _ N I N O
et ~ I et C~
~ ~ U
O C~J U . . . O~
._ ~ _ ~ L171~ N
er (D
v~ ~1 ~n , a~ '' ` I u7 u~l ` ~' ~
~~ ~ ~ 1 at U~ l ^
O If) O ~ N
s 8 v _ o c~
J O L~
~ U~ ~ _ In O _I In U~ . . = o ,~,~ ~ r~O ~ O
~C O ~ C
~D

c~.l I O O ~I~
a~_ ~ ON O C~J ~ ~ _ 0 ~ O ~ ~ N
'L~
o 3_. u~ ~1 .D ~ ~ ~ ~
~ ~o ~ ~ ~~ ~
~t _ , _ _ a~
c~ ~ ~r r~
_ _ _ _ ~o ai ~ o c~

-.. _ _ C
N ~ I ~
c ~ c~ T I 11 1l 0 ._ I ~ I I C~
n~ O ~
c C ~ cn lJ O ~ _ O
o ~

-65- ~2~73 The percentages of various typ8s of olefinic hydrogens, are shown by Table XIV. From the hydrogen distributions, the weight percentages of the various types of olefins were estimated. As it is shown by Table XII, the estimate of total olefins including dienss is between 50.4 and 61.7X.
It is noted that the 61.7X value is for the C16 fraction which was distilled with decomposit$on. As a result of cracking this fraction contained not only C16 but lower molecular weight olefins as w811. In the case of the C12 range, four narrow cut fractions were analyzed to determine changes in the proportion of different types of compounds. Only moderate changes were found in total olefiD concentration (45.5 to 54.4X).
To illustrate the detailed composition of ths present gas oil feeds, more detailed data are provided on a narrow Cl2 fraction on the basis of GC/MS analyses. Such a cut cannot be separated on a nonpolar (boilin& point) methylsilicone GC column. However, it was found that a highly polar type CP Sil 88 column (with a cyanopropylated silicone stationary phase) separated the various types of components according to their polarity. [This column is particularly suitable for the analysis of high boiling fractions since it has a high use temperature limit (about 275C)]. These components could then be largely identified via GC/MS
studies. Two capillary GC traces with the groups of components identified are shown by Fi&uxe 9.
The effluent of the above polar capillary column was splLt and led to a flame ionization and a sulfur specific detector. The chromato~ram of the flame ionization detector shows tha distribution of the organic compounds according to polarity in the lower part of the Figure. The upper chromato~ra~ produced by the sulfur specific detector shows the elution of the sulfur compounds in the order of their polarity.
The lo~er GC of Figure 9 shows good separation of ehe aliphatic, monoaromatic and diaromatic hydrocarbon components of the Cl2 fraction.
With the help of GC/MS the aliphatic components could be broken down eo paraffins, olefins plus diolefins. Their percentages were 18.6 and 50.5X, respectively. The monoaromatics included alkylbenzenes, naphthenobenzenes and trace ~mounts of alkylthiophenes. The total a~ount of monoaromatics was 28.2X. The main diaromatic compounds were indene, nephthalene and benzothiophen~. Surprisingly, trace amounts of trimethyl phenols were also found.
The upper, sulfur specific GC of Figure 6 shows th~t essentlally all ehe sulfur compounds of che C12 fraction were aroma~ic. The ma~ority ~4~ ~

~ O e~ ~ ~ ~ I_ ~ ~ o~ _ v ~ r~ ~ ~ o 1~ et O
~o ~ ~ o o d- ~ ~ ~
' ~'1 o~ cn cn ~ e~ a~ co ~ o _ a~
C:l V~
~ ~ C O CO 0 1~ 0~ 1~ It~ _ r~ 0 ~ i~
~ I co ~ O 1~ O cr~
~ 8 ~ _. ~ ol N CO O Cl~ O ~ O~ O
,Co I _ _ _ __ O ~ N ~ Q
L

vl 4 e:t O et :) O 1~ r~ N O ~`J
o OU C ~r ~r OD _ e~ n CS ~) N U~ _-.v U L ~ U~ 0 a~ ~ CO O 0 ~ cr~ O
~ L j~ 2_ 1~ CO CO Ct) 00 1:~ O~ 00 CO CO C~t ~

X O ~ ~ --~ C ¦ N 0~ ~) N N ~) 3~ ~, ~ ,,. o ',1 -O N Ir) Ir~ ~D Cl~ ~r ~ ~ C~ ':t 0 ~ C~J ~7 _ C~J ~ ~t ~ ~ ~
~_ O o O O O o o o o o o ~:L L
~ 3~ a~ -- ~ ~ a) ~`J ~ oo u:~

0~11 I--_ .- O O O O O O O O O O
' dl _ _ o ~ ~ o o Cll ~1 a~ ~ 0 ~
L T O -- -- O _ O O O O O O O
V~ .Y
a ~3 ~o _ ~ co ~ ~ ~o ~ J ~ 1~
e ~ _, r~ r_ _ ~ I~ ~ ~ a~ Ln o Vl a ~ ~ ~ c~
L LL
~ 0:
C ~ ~ ~ O U> o o o o .-- ~ E ~ ~C ~ O _ Lrl C~ C~
e O O ~ ~ ~ _ ~ co ~ U7 u~ o ~ ~ ~ _ O O c~J ~
o _ _ o c~
~ O L
_ ~L n ~ ~ E

~2~ 3 wers alkyl thiophenes. Benzothiophene was also present in significant amounts.
A similar analysis of the C14 fraction showed an even better separation of the components according to their polarity. In this case the distribution of the aliphatic components was similar but the ma~or aromatic componsnts were dinuclear: methylnaphthalenes and methylbenzothiophenes.
The dlstillate fractions of light gas oil were also analyzed for elemental composition, particularly for sulfur and nitrogen compounds and mercaptans. The data obtained are summarized in Table XV.
The percantages of carbon and hydrogen were rather well maintained with increasing molecular weights. They indicate that the aliphatic character of the gas oil was f~irly maintained. The total sulfur contant remained at about lX in the Cg to C12 range. Thereafter, thers was a rapid increase of sulfur up to 2.82X in the Cl6 fraction. It is noted that ehere was increas~ng decomposition during the distillatLon of these fractions. When the C16 fraction was red~stilled a broad molecular weight range of l-n-olefins was found in the distillates. This suggests the breakdown of nonvolatile aliphatic sulfur compounds to generate olefins and mercaptans.
The total nitrogen contents of the distillates were more than an order less than that of the total sulfur. The mercaptan content is generally even lower. However, both the nltrogen and mercaptan contents rose sharply in the Cls and C16 fractions.

Exp~rlmental proç~e~ures Except as otherwise specified in the examples, the processes found in those examples were carried out using the following experimental procedure~.

Low ~d 4Odlum P~es~urs Hy~o~o~myla-tio~
The low and medium pressure hydroformylation experiments employed 300 ml and 150 ~1 ste~l autocla~es, resp~ctively. Both autoclaves ~ere equipped with impeller type stirrers operating at 1500 rpm. The co~al liquid feed was 100 g and 50 g, respectively.
In a standard hydroformylation experiment, 80Z of the feed was placed into the autoclave and deoxygenated with repeated pressuriæation with nitrogen. The solution, now at atmospheric nitrogen pressure, was then seal-d and pressured with 1:1 H2/C0 to 50~ of the reaction pressure.

-` ~2~7~

u~ ~n r~ c~ c U~ 1~ ~r~ ~I OD O _ N ~
J O1~ ~ --C~.~
u~
U7 O C~J U~ D O O ~ ~ g ~ u7 U~ d ~
::~ ,o e~ u~ ~ ~ ~ O O c~
d- ~r OD= ~ -- ~ O
U~ ~o ,C _, C~ ~ ~ ~ ~ O 0 1 ~ ~t ~ CO~

-- 1~ ~1 _ co ~ o o (O ~ -~
~
U~ -- C~l ~ o ~
a- ~1~o ~ ~ _ v U- o~
~ ~ n= ~~ o o e~ ~ ~c ~ ~I _ _a~ N _ ,~ o ~

J O _ ¦ . q O
~1 - ~ _ _ W C ~a~
t~
O æ _ N _ _ O 11'~
._ ~ ~ ~ _ O ~ n~ L

~ N _ ~O U _ ~o co ~ o o o co o ~tl C~
~L~ O ~ . ~ ~ ~
_ ~ ~N O o o ~ O O
tJ-o s_~ a~ c~--~ o o o _ _ N ~ U~ C
O
~ _ O

V ~

8 , W' , ~
_ ~~ o V~ E ~ ~D
v~ 2 ~1 ~ 3 ~ O
O ~_ ~O 0~ ~ ' ~ ~_ .~ ~ ~ " ' ~

The catalyst precursors, l.e., rhodium carbonyl acetylacetonate, dicobalt tstracarbonyl or dicobalt octacarbonyl plus the appropriate phosphorus ligand, were dissolved in 20~ of the feed and placed into a pressure vessel connected to the initial H2/~0 feed line and the autoclave.
The autoclave was then heated to the reaction temperature.
Thereafter the catalyst solution, about 40 or 80 ml dependent on the volume of the autoclave, was pressured into the autoclave by the inltial feed gas and ths d2sirsd reaction pressure was established without stlrring.
Thereafter, a switch wa~ made to the feed Bas pressure vessel of known volume which contained an appropriate mixture of H2/C0 at higher initial pressure. Then the stirring of the reactlon mixture started. This resulted in efficiant contact of the gaseous H2/CO with the liquid reaction mixture. As ths reaction proceeded, the reactor pressure dropped due to the H2/CO reactant gas consu~ption. In r~sponse, feed gas was automatically p~ovided as needed to maintain the pressure in the reactor.
The feed gas had an appropriately hi~h H2/C0 ratio above ons so as to provide H2 not only for the main hydroformylation reaction but the hydro~enation side reactions as well.
The progreqs of the hydrofor~ylation was followed on the basis of the C0 and H2 consumed. The latter was calculated on the basis of the pressure drop in the l liter H2/C0 cylinder. Reactant conversion was estimated by plotting the C0 consumption against the reaction time. In soma cases, reaction rates were also estimated in spite of the complexity of the feads and were expressed ns ~he fraction of the theoretical H2/C0 consumed per minute. Reaction rate constants were normalized for lM
transition metal concentration, assuming a first order rate dependence on the metal concsntration.
When the reaction was discontinued, the H2/C0 valve was shut and the autoclave immediately cooled by water The synthesis gas in the head space of the autoclave was analyzed to determine the H2 to C0 ratio. After the release of excess H~/C0, the residual liqu~d reaction mixture was also analyzed to determine conversion selactivity. For these analyses a capillary gas chromatograph with a 50 m fused silica colu~n was used.
Reactant conversions an~ product selectivities were also estimated on the basis of ths gas chromatograms of the reaction mixture.
The conversion of l-n-olefins could be usually determined on the basis of the reduction o~ their peak intensities compared to those of the inert paxaffins. These conversions could be correlated with the formation of che ~26~4~ 3 correspondin~ n-aldehyde and 2-methyl branched aldehyde products. When comparing hydrocarbon slgnal intensities with those of aldehydes and alcohols, a correction factor of 0.7 was assumed for the oxygenated compounds.
When the ma~or products of the present hydroformylation process were alcohols, e.g. in cobalt-phosphine catalyzed reactions, samples of the reaction mixtures were silylated prior to GC analyses. An excess of N-methyl-O-trimethylsilyl-trifluoroacetamide wa~ ussd to convert the alcohol! to trimethylsilyl derivatives:

RCH20H CF3COSi(CH3)3 RCH20Si(CH3)3 These derivatives of increased retention time are easier to chromatographically resolve and determine than their alcohol precursors.

High ~
In the high pressure hydroformylation experLments, a 1 liter and a 1 gallon stirred autoclave were used. In these experiments, the amounts of synthesis gas consumed were not monitored quantitatively. However, the liquid reaction mixture was sa~pled, usually after 10, 30, 120 and 180 minutes, and analyzed to determine ol~fin conversions and product selectivities, Also, the relative reaction rates were estimated by periodically shutting off the synthesis ~as reactant supply and determining the rate of pressure drop per minute in the reactor.
In the one liter autoclave, the thermally cracked distillate was usually dilut~d with an equal amount of n-hexane, to provide a hydroformylation feed for standard experiments. However, about 20X of the diluent was employed to dissolve the catalyst, usually dicobalt octacarbonyl. In the one gallon autoclave, the cracked distillate was placed as such without solvent. The catalyst was usually dissolved in toluene solvent a~ounting to about 5X of the distillate reactant.
The high pressure experiments were carried out in a manner basically similar to those employ~d in the low pressure experiments~. The distillate reactant was typically preheated to the reaction temperaeure with stirr$ng under an initial H2/CO pressure equally about 3/4 of the final rsaction presiure. The catalyst solution was then pressured into -71- ~2~C~3 stirred mixture using the in~tial H2/C0 at reaction pressure and the pressure waq ma~ntained with additional, ~2/C0 feed gas as the reaction proceeded. During the periodical sampling of the liquid mixture, significant losses of H2/C0 occurred, thus the H2/C0 ratio thereafter was that of the feed gas rather than the initial gas. At the completion of the experiment tha reaction mixture was rapidly cooled under H2/C0 pressure and discharged when cold.
For a more detalled study of some of the products of high pressure cobalt hydroformylation, particularly those prepared in the one gallon reactor, the reaction mixtures were fractionally distilled. To avoid decomposition, the cobalt was removed as cobalt acetate by hot aqueous acetic acid plus air treatment. In a typical procedure, a 200X
excess of acetic acid is used aa an about 6X aqueous solution. As a reaction vessel a three necked glass vessel equipped with a mechanical stirrer, sintered glass bubbler, reflux condenser and a bottom valve for liquid takeoff, was used.
The stirred mixture of the cobalt hydroformylation reaction mixture and the theoretical amount of aqueous acetic acid was heated to reflux temparature while introducing air. Thereafter, stirring and aeration were continusd for 20 minutes while refluxing. As indicated by the lightening of the color of the reAction mixture, cobalt conversion was usuall~ substantially complete by the tima refluxing started. The mixture was then allowad to cool and settle. Theraafter, the bottom pink aqueous phase was separated. The organic phase then was treated the same way again. After ths qacond acid wash, tha mixture was filtered if there were any solids present. Thereaftar, two washed with distilled water followed.
Lack of color of the aqueous washings indicated a complete prior removal of cobalt.
The cobalt free orgsnic phase was fractionally distilled in vacuo using a l to 2 ft. long, glass beads packed column or an Oldershaw column with 22 theore~ical plates. The composition of distillate fractions was monitored by capillary GC to halp appropriate fractionation. Many of the fractions were also analyzed by a sulfur specific GC detector. Selected fractions were also analy~ed by a combined gas chromatography/~ass spectrometry ~GC/MS).

.

125~73 ~dsh~d~ H~dro~en~on,~o Produc~ ALcoh~ls Typically, the aldehyde hydrogenation~ were carried out at 3000 p~i (206 atm) pressure in a 1 gallon (about 3.8 liter) rocking aucoclave using about 1800 g reactant. The aldehyde rsactant was used as such or in a hydrocarbon solutlon. Five per cent by weight of water was added to the aldehyde to inhibit the formation of dimesic and trimeric by-products durin~ hydrogenation.
A~ a preferred hydrogenation catalyst, cobalt sulfide molybdenum sulfide on alu~ina was used. Alternatively, molybdenum sulfide on carbon support was employed. Ten percent by weight of catalyst was used. In the presencs of the CoS/MoS bascd catalyst, the hydrogenations could be carried at lower temperatureQ in tha ran8e of 130 to 170C. The low temperatures are important for avoidin~ the undssired conversion of aldehydes to paraffins and sulfur transfer from the metal sulfides co form sulfur containing by-products. In the presence of molybdenum sulfide the hydrogsnations were carried oue a~ 232C (450F). At this temperature, paraffln formation was significant (10 to 30~).
The hydrogenations were substantially compleeed in five hours.
However, they were gen~rally continued for a total period of 20 to 24 hours to assure a completa conversion of ths aldehydes. Ihe alcohol products were usually colorless or very light in color. They were characterized by GC and GC/MS and fractionally distilled in vacuo to provide colorless liquids. Som~ of the alcohols were washed with 10~ aqueous sodium hydroxide to remove hydrogen sulfide and other potential acidic impurities.

~ Q~ Pr~ Llbob3a~Q~ ation of C4-C12Naphtha Frace~,,o~s i~ tbo ~ u~Lsa~_~b:~phln~-ahQdium Com~le~es ~xamplc3 1-12~
Tho previoucly described C4 to C12 Fluid-coker naphtha and its di3tillate fractions were hydroformylated without prior treating in th~
pre~ence of rhodium co~plexes of various phosphines under varying low pressure conditions.
The rhodium catalyse systam~ employed and the reaction conditions used are u~arized togethar with som2 r~ults for orientation in Table ~21- In general, in the presence of sufficient a~ounts of phosphine-rhodium cstalyst complexes, rapLd and selective hydroformylation occurs ae low pressure. Very little hydroesnation occurs. CC analysis provideq a quantitative measure of tha two ~a~or aldehyde produces and a more qualitative estimate of the toeal a1dehyde products. At low pressure, ~Z~ 3 ~7 cl o ,~ ,~ E ~ ~ o ~oO o O co o ~D o ccl c~ _~ ~ ~ I ~ r~ o~ ~ ~ Q
C O
_ ~ V
_ E ._ n l O C 2r~ 00 ~ o _~ o ~ a a~
c~lo ~ c~ _ _ o o lel ~: O C~ l r~ ~ C~l C`J cr~ N C~J ~ I
~ CJ
al ~ ~7 ~ n c o a~ ~ n o ~:L O ~ ._ I~ CO C~ D O aJ
1~ ~ C~ ~ ~
~:3 rn O C
,9, L. 1~ ~ o a) la _ ~2~ o v~ ~ _ _ a~ o 1-- o o c~
Ecc ~ o ~ __ __ _ __ ~ c c C~~E ~ ~ ~ "
tU~ 3 t_ --~ c~ ~I ~ ~ ~ ~7~c~ ~ ~oLI~ .~
C ¦ t _ N _ N C~J C~ 0 _ o _ O

H ~: O . U'l 0~:) Ln ~ 1~. LO I ~ C ~ e ..... ~c O~ 0OO ~ O _.~0 _ c, ~ c ~_ . ... . . . ... , ._ ~
OJ J~ T Ll_ _ _ _ ~ _ ~ O _ Ir~ ') _ r~ O C
,1 o E c o o o oN O O ~1 o O o CL C 1~ O ~O ~ r.~l ~O _~O ~ ~O
O U
V~ ._ O,J r_ n OO O O O O O O OO O O
O ~_ t_ ~ o o 1~7 0 0 0 00 01~ Ln O O
Q ----~-- ------ ~
~ ~ C
'~ ~ C C
E~ I O O O O O o O o o o Ln o ~ ~ ~ c~ o ~ co o ~ ~ o O ~ o -- -------- E s '- s G c~ c I o c~ ~ a~ c~. c~ ~ ~ ~ ~ ~ _ '-- C t-- c~ ~ t~c~ ~ c~. ~~ I--el Jl ~ ~ ~ o c c ~
--~ _.1 ~ X T I oT ~-- :~ TI Q L7 ~ ~ ~ Q
~ ~ ~ .) ~ S ~_ 5) 0 0 ~t) . C ~ ~ Q `_ L ~J
o ~ . e ~ U- S~
~n r,~ ~ ~ S ~~ ~ e~ ~ o t_ O~ _ ___ _ _ O oo_ C~ O O O O O O O O r O_ _ O ~ C C
~:: O O
~IJ
C I Q
~y ~ O ~ ~~ ~ ~ C o o C . ~
q~ ~o o ~ ~ ~ ~ o a~ ~ z _- o o o 1--1~ 1~ o o o o o -- --.~__ O O
L-I ~ CS 1-~ Z~ 3 the total aldehyda products could be more reliably estlmated, on the basis of the H2/C0 consumed, by comparing the found values with the amounts calculated for converting the l-n-olefin component. Based on the initial rates of H2/C0 consumption (0-1 minute) the hydroformylation rates of t'ne most reacti~e l-n-olefin components were also compared in the presence of different catalyst complexes.
Comparative l-n-decene hydroformylation experiments with the Clo naphtha fraction as a feed showed that the activity and selectivity of rhodium complex catalysts could be controlled by the chemical structure and excess concentration of the phosphine ligand added, as it will be discussed in the individual examples.

Ex~mple 1 ~ydroformyl~tion of ~ C4-C12 Naphth~ ~ith a Tr~butyl Phosphlno Rhodlu~ Complo~
A broad naphtha cut previousiy dascribed was hydroformylated in the presence of a catalyst system containing 10 ~M rhodium, employed as dlcarbonyl acetylacetonate, and 0.14~ tri-n-butyl phosphine. The reaction was run at 180 under 1000 psi (6900 kPa) pressure for 40 minutes. The inltial H2/C0 ratio was 1, the H2/C0 feed ratio employed during the run 1.22 and the final head space ratio 1.95. The increase of the H2/C0 ratio during the run indicated that very littl0 hydrogenaeion side reaction occurred.
The final roaction mixture was analyzed by GC. The chromatogram showed no l-n-olefin components, indicating their complete conversion. The main products were the n-aldehydes. Among the minor aldehyde products, thos~ of the 2-methyl substituted aldehydes were readily recognizable.
i~L~YII shows the signal intensities of these two types of aldehyde products and those of the n-paraff~n components. The paraffin components repres~nt ~ultiple internal 3tandards which were present in the starting reactants in amount~ comparable to the l-n-olefin reactants of corresponding carbon nu~bers. The data of the table qualitatively show that the convsrsion of the l-n-olefins resulted in the formation of the expectsd normal aldehyde and 2-methyl branched aldehyde products:

CnH2n+lcH-cH2 /i2~ CnH2n~1CH2CH2CHO + CnH2n~lCHCHO

~2~

~able XVI~
Ma~or Aldahyte Product~ snt n-Paraffin Components of ~luid Coker Naph~h-Alkyl GC Sign~l ~ntensltv. X
Carbon Normal2-Methyl Normal No, B~ Aldehyde ~3E~ LB
1.104 0.926 0.798 6 1.837 1.468 7 1.796 2.927 8 2.259 1.586 3.064 g 2.0~7 1.350 2.208 2.182 1.115 2.043 ll 1.~23 0.715 1.409 12 0.514 0.239 0.393 5-12 13.162 14.310 The n/i ratlo of these linear versus branched aldehydes is about 2. Using the present cataly3t system and conditlon~, this ratio is in the ran~e oi n/i values obtained on the hydroformylation of pure l-n-olefins and Type I olefins, in general. As eh~ l-n-olefins were converted, the reaction rate d&creased and the reaction was discontinued. Thus, the results of this example indicate that the l-n-olefin components of the diqtillate ieed can be selectively hydrofor~ylated in the presence of phosphine rhodium complex based catalysts.

E~ample 2 Hyd~ofor~ylation of Clo N~phth~ ~ith a Tri-n-octyl Phosphine Rhodlu~ Co~pls~ at 1000 p~i The pr~viously described Clo fraction of the Fluid coker-naphtha was hydroformylated at 180C under 1000 psi, using the low pressure procadure. The catalyst system was derivad from 2mM rhodium dicarbonyl acetylacetonate and 0.14M tri-n-octyl phosphine. The reaction period was 60 minutes. The ratio of the initial H2/C0 wa~ 1; the H2/C0 feed was of 51 to 49 ratio. The final H2/C0 ratio of the head space was 52 to 48, indicating a virtual absence oi hydrogenation.~
The reactio~ was very fast during the initial period of about 5 minute~, then the reaction became slower and slower. Apparently, the l-n-decene component of the feed was rapidly hydroformylated while the isomQriC Type II and Typs III deoenea ~ere more sluggish to react.

-76- ~2~ 7~

A GC analysis of the final reaction mixture showed that l-n-decene was absent. Apparently, it reacted to form n-undecanal and 2-methyl decanal. The latter compounds constituted about 69X of the total aldehydes fsr~ed. The ratio of the normal to the iso aldehyde produced ~as 1.88.
On the basis of the origlnal conceneration of l-n-decene in the feed, the theoretical amount of Cll aldehydes was calculated. The total aldehydes were 171X of the amount which could have been derived from l-n-decene. Apparently ma~or amounts of the Type II decens components of the feed were also hydroformylated On the other hand, the GC showed that 2-methylnonene was still substan~ially unconverted in the rPaction mixture.
This indlcated that the Type III olefins of tha feed are of low reactivity in the presence of this catalyst system.

Example 3 Hydroform~lat~on of Clo Naphtha ~ith a Tri-n-octyl Pho~phine Rhodiu~ Complex at 350 psi The experiment of Example 2 was repeated-at 350 psi instead of 1000 psi pressure. Qualitativ~ly, the reaction was very similar. The reaction rate was only slightly lower. The final H2/CO ratio in the head space was 51/49.
The ratio of the two major products, n-undec~nal versus 2-methyldecanal was about 2. These two aldehydes represent ll9X of the calculated yield based on the starting l-n-decene. The total aldehyde yield Ls 187~ of the l-decene based value. Thus, the amount of the above two aldehydes is about 62X of the total.

Example 4 Hydrofor~ylatlon o~ Clo Naph~ha With a Tri-l-octyl Ph~sphine Rhodiu~ Comple~
Ex~mple 2 was repeated using the rhodium complex of tri-i-octyl phosphine ~tris-(2,4,4-tri~ethyl-pentyl)pho~phine] as the catalyst instead of that of tri-n-octyl phosphine. The reaction was very similar to that of Example 2 except for the lower n/i ratio of the two maLn products. The ratio of n-undecanal to 2-msthyl decanal was 1.64 in the present experiments whila a ratio of 1.88 was found in Example 2. The reduced n/i ratio wss apparently a result o the steric crowding effece of the bulky tri-i-octyl phosphine ligand.

77 123~73 The two main aldehyde products represent 94X of the theoretical yield based on the l-n-decene content of the feed. On the same basis, the yield of the total aldehydes was fond to be 128Z. Thus, the two main aldehydes amounted to about 74Z of the total aldehydes produced.

Example 5-7 Hydroformylation o~ C7 Naphtha ~ith Tri-n-butyl Phosphin~ Rhodlum Comple~
The previously described C7 fraction of the Fluid coker naphtha was hydroformylated at 180C uDder 1000 psi pressure with the standard low pressure procedure using 1/1 H2/CO as reactant. Three hydroPormylation experiments were carried out using different concentrations of rhodium in the presence of excess tri-n-butyl phosphine at 0.14M concentration. The rhodium was provided as a dicarbonyl acetylacetonate derivative in 1,2 and lOmM concantration. Reasonably fast reaction occurred with 2mM rhodium.
The results of ~his experiment (Example 5) will be discussed at first.
Gas consumption data indicate thae initially the reaction rate was very high, but started to drop in 2 minutes. When the reaction was discontinued after 12 minutes, gas absorption was minimal. The H2~CO ratio remained close to 1 during the reaction.
Gas chromatography showed that 42X of the l-n-heptene component of the faed was reacted. The l-n-heptene derived component of the product was mostly n-octanal and 2-methylheptanal. The n/i ratLo of these products was 2.3. The amount of the two compounds was 115X of the calculated value based on the converted n-l-heptene. The total aldehyde products correspond to 133X of that value. Apparently, minor amounts of other heptene isomers besides I-n-heptene were also raacted.
In another experimant (Example 6) the same reaction was run in the presence of lOmM rhodium. This resulted in an extremely fast reaction.
About 0.645 moles of H2/CO mixture wa~ consumed within the one minute reaction time. The run gas used had a 52/48 ratio. The final ratio of H2/CO was 1.47, a substantial increasa over the initial H2JCO ratio of 1.
Apparently, no significant hydrogenation occurred.
The gas chromatogram of the reaction mixture showed that all the l-n-heptene was converted. The two main products were again n-octanal and l-methyl heptanal, in a ratio of 2.15. The sum of these two corresponds to 18% more than the 2~ount which could ha~e been theoretically derived from l-heptene. The total amount of aldehyde product is 165X of the amoun.

~2~ 3 derivable from l-heptene. Thus, the n-octanal formed equals to 48~ of the total aldehydes formed.
In a third expariment (Example 7) only lmM rhodium was employed.
At ehis low catalyse concentration, little reaction occurred. In 20 minutes only 15Z of the l-n-heptene wa~ consumed. The n/i ratio of the two main products was 2.3.

Example 8 ~ydrofor~yl~tion of Clo Naphtha with Rhodiu~ Co~ple~
in tha Presence of 1~ Trlb~tyl Phosphina The Clo fraction of the coker naphtha was hydroformylated under the conditions of Example 2. Ho~ever, LM tri-n-butyl phosphine was used insteat of 0.14~ tri-n-octyl phosphine to ascertain the effect of an increased excess of phosphine ligand. Also, 4mM Lnstead of 2~M rhodium was used to counteract the inhlbitory effect of the added ligand.
The initial reaction was very fast. All the l-n-decene was converted in about 140 seconds. Thereafter, the internal decenes were being converted at a much slower rate. At 60 minutes, the C0/H2 consumption rate was quite low. Tha reaction was discontinued after 60 minutes.
A GC analysis of the r0action mixtura showed that the two main reaction products, n-undecanal and 2-mathylnonanal were formed at an increased ratio. Due to the increased excess trialkyl phosphine ligand concentration, the n/i value was significantly higher, 2.02. (In the presance of the smaller ligand concentration Example 3t the n/i ratio was 1.88). The amount of ths two maJor products was 102X of the v~lue calculated for the ~ounts derivable for l-n-decene. The total amount of aldehyde product~ form~d was 130X of the theoretical value calculated for l-n-dacene.

Example 9 ~ydroPormylation of Clo Naphtha ~ith Rhodiu~ Dicarbonyl Ac~tyl~cetona~a The sam~ Clo naphtha was also hydroformylated under the conditions of the previous example, but wlthout any phosphine catalyst modifier. In this example, tha usual rhodiu~ catalyst precursor, rhodiu~
dicarbonyl acetylacetonate was used alone in amounts corresponding to 2~M
rhodiu~ conc~ntration~

., 73 Apparently due to the absence of phosphine modifying ligand, ehe reaction was slow. Although the reaction time wa~ lncreased to 120 minutes, even the converstion of the most reacti~e olefin component of the feed, l-n-decene, remained incomplete. Also, the amount of the C0/H2 reactant gas consumed WAS only about half of that of the previous example (The 1/1 ratio of H2/CO wa8 well maintained during reaction).
The main products of the reaction were again undecanal and 2-methyldecanal derived from l-n-decene. They represented about 77X of the aldehyde products. No alcohol product was observed. The n/l ratio of the two main products was 1.93.

Exæmple lO
~ydro~orDylation o~ Clo ~nphth~ vleh T~i-n-butyl Phosphina Rhotium Co~ple~ at 350 psi S/l ~2/GO ProQsur6 The Clo naphtha was hydroformylated under the conditions of Example 8, but at reduced pressure, at 350 psi of 5/1 H2/C0. The amount of rhodium was cut to 2mM. The trl-n-butyl phosphine concentration was ehe sa~e, LM. The 5/1 H2/CO ratio wa_ maintained by a feed gas ratio of 53/47.
Tha sharply reduced C0 partial pressure of this reaction signlficantly increased the n/i ratio of thc two ma~or aldehyde products without a ma~or drop in the raaction rate.
Compared to Example 8, the n/i ratio of the two main products increased from 2.02 to 3.2. These two products represented 68.5X of the total aldehyde yield. No alcohols were formed during the 60 minutes resction time. The yield based on l-decene was lOlX ~or the two main aldehyd~s. The total aldehydes amounted eo 147X of the l-decene based calculated yield, indicating a qigniflcant con~ersion of some of the other olefin co~ponents of the feed. The amount of H2/CO needed to hydroformylate all the l-decene was consumed during the first 7 minutes of the experiment.

~xample 11 ~ydroformrlation of Cl~ ~aphtha ~ith a ~hodium Compl~
of n-Oc~adsc~l Diph~n~l ~hosphln~ at 14S~C
The Clo naphtha fraction wa3 hydroformylated with the rhodium complex of an alkyl diaryl phosphine ~o produce a higher ratio of normal versus iso aldehyde products. To deri~e the catalyst system, 2mM rhodium and lM n-octatecyl diphenyl phosphine were used. The reac~ion was run at -80- ~2~ 3 145C und~r 350 psi 5/1 H2/C0 pressure. During the reaction a 53/47 mlxture of H2/C0 was fed. This feed gas more than maintained the initial H2/C0 ratio during the 60 minutes run. The final H2/C0 ratio was 5.75, indicating the absence of ma~or hydrogenation side reaction. Compared to the previous example the difference is in the type of phosphine ligand used and the reaction temperature.
The use of the alkyl diaryl phosphine ligand resulted in a much increased selectivity of l-n-decene hydroformylation to n-undecanal. The n/l ratio of the two main aldehyde products was 6.76. Also, in the presence of this ligand a faster hydroformylation rate was observed. An amount of H2/C0 sufficient to convert all the l-n-decene was consum~d withln 3 minutes.
After the 60 minutes reaction time, GC analyses indicated that the amount of the two main aldehyde products was 106X of the calculated yield for l-n-decene. The total aldehyde product were 164X of this yield and no alcohols were formed.

Example 12 Hydroformylation of Clo Naphth~ ~ith a ~hotiu~ Compler of Trl-i butyl Phosphine Tha C7 naphtha fraction was hydroformylated under conditions similar to those in Examples 2, i.e., at 180C under 1000 psi 1/1 H2/C0 pressure. However, instead of a tri-n-alkyl phosphine, a sterically crowded tri-i-alkyl phosphine, tri-2-methylpropyl phosphine (tri-i-butyl phosphino) was used. The phosphorus ligand concentration was 0.14M, the rhodiu~ concentration 2mM. Feeding a 51/49 mixture of H2/CO as usual maintained the equimolar synthesls ~as reactant mixture during the 60 minutes reaction tim~.
The use of the tri-i-butyl phosphine ligand resulted in a fast resction of low n/i selectivity. Enough H2/C0 reactant was consumed during th~ first minute of the reaction to convert all the l-n-decene in the reaction mixture. The n/i ratio of the two main aldehyde products was 1.25. After the complete run, GC showed that the co~bined yield oi` the two main products formed was 90~ of the value calculated for l-n-decene. The total aldehyde yield corr~sponded to 161X of ehis valu~. In this reaction ~inor amounts of alcohols were also formed. Thus, the combined yield of aldehydes and alcohols was 165% of the theoretical yield of the hydrofor~ylation of the l-n-decene com?onent.

~xample 13 ~ydrofor~ylation of C16-Cl~ Gas Oil with Tri-i-butyl Pho~phine Rhodiu~ Comple~ st 180C and 1000 psi A broad cut light gas oil from a Fluid coker was distilled in vacuo to provids a C16-Clg fraction, having a bolling range of 74-82C at 0.lm~. A capillary GC analysis of this fraction showed that it contained approximately the follo~ing percenta~es of l-n-olafins (Cn~) and n-paraffins (Cn) : Cl~, 0.30; Cls, 0.28; Cl~, 10.06; C16,6.25; C17, 9.55; C17, 7.90; Cl~, 3.34; Clg, 3.10; Cl~, 0.78; Clg, 0.62.
About lOOg of the above distillate feed was hydroformylated using the low pressur2 hydroformylation procedure under 1000 psi 1/1 H2/CO
pressure at 180C in the pres~nce of 2m~ rhodium and 140 ~M triisobutyl phosphine.
The gas consumption data indicated a very fast initial reaction, apparently a very effective conversion of the l-n-olefin components. After this initial stage, the rate was steadily declining as the less reactive olefins were being converted. At a gas consumption calculated for a 50X
conversion of a C17 feed of SOX olefin content, the reaction was di3continued.
Capillary GC analysis of the reaction mixture showed a complete conversion of the l-n-olefins and the formation of the corresponding l-n-aldehydes and 2-methyl substituted aldehyd~s having one carbon more than the parent olefin. The ratio of these n- and i-aldehyde products was 1.35.
Together, they represented 69X of ths total aldehydes formed. A comparison of the intensities of the peaks of the two ma~or ~ypes of aldehyde products and the n-paraffins showed that the yield of these aldehydes is about 61X
of the calculated value for the l-n-olefins. Thus a significant l-n-olefin to int~rnal olefin isomerization occurred during hydroformylation. The linear olefins formed were converted to 2-ethyl and higher alkyl sub~tituted aldehydes which constitute moqt of the minor C17-Clg aldehyde products.
The reaction mixture was distillsd in vacuo to separats the feed from the products. About 15g of clsar yellow-greenish product was obtained as a distillate, boiling in the range of 102 to 124G at O.OSm~.

129 L~97~

~etium Pressure HydroformYlation in the Pre~ence o~
Phosphlne-Cobalt ~9=~ uL ~ =e~e~ 14-18~
The previously described, untreated C4 to C12 Fluid coker naphtha and i~s distillate fractions were also hydroformylated in the presence of cobalt co~plexes of trialkyl phosphine complexes. The reaction conditions used and results obtained are summarized in ~ LISYILI
In general, the substitution of cobalt for rhodium in these phosphine complex catalyst systems changes the activity and the selectivity of the system. The inherent activity of the cobalt systems is about 2 orders of magnituda smaller. In contrast to rhodium, the cobalt complexes are multifunceional catalysts. Olefin isomerization is extensive; this results in an increase of the n/i ratio of the products. Aldehyde to alcohol hydrogenatlon is also extensiva. Since the ma~or products are alcohols and the reactions are performed at medium rather than low pres ure, syn gas consu~ption based olefin converstions are relative rather than absolute values.

Example 14 Hydro~or~ylation of a C4 to C12 Naphtha With a Tributyl Phosphine Cobalt Complex About 93.8g of the broad cut naphtha feed previously described was hydroformylated in the presenca of a catalyst system containing 80mM of cobalt, added as dicobalt octacarbonyl, and 0.24M tri-n-butyl phosphine tP/Co - 3). The reaction was run under the conditions of the first example (180C, 1000 psi) but, for a longer period (60 minutes). While the initial H2/C0 ratio was again 1/1, the synthesis gas added during the run had a significantly higher H2/C0 ratio of 3/2. This higher run gas ratio was amployed because cobalt phosphine complexes catalyze both olefin hydroformylation to aldehydes and aldehyda reduction to alcohols.
During the reaction about 1 mole of H2~C0 mixture was consumed.
In contrast to the first example, no significant reduction in the reaction rate W85 observed. The final hsad space ratio of H2/C0 dropped to 0~68, indicating that hydrogenation took place to a major degree.
The final reaction mixture was again analy7ed by GC. The chromatogra~ obtained showed an essentially complete converstion of the l-n-olefin components and the formation of major amounts of the corresponding n-aldehydes and alcohols.

O o ~ ¦ '` .D O U ~
C a: ol _ ,o c S ~ V ~
t`') O ~ 1-- C~J ) ~ N ~ O O
~ " ~ ~ ,c c E E

r E ~I o ~ -- O u~ ~ e O _ _ ~Co ~ VC~ ~ ~ ~ ,c ._, ~ o ~ ~ ~ oo u. ~ oC O 8 O L ._ I~ N ~ ~ L
~- a ~ ~ e v ~

. ~ CL~ ~ C
~ ~ .C~ ~ O C O O U~ ~ O ~ C

e~ :~ CO ~o _ ~0 ~

_ ~ V~
V (~- ~ O C
O ~ O C ~U~ lS~ O O _ ~ ~ o _ -~

O~ ~ r_ 3 ~ C
cl o o o o o >~
_ C C ._._ L ,~, O O ~ E ~) ~LI ~J ~ C
~ U~ O O O o o ~V~ C ~ ~ _C
oc ~ ~._ o o o c:~ o L, ~_ ~ O ~ V) 0 1~- 0 U~U~ O C ~_ O
O'~ ) ~ C!. _ ,~
LU~ ~') C ~ ~
o t ~ O cl o o ~. o c`J ~ o ~ O O
o~ ~J L 3 C
~ ~ ~ ~ l_ O O I e ~ U V

U~ O ~
X E ~ ~ ~ ~n ~D 1~ X C X C C

Example 15 ~ydroformylatlon o~ Clo N~phths ~ith ~
Tri-n-oct~l Phosphlna Cobalt Co~plo~ at 1500 p~i The Clo fraction of the Fluid-coker naphtha used in the pre~ious examples was also hydroformylated using a catalyst system based on dicobalt octacarbonyl and tri-n-octyl phosphine. The concentrations were 40mM
cobalt and 120~M phosphine ligand (P/Co - 6). The reaction was carried out at 180C under 1500 psi for 2 hours. The initial H2/CO ratio was 1.
During the run an H2/C0 ratio of 60/40 W~9 used. The final H2/C0 ratio of the head space was 48/50. Ther~ was no apparent decrease of hydroformylation rate during the reaction. The maximu~ rate wa~ reached after about 10 ~inutes. In 120 minutes, the H2/C0 feed consumed was about 155Z of tha a~ount theoretically required to convert the l-n-decene component to undecyl alcohol.
The gas chromatogram of the final reaction mixture shows no significant amount~ of l-n-decene present. However, other decene isomers appear to be present in increased amounts as a consequence of concurrent isomerization-hydroformylation.
The hydrofor~ylstion produccd tha expected two significant aldehyde products derived from l-n-decene. However, thesa wera largely hydroganated to the corresponding alcohols, as shown by the reaction scheme:

iso~erization CH3(CH2)xcH-cH(cH2)ycH3 ~ CH3(CH2)7CH-cH2 x+y - 6 ~ C0/H2 C8H17CH2CH2CHO + C8H17CHCHO

! H2 C8H17CH2CH2cH20H C8Hl7cHcH2oH

The a~ount of the above 4 products i9 about 75.5X of the calculated yield for l-dacene.

-85~ $~3 The total yield of aldehydes plus alcohols was also estimated on ehe basis of the capillary GC analysis of the final reactlon mixture. It was 139X of the products calculated for a complete conversion of the l-n-decene component. The n-aldehyde plus n-alcohol amounted to 52.lX of the total products. Most of the products, 92.lX were alcohols. Only about 7.9X ware aldehydes. The n/i ratio of the 4 ma~or products, mostly derived from l-n-decene was high, 7.62.

Example 16 Hydroformyl~ion of C7 ~aphtha ~ th a ~ributyl Phosph~ne Cob~lt Compl~
The C7 fraction of the Fluid-coker naphtha employed in Examples 5, 6 and 7 was also hydrofor~ylated with a catalyst system derived from dicob~lt octacarbonyl and trioctyl phosphine. Forty mM cobalt and 0.12~M
l~gand were used tP/CO ~ 3). The reaction conditions were similar to those in Example 6: 180C, 1500 psi and 1 hour using a 60/40 ratio of run gas.
The initial and final ratio of H2/CO in the reactor were both very close to 1. The H2/CO feed consumed was about 70X of the amount calculated for the conversiton of the l-n-heptene component to octanols.
According to GC there was no unconverted l-n-heptene left in the react~on mixture. Besldes hydroformylation, isomerization occurred. The ma~or hydroformylation products prcsent were n-octanal, 2-methylheptanal and tha corresponding alcohol hydrogenation products. The overall n/i ratio of these products i9 about 10.06. These four products represent about 56X of the total aldehyde and alcohol products. About 58.3X of the total product.~ were alcohols. The significant percentage, 41.7%, of the aldehydes present lndicatqd that the hydrogenation reaction was incomplete.

Examples 17 and 1~
~ydrofor~ylation of Clo ~aphtha ~ith a Tr~-n-bueyl Phosphina Cobalt Co~ple~
The Clo fraction of the coker naphtha was hydrofor~ylated in the presence of dicobalt octac2rbonyl plus tri-n-butyl phosphine catalyst systems having a P~Co ratio of 3. The reactlons were run at 180C under 1500 p9i 1/1 H2/CO pressurq. The high H2/CO ratio was maintained by the addition o~ a 60/40 feed gas mixturs during the reaction.
Th~ rate of absorption of the H2/CO reactant gas showed that the reaction has an inital inhibition perlod, dependene on the concentration of -8~

catalyst. At 40mM cobalt, this inhibition period is about S minutes; at 120mM Co, it is less than 1 min. At 40mM cobalt (Example 16), it takes about 35 minutes to consume enough H2/CO for a complete converstion of the l-n-decene component of the naphtha cut. At 120mM cobalt (Example 17), only about 10 minutes are required to achieve thi~ conversion. The rate of absorption indicate a first order reaction rate dependence on cobalt concentration.
The first reaction with ~OmM cobalt (Example 16) was run for a total of 1290 minutes. In that time 0.254 moles of H2/CO was consumed.
This is about two and a half fold of ths amount necessary to convert the l-decene component to the corresponding aldehydes. However, most of the primary aldehyde products were reduced to the cor~esponding alcohols. The two main aldehyde products and the corresponding alcohols are derived from l-decene via combined isomerization hydroformylation a~ described in Exa~ple 14. Capillary GC indicated that the yield of the total oxygenated products 63.2X of the value calculated for a complete conversion of the l-decene component. About half of the products were of straight chain.
Most of the products, 91.2X were alcohols rather than aldehydes. The n/i ratio of the four major products was 7.
The second reaction with 120 mM cobalt tExample 17) was run for a total of 60 minutes and consumed 0.292 moles of H2/CO. This is almost 3 fold of the amount needed to convert l-decene to aldehydes. Again most of the aldehydes formed were reduced to alcohols. Caplllary GC indicated that the increased cat~lyst concentration resulted in approximaeely doubling the total product yield to 129X of the cslculated value for the l-`n-decens feed component. The yield of the ~our major products which could be derived from l-n-decene was 64.8X. The n/i ratio of these product~ was 8.45.
About 44.8X of the total products was completely linear.

Examples 19 and 20 Hydroformylation of 2-Butene with a ~ri-n-Butyl rho~phlne Cobalt Co~ple~ ant Adted Th{oL
Comparative hydroformylatlon experiments were carried out with 2-butene as a model olefin reactan~ under the conditions of Example 13 to demonstrate that thiol inhibition can be overco~e by the use of cobalt phosphino complex catalysts in the present process.
Two reactions were carried out, each starting with lOQg reaction mixture coneaining 20g (0.1 mole) 2-butene, 2.43g (12 milimole) -87- 12~

tri-n-butylphosphine and 0.68g (2 milimole) dicobalt octacarbonyl in 2-ethylhexyl acetate as a solvent. One of the reaction mixtures also contained 38.8 mg (0.626 milimole) ethyl mercaptan to provide 200 ppm mercaptan sulfur. Both reactant solutions were reacted with 1/1 H2/Co undar 1000 psi pressure at 180C. An equimolar ratio of H2/CO was maintained during the run by supplying addi~ional H2/CO in a 3/2 ratio during the re~ction.
Both reaction mixtures were hydroformylated with simLlar selectivity. The only slgnificant difference was in the reaction rates.
The 2-butene was more reactive in the absence of ethane~hiol. In the absence of the thiol, 50X olefin conversion was achieved within 18 minutes.
In the presence of the thiol, a similar convsrsion took 36 minutes.
After the reaction, both m1xeuras were analyzed~ The most signiiicant difference between the mlxtures was the selectivity to l-butene; 10.5X in the absence of thiol versus 5.8X in its presence. This indicated inhibition by the thiol of the isomerization of 2-butene to produce the more reactive l-butene which is then hydroformylated to produce n-valeraldehyde with high selectivity. The latter is largely converted by hydrogenation to n-amyl alcohol.

CH3CH-CHCH3 ~ CH3cH2cH-cH2 COtH2 ~ CH3CH2CH2CH~CHO
~H2 CH3CH2CH2CH2cH20H

The selectivities toward the various oxygenated products were similar in the absence and presonce of thiol: overall n/i 8.15 vs. 8.92;
alcohol/aldchyde 0.52 vs. 0.57; aldehyde n/i 6.81 vs. 7.34; alcohol n/i 12.6 v~. 13.8.

Example 21 ~ytrofor~ylation of Cg-C16 Li8ht Gas Oil With Trioctyl Phosphine Cobalt Co~ple~
Th9 previously described Cg-C16 light gas oil was hydroformylated using a tri-n-octyl phosphine cobalt complex based catalyst system at 180C
under 1000 psi pressure and a 3~2 H2/CO reactane ratio. Cobalt carbonyl was employed as a catalyst precursor; its concentration was 40~M, i.e., 0.0472X cob~lt metal. The phosphine ligand was employed in 240mM
concsntration to provide a 3/1 P/Co ratio. It was added to stabilize the cobalt and to obtain a more linear product.
The reaction was carried out without solvent. No induction period was observed. The reaction was discontinued after 60 minutes, although H2/CO uptake continued throughout the reaction period The amountof H2 and C0 consumed indicated that hydroformylation and hydrogenation both occurred to a great extent. GC indicated that the products were mainly alcohols. To enhance the analysis of the alcohol products in the GC, the rsaction mixtures were treated with an excess of a silylating reagen~ which acts to convert the -CH20H groups of the alcohols to -CH20Si(CH3)3 ~roups. The retention time of the resulting capped alcohols in the GC column is significantly increased. The shifts of retention times by silylation confirmed that ~he main products were alcohols.
The GC of the final silylated reaction mixture is shown by Figure 10. The GC shows that none of the l-n-olefin componsnts of the feed remain in the product stream. The capped alcohol products are mostly n-alcohol derivatives. Although many branched alcohol derivatives are present, they are mostly in minor amounts. Due to their increased retention time, the peaks of most of the capped alcohols is beyond those of the hydrocarbon feed.
A comparison of the peak heights of the capped n-alcohol products derivad from gas oil indicated a distribution similar to that of the starting l-n-olefins (and n-paraffins). Thus, the reactivity of the feed l-n-olefins Ls essentially independent of the olefins' carbon number in the presence of the phosphine cobalt complex catalyst.

Example 22 ~ydroior~ylation of Clo Gs3 0$1 ~ith Triethyl Pho~phine Gobalt Comple~
The hydroformylatlon of the previously described Clo coker gas oil fraction was also attempted in the presence of a tri-n-alkyl phosphine cobalt co~plex catalyst at hi8h pressure, i.e., 3000 p5i . Examples 14-18 have shown us that phosphine cobalt complexes catalyze coker naphtha hydrofor~ylation under low pressure, i.e., 1000 psi at 180C and mediu~
pressure, i.e., 1500 psi at 180C. The purpose of the present experiments wa~ to detemine the effect of pressure on the stability and selectivity of the catalyst system.

~2~ 3 Triethyl phosphine was selected as the ligand because it is potentially applicable in the present high temperature process. Triethyl phosphine is fairly volatile (bp. 130lC), thus excess li~and can be removed as a forerun by distillation if desired. Triethyl phosphine can be also readily removsd from the reaction mixture by an aqueous acid wash and then recovered by the addition of a base.
As a precursor for the phosphine complex, dicobalt oceacarbonyl was employed. An amount equivalent to 0.472~ Co was used [0.04M Co2(CO)g]
The triethyl phosphine added was 2.9% (0.24~). Thus the P/Co ratio was 3.
The triethyl phosphine catalyst was dissolved in the naphtha feed which was then heated under ~2/C0 pressure. Under tha reaction conditions, a concantrated solution of the dicobalt octacarbonyl was added to the reaction mixtura to preform the catalyst and start the reaction.
The reaction was followed by capillary GC analyses of samples taken after 10, 30, 60, 120 and 180 mlnutes. Extensive isomerization of l-n-decene to internal decenes occurred in 30 minutes. Hydroformylation and hydrogenation of the aldehyde were rather slow. As expected, the phosphine complex of the cobalt is a more stable, but less active, hydroEormylation catalyst.
To increase the GC and GC/MS sensitivity for alcohols and to increase their retention time, the reaction mixture was treated with a silylating agent. The capillary GC of the resulting mixture is shown by Fi~ e 11.
The GC/MS established that mose of the reaction products were primary alcohols. The only detectable aldehyde components present were minor a~ount~ of n-undecanal and 2-methyldecanal. They are present in amounts le~9 than 5X of the total oxygenated products.
As it is apparant from the figure, the main product of the reaction wa9 the n-Cll alcohol, undecanol. It represents 50X of the total reaction mixture. Thus, only about 50% of the products have branching.
Significant a~ounts of 2-methyldecanol wera also formed. The n/i ratio of these two products was about 10. This means that the hydroformylation of l-decens was highly selective, sinc~ both of these compounds were derived from it. The minor alcohol components could not be identifi2d because of similariti~s in th~ir mass spectra. Baqed on the relatively short GC
ret2ntion time the isomeric C12 alcohols wer~ probably dibranched compounds.

-so-The reaction mixture was also analyzed using packed column CC to estimate the amount of heavies formed. The heavies were only about 0.3X in the residual product. The presence of the phosphine ligand apparently inhibited the formation of the hea~y by-products.
The reaction was stopped after 1~0 minutes. Thereafter, the remaining 1704g of the product catalyst mixture was worked up. The excess phosphine and then the unreacted components were first removed in high vacuo at room tempcrature. However, in the absence of excess phosphine, the remaining product plus catalyst mixture was unstable when heated to 90C in vacuo. Ther~al decomposition was indLcated by a loss of vacuum.
Thersfore, th~ attempted distillation was discontinued and the catalyst was removed fro~ the residue by aqueous acetic acid plus air treatment as usual. The water-organic mlx~ure was diluted with he~ane to facilitate the separation of the organic phase. After the removal of the solvent in vacuo, the residual product weighed 420g. This is about 25 weight percent of the crude reactant mixture. Disregarding the weighe increase of the olefinic reaction mixture during the reaction, the above amount of total oxygenated products corrasponds to the conversion of 25X of the gas oil fraction employed as a feed.
The cobalt free residual product was distilled under 0.12mm pressure. The isomeric undecyl alcohol products were obtained as a clear, colorless liquid distillate between 80 and 90C. The dark residual heavy by-products amounted to about 5X of the total oxygenates.

Examples 23-25 ~ffect of Aging on the ~ytrofor~ylatlon of C16-Clg G~s Oil ~ith Triethyl Pho-~phlno Cob~lt Comple~ at 140~C and 1500 psl The broad cut light gas oil of the previous example was hydro-formylated using the medium pressure procedure in the pxesence of 0.23M
cobalt and 0.72M triethyl phosphine. The reaction was carried out at 180C
using an initial 1/1 H2/CO reactant at a pressure of 1500 psi. The pressure was maintained with a feed gas of 3/2 H2/C0 ratio.
In the first example (23), a rapid initial reaction too~ place.
GC analyses indica~ed that, assuming 50X olefin content for the feed, about half of the Qlefins were hydroformylated in 12 minutes. The ma~or reaction products were the C17-Clg n-aldehydes and 2-methyl aldehydes in a n/i ratio of about 5.

~.2~ 73 In the second example (24), the same feed was used under the same conditions, but after about a month's storage at room temperature, without an antioxidant. No reaction occurred. The cobalt was precipitated.
Testing of the aged feed for peroxide was positive.
In the third example (25), the aged feed was disti.lled in vacuo prior to being used in another hydroformylation experiment under the same conditions. The re~ults with the redistllled feed were about the same as those with the fresh feed of Example 23.

Hi~ Pre~ Hydrofo~YlA~ion o~ C4_to Cl~ N~phtha Fractions in th~ Pr~sance of Cobalt Comple~s ~amples_26-472 The previously described C4 to C12 Tluid coker naphtha containing l-n-olefins as the major type of olefin reactant was also hydroformylated successfully in the presence of cobalt complexes withollt phosphine modifiers at high pressure. Clo and C8 feeds were studied in detail. The reaction conditions used and some of the results obtained are su~marized in Table~

In general, the omission of the trialkyl phosphine modifying ligand from these cobalt carbonyl complex catalysts resulted in greater hydroformylation activity. However, the ratio of n-aldehydes to the 2-methyl branched aldehydes was drastically reduced to values between about 1.9 and 3.2. The cobalt catalysts could be used not only at high, but at low temperatures as w~ll. In the low temperature region of 110 to 145C, the process was selective for the production of these major aldehyde isomers. The rata of olefin isomeri~ation was drastically reduced. The n/i ratio of the products and the amount of aldehyde dimer and trimer products w~re inversely proportional with the reaction temperature.

Example 26 Hydrofsrmylation of a C4 to C12 Naphtha by H2~CO ~lth D~cobalt Octacarbonyl at 150C and 4500 psi The previously described broad naphtha cut was hydroformylated as a 1/1 mixtur~ with hexanc in the presence of 0.2X CO at 150C by an approx-imat61y 55 to 45 miXturQ of H2 and CO at 4500 psi, using the high pressure 7~3 V-_ ~ ~ ~ ~ o _ o ~ CO o r~ o~
o a~
o ~ .~

o ~ .- Ur~
_~ _ ~ ~ ~~ ~ o o ~o o ~ r~U~
c -~ ~ o o C o o o -- o ~
O I N O O C~J O O C~l O C~i O
_ ,a ~ ., 'O

C~ $ N ~ 0 O O O
C~ ~
3~ ~
~" o o ~ oo o o~
, a~ ~ u oo oC o~ ~C~
~ ~ o~ I o 88 oo gc~o g~o ~ U o o o C~ o ~ o ~ o o E~ ~
1~. iit ~.1 G
O Itl li~ O O O O O O O O O O O
g ~t ~ t tn ~ ~7 ~ ~ ~ L t u~

~t ~ O O
E~o ~ ~1 N C`J O N ~N N N ~1 N C~l ,S ~ O O O _ O O O O O O O O
O ~
`~:1 0 ' W ~ ~ u~ ~t ~St a~ ,' at o Cl J ~ 2 Z
_~
--O

O -~t ~ z -- o o :~ ~ 0 a~ o o ~ ~ t _ _ __ ~.
~ O ~ I~ ~ ~ O
:~ Z
!~1 ~2~$73 procedure. The reaction mixture was sampled after 10, 30, 60, 120 and 180 minutes to follow th~ progress of the reaction by capillary GC analyses.
The GC data indicated a long induction period. Up to 30 minutes, no n-l-olefin conversion was observed. For example, the ratio of n-l-decene to n-decane component remai~ed the same. However, thereafter a fast reaction occurred. The GC of the 120 minute sample showed that all the l-n-olafin components were completely conYerted. The ma~or product peaks of the GC are those o the correspondin~ n-aldehydes. The minor, but distinct aldehyde products are 2-methyl substituted aldehydes. The n/i ratio of these major products is about 2.8.
The GC of the final reaction mixture is shown by Fi~e 12. It expressly shows the ma~or Cs to C13 aldehyde products formed and the Cs to C12 n-paraffins. A comparison of the hydrocarbon reglon of the figure with Figure 1 of th~ naphtha feed clearly indicates that on hydroformylation the l-n-olein components were essentially completely converted to provide mainly the n-aldehyde products. Figure 7 also shows that the peaks of the hydrocarbon and sulfur compound components of the feed in the Cg to C12 n-paraffins region overlap with those of the C7 to Clo aldehyde products.
Since the GC retention times of components are approximately proportional to their boiling points, this indicates that thc overlapping components cannot be separated by fractional distilla~ion.

Examples 26 and 27 Hydroformylation of Cs Naph~ha by 1/1 H2/CO
~ lth 0.2X CobAlt at 130C and 3000 psi a~d tha ~ydrogenation of tha C6 Aldehyde Product About 2500g of a broad Cs Flexicoker naphtha fraction with a boiling range (bp.~ of 24 to 34C was washed three times with 1250ml cold 25X agneous NaOH solution and once with distilled water to remo~e the thiol components. Thereafeer it was fractionally distilled using a 22 plate Oldershaw column to obtain hydroformylation feeds free from higher boiling disulfides. The feed composit~ons and the results of two hydroformylation experiments are shown in TablQ~ 9g~
A Cs f ed of bp. 25-28C, cont.aining about 33X l-pentene and 13X
n-pentane, was hydroformylated in the presence of 0.2X Co addPd as Co2(CO)g at 130C by a 1/1 mixture of H2/CO at 3000 psi for 6 hours. The reaction mixeure WaQ periodically sampled fcr packed column and capillary GC
analyses. A highe. boilinz Cs feed of bp. 28-32C, - ~2~73 Table XX
~ydroformylation of Cs Olsfinic Fr~ction~ of Flexico~er Naphtha at 130~C
in the Prcsence of 0.2X Cobalt Catslyst Derived ~rom Co2(CO)g with 1/1 H2/CO at 3000 p3i Components of Total Mixturç
Reaction Pressure mbY Packed Column GC. Z
Example Time Drop Un- Alde- Alcohols Dimers __~Q~ B_reac~ hYdes Form~tes Trimers 26 60 21 92.4 6.2 0.3 1.1 120 160 78.2lg.5 0.3 2.0 180 32 28.965.9 1.2 4.0 240 12 22.568.5 3.8 5.2 27 60 94 93.5 4.8 0.4 1.3 120 188 66.329.0 0.5 4.2 180 20 29.861.8 2.9 5.5 240 8 25.963.5 4.2 6.4 -95- ~. 2~ 3 Co, i ~ un ~ ~ I~ a~ O O~ u~
t, ~ ~ I I o o o o o, o o n ~~ ~ ~ O ~ ~ O
Z Y I ~1 '' o ~ I O ~D ~ O ~ O Ul O ~ Z a o ~ ul '~
O U

~ ~ O ~ E ~ D ~ 0~ O C`l _~ ;; e _ ~ I ~ O C`~
O h ~ C ~ ~ o V
O a ~ c ~ 5' al '' -' O ~ c ~: . U ~ ~ ~
E-~ O C~ N ~ ~ O O O O
0~ ~ ~ ~ C C
C ~ C
i~ c~ ~ 1 C Cl.
V C~ 1~ _1 (~
C CC~ ~ ~ U~ ~ r~ N CO ~0 O CO C~ CL .C
~O U ~ Cl U~ O ~ O ~C ~, " 3 ~ ~ C ~
~ L ~13 0 J
a c cr~ ~ co r~ oo o o o~
C C\ C~ N ~ ~D 1~ 3 0 N 00 r~ ~ I e ~4 1 ~ U ~ I ~ ~N N N U~ ~:1 0 ~ It ~ ~ ~ ~ U C ~ UO S
1 ~ ~ O N u~ ~ C`J. , , . C C~ IJ
O '~:1 C' ~ C~ u~ ~ u~ o ~ Q. 3 U s _~ u ~ a o 'Yl ~ U , _~ O N Cl~ t 3 3 c c ~ ~ ~ 3C~ ~ ~ O O ,~ O _l o ~ r~ u~ C~
4 U~ ~ C`J t~ D N ~ ~C E~
.C . g . ~ ~ 3 S

Sb C~ C, Z; O ~ :~ O O O O O O O C3 0 0 C~
E~ ~ L O O
C ~_1 _ $ ~3 -96~
ontaining about 31X l-n-pentene and 20X n-pentane was similarly converted.
The packed GC data of Table X~ show a selective conversion of the olefinic components to aldehydes. The observed rates of pressure drop indicate that maximum reaction rates were reached betweeen 1 and 3 hours.
By the end of the 4 hour reaction period, the hydroformylation was practically complete and according to GC, the final reaction mixture contained more than 60X C6 aldehydes.
The capillary GC data of Table XXI show the selectivity of the olefin conversion and the isomer composition of the aldehyde products formed.
The change in the distribution of the major hydrocarbon compon-ents of the Cs fe~d indicate that l-pentene, 2-pentenes and the methyl substitueed l-butenes are converted to n-hexanal and the corresponding 2-, 3-and 4-methyl branched pentanals. The 2-methyl-1-butene component is much less reactive and as such is only partially converted under the reaction conditions used.
The main aldehyde product is n-hexanal. According to capillary GC, it is more than 45X of the C6 aldehyde products. The three methyl branched C6 aldehydes, 2-, 3- and 4-methylpentanal, are present in compar-able ~uantities and not completely separated by GC. n-Pentanal and these monobranched aldehydes amount to more than 95X of the reaction mixture.
Slightly more than lX 2-ethylbutanal is present. Similar amounts of 2,3-dimethylbutanal are found.
Sulfur specific GC of the reaction mixture did not indicate any sulfur containing impurities in the aldehyde range. However, there were low boiling sulfur compounds includLng H2S in the feed range.
The reaction mixture was distilled to isolate the products. The C6 aldehydes were obtained betwen 47 and 51C at about 50/mm pressure.
During the distillation, most of the cobalt complex catalyst decomposed and precipitated. Significant aldehyde dimerization and trimerization occurred during distillation as a side reaction. The recoYered Cs hydrocarbon feed was free o~ sulfur indicating that desulfurization by cobalt also occurred during the distillation.
The distilled aldehyde product contained 37.8X n-hexanal, 55.8X
isohexanals, 1.8X alcohols and 4.6X formates according to packed GC. The ~ t 3 reduced percentage of n-hexanal in the distillate compared to the reaction mixture was due to its preferential aldolization over the isohexanals.
The distilled aldehyde washed with lOX aqueous sodium hydroxide solution to remove the small amounts of HCo(CO)4 which codistilled during the separation. The washed aldehyde (1730g) plus 5Z distilled water (86.5g) was then hydrogenated in the presence of a 160g (270~1) CoS/MoS
based catalyst. The reaction mixture was presured to 1500 psi (103 atm) with hydrogen and heated to 130C. The pressure was set to 3000 psi and th~ temperature was increased by 10C every hour. Once the temperature reached 160C, it was kept there for the total reaction time of 20 hours.
Subsequent capillary GC and 400 MHz lH NMR analyses indicated that essentially all aldehydes were hydrogenated to the corresponding alcohols.
Capillary GC indicated that 38.4X of the C6 alcohols formed was n-hexanol.
According to packed GC, dimeric and trimeric by-products were formed in comparabla amounts. They amounted to about 15X of the reaction mixture.
No paraffin by-products were observed. Sulfur specific GC detected no sulfur.
A 2 to 1 mixture of the crude, C~ alcohol product was ~ashed with a lOZ agneous solution of NaOH and then with water. After drying over MgS04, the alcohol was distilled to recover the hexanols as a clear, colorless liquid mixture between 109 and 115C at 200 mm. The n-hexanol content of the distillate product was 35.8X The dimer by-product was distilled at 12 mm. It was obtained as a colorless liquid between 103 and 113C. Capillary GC indicated that it contained isomeric C12 aldol alcohols. The distillation residue was mostly the trimer presumably formed from the aldol adduct of the aldehyde via the Tischenko reaction:
C4H9CH2CH-CHCH ~ C4HgCH2CHO ~ C4HgCH2CH-CH-02CCH2CH2C4Hg OH R OH C4Hg + C4HgCH2CH-CHCH20H
C4HgCH2CH20CO C4Hg Similar trimerization side reactions occur with the other aldehyde products of the present process.

Examples 28 and 29 Hydroform~l~tion o~ C6 Naphtha by 1/1 H2~GO vith Cobalt at 130 and 150C
under 3000 psl ant the Hydrog~nat~on o~ tho C7 Aldehyd~ Pr~duct A heart cut C6 Fluid-coker naphtha of bp. 56-65C fraction was used as a feed for hydroformylation. It contained about 42X l-hexene. Its detailed composition was previously discussed and given in Figure 3. The reactions wsre carried out with an equimolar mixture of H2 and C0 at 3000 psi, (206 atm). About 2000g of the feed was used per run.
As a catalyst C02(CO)g was added in benzene solution. In the first run, the cobalt equivalent of the catalyst was 0.4% and the reaction temperature 130C. In the second experiment, 0.2Z Co was used at 150C. Rapid olefin conversion was observed in both experiments.
Analyses of periodic samples of the reaction mixtures by packed colu~n GC are shown by Table XXII. The pressure drop data of the Table indicate that the hydroformylations were essentially complete in both experiments in 180 minutes. By that time, the percentage of unconverted hydrocarbons in the reaction mixture was reduced close to the mini~um in the 30X range. The combined percentages of aldeh~de plus some alcohol and formate ester products reached a maximum in about 180 minutes. More aldehydes (62.3X) were obtained at 130C than at 150C (53.6%). The significantly reduced aldehyde concentration after 360 minutes (52.1 and 37.6X, respectively) is clearly due to dimer and trimer formation (19.4 and 33.4Z) respectively.
Capillary GC provlded an effective separation of the volatile isomeric components o~ the reaction mixture. Most of the isomeric C7 aldehyde, C7 alcohol and C7 alkyl ~ormate ester products could be identi-fied by a combination of capillary and mass spectrometry (~S). In the capillary GC of the isomeric aldehydes product, shown by Figure_13, all the aldehydeg which can be derived from linear hexenes and four of the aldehyde-~ deri~ed from monobranched heptenes were separated and identified.
The reaction schemes of the presumed hydroformylations leading ~o the various heptanol isomers are shown in the following:

CH3CH2CH2CH2CH-CH2 ~ CH3CH2CH2CH2CH2CH2CH0 : Normal CH3CH2CH2CH-CHCH3 - ~ CH3CH2CH2CH2CHCH0 : 2-Me -CH3CH2CH-CHCH2CH3 - ~ CH3CH2CH2CHCHO : 2-Et ~$~ 3 - Table XXII

COMPOSITION BY PACKED COLUMN GC OF PERIODIC SAMPLES OF REACTION MIXTURES
OF ~HE HYDROFORMYLATION OF A C6 OLEFINIC DISTILLATE FRACTION OF
FLUID-COKER NAPHTHA BY H2~CO AT 3000 PSI (20680 kPa) Pressure _ _ GC Composition~ X Alde-No. Dropa hydes, Temp. Ti~e ~i Hydro Ald~hydes Dimers n/i _Cat. Min. ~ carbonsb nC 1 Trimers Ratio 28 10 94 97.0 1.4 0.8 0.8 0.56 77 45.127.0 24.1 3.8 1.12 130C 60 32 39.729.6 27.8 2.g 1.06 0.4XCo 180 11 30.331.5 30.8 7.4 1.02 240 8 29.629.6 29.9 10.9 0.99 360 - 28.525.0 27.1 19.4 0.92 29 10 118 89.0 3.7 3.5 3.8 1.06 273 74.612.1 9.2 4.1 1.31 150C 60 201 42.526.4 ~4.4 6.7 1.08 0.2~Co 180 11 28.726.4 27.2 17.7 0.97 240 10 26.323.8 25.1 24.8 0 95 360 - 29.020.0 17.6 33.4 1.14 ... . ~

aPressure drop while supply of additional H2/CO is shut.
bUncon~erted feed components.
CThe percentage of n-aldehydes also includes alcohols and formate esters.

CH3CH2CH2C-CH2 ~CH3CH2CH2CHCH2CHO : 3-Me ., CH3CH-CHCHCH3 ~CH3CHCH2CHCHO : 2,4-DiMe CH3cH2cHcH-cH2 ~CH3CH2CHCH2CH2CHO : 4-Me Capillary GC also indicated the presence of minor amounts heptyl alcohol and heptyl formate secondary products. The main isomers were ehe normal heptyl and 2-methylhexyl derivatives derived from normal heptanal and 2-methylhexanal as indicated by the following reaction scheme:

CH3(CH2)5CH20M CH3tCH2)3CH(CH3)cH20H

. I H2 CH3(CH2)3CH--CH2 - ----~ CH3(CH2)5CH + CH3(cH2)3cH(cH3)cHo ~ H2/CO ¦ H2/CO
CH3(cH2)6ocH CH3(CH2)3cHcH20cH
Il 1 11 n-ROCH 2-MeCH20CH
Il li O O
, In case o the alcohol by-products only normal heptanol and 2-methylhexanol, the two main isomers, were identified. However, all ehe significant isomeric heptyl formaee by-produces could be recognized since their GC peak patterns werç the same as those of the corresponding aldehydes.
Analyses by capillary GC of samples periodically taken from the reaction mixtures of 130 and 150C hydroformylations provided detailed ~nformation on the progress of the reactions and side reactions. The data obtained are sum~arized in Table XXIII.
In general, the capillary GC results also indicate that the primary reaction, i.e. hydroformylation, was essentially complete in 180 mlnutes. In this period, the hydrocarbon content of the mixtures decreased to about 35X. Deter~ination of the concentrations o f l-hexene and 3-hexene reactants relative to that of the unreactive 3-methylpentane component indlcated rapid olefin conversion. l-Hexene conversion was essentially complete within one hour. 3-Hexenes convsrsion took about three hours. In that period 2-methyl-1-pentene was also reaceed. The residual olefin content of th~ hydrocarbon feed after 3 hours was about 5X. Thus the total olefin conversion is about 92X.
Determinations of the 3-hexene concentrations in the 150C
reaction mixtures indicated a slight increase rather than decrease during the first hour. This increase is apparently due to the isomerization of l-hexene to internal hexenes concurrent with the hydroformylation. Olefin isomerization is much reduced at 130C.
Hydroformylations were continued at both 130 and 150C for a total reaction time of 6 hours (360 mln.). During the last 3 hours largely secondary reactions took place. The concentration of formate esters more than doubled. Formates were 7.1% of the oxygenated products at 130C and 8.2% at 150C. The alcohol concentraiton decreased from 3.2 to 1.7X during the last 3 hours at 150C. This was apparently due to the ~ormation of heavy by-products not observable by capillary GC.
Table XXIII also shows the percentage distribution of the main aldehyde products in the reaction mixture. After 3 hours reaction time, the maln aldahyde products amounted to 93.lX of the total oxygenates at 130C and 95.7X a~ 150C. The n-aldehyde component was 36.2 and 31.2~, respectively. As sxpected the n-aldehyde conc0ntration decreased with increasing reaction time. Mor2 and more of the internal and branched olefin components reacted to form other branched aldehydes.
Table XXIII separately lists the percenta~e in the oxygenates of ths three iso~eric heptanals, i.e. n-heptanal, 2-methylhexanal and 2-ethylpentanal. These aldehydes are derived from linear h2xenes as it was ~ ~ o~ ~ S
æ c:~ ~ s~ c~ o o _i , ~ ~ ~ U
o CO ~ ~
~i ~ ~ C I U~ O
o o ~ ~ ~ O o o ~ ~ 'O u~ I I ~ o ~, o æ~ C O O OO OC, O O O O
~o 0 u O o 3; ~ 3 I ` 1 ~ ~ ~ ~ ~ ~ ~
~ O a~ ~1,~ ~ ~ ~ ~ ~ ~

V ~ N 00 ;S ~ rf ~ ~d aJ
U ~ C
O C~
~X ~ ~ ~ O O C~ C
O ~ ~1. .C c~ D X
_l ~: co o o~ ~ ~ ~ n 1~ ~ ~ ~ 1 ¢ ~ o . u~
~ ~ C7 ~
O ~ ~ I~ ~ X ~ ~ ~ ~ ~ ~ C
u~ u 3 ~ ~ ~ ~ 0 ~ ~ " , C
~ ~ X V~ ~C
O a~ ~
o ~ o~ I ~ o ~ Z ~ C
o o o O O ~ _I ~ ~1 VD N ~1 ~ . o C
Q.
E~~ h ~ . ~ ~ ~ O _~ o u ~a ~ o Z C~
æ~ cl O ~ O O`~ ~oO O O O O O ~ O

O ~ ` O
X z ~ O U

-103 1~ 73 praviou~ly ~hown by the rea~tion schemes. Their combined percentage after 3 hou~ i9 57.3X at 130C and 52.6X at 150~C. It is noted that these per-centaga~ are way balow the percentage of linear olefins in the total identified olsflns of the feed (86X). This and particularly the lower than expected perc~ntags of ehe n-aldehyde component are due to the presence of signlficant amounes of unidentified methylcyclopentenes in the feed and the preferential convçrsion of the primary normal aldehyde iso~er products to higher boiling secondary by-products.
The percentage distribution of identlfied lower boilin~ oxygen-ated compounds is shown in TablçL~ V It is noted that in this table the su~ of the aldehydes derived from linear hexenes (Normal, 2-Me and 2-Et) in 180 mlnuta~ is 65.6X at 130~C and 61.lX at lSO'C. These increased percent-age~ ~re due to the exclusion of cyclic C7-aldehyde products from the calculations.
The primary C7 aldehyde products of the hydroformylation of C6 olefinic coker distillate feed and the secondary products derived ~ia ehe condensation of these aldehydes were separated by distillation for further studies. At first, the two reaction mixtures resulting from the hydroformylation of the C6 olefinic naphtha fraction at 130 and 150C were separately decobalted with aqueous acetic acid plu5 air treatment as usual.
Neither precipitation nor separation prob~ems were encountered. In case of the 150C reaction mixeuro, the percentage of dimers plus trimers was reduced from 33.4~ to 18.5Z duri.ng decobalting, probably due to acetal and ester hydrolysis.
The decobalted yellow liquid reaction mixtures (1834 g and 2288 e) wero fractionally distilled under atmospheric pressure using a two foot packod column.
Mose of the unconverted hydrocarbons were removed from the reaction mixtures as colorless liquid distillates boiling between 55 and 65C, u~ing a heating bath of 120 to 130C. Thereafter, the aldehyde produots were distilled at reduced pressure. Th~ aldehydes from ~he 130C
reaction were distilled at 29 to 31~C under 0.2mm. Those of the 140C
hydroform~lation were received between 30 and 50C at lOmm pr~ssure. The heating, particularly during the at~ospheric distillation, resulted in the formation of additional amounts of dimers and trimers. In the cas~ of the 130C reaction mix~ure, the heavy by-products increased by 60% (from 18.9 to 31.9X) during dlstillation. In tha case of the 150'C reaction mixture, the incre~se was 82X (fro~ 18.5 to 33.6~).

-104- ~ 3 Table XXIV

OF BILI.INGS FLUID-COKER NAPHTHA TO VARIOUS ISOMERIC C7 ALDEHYDES

_ Capil~GC Con~osition of Mai~ Products and Bv-P~ducts. X
Aldehvdes Alcohols Formates Temp. Timc 2,4- 2- 3- 2- 4- 5- Nor- 2- Nor- 2-Nor-Cae, ~ ~ ~ Me Me Me Me mal ~ mal Memal 130C 10 0.9 4.39.0 17.3 6.4 4.857.3 - - - -0.4X Co30 2.3 ~.29.1 20.9 7.5 4.047.4 0.3 0.40.3 0.6 2.2 6.99.9 19.8 8.2 4.646.1 0.4 0.60.5 0.8 180 2.5 6.511.9 17.6 9.0 5.941.5 0.6 1.01.4 2.1 240 2.4 6.412.2 17.3 8.9 5.940.:20.7 1.21.9 2.8 3~0 2.5 6.812.5 17 2 8.5 5.837.7 0.6 1.52.8 4.1 150C 10 2.0 7.57.5 24.2 6.0 4.047.2 - 1.0 - 0.6 0.2X Co30 2.2 7.66.9 24.8 6.6 3.547.1 0.2 0.7 - 0.4 2.4 7.38.5 21.7 7.9 4.345.4 0.6 0.90.4 0.6 180 2.7 6.611.0 17.8 9.0 5.83~.7 2.7 3.71.7 2.3 240 2.7 6.510.7 16.7 8.4 5.533.6 4.3 5.92.4 3.3 360 2.9 8.011.5 18.1 7.7 5.030.7 2.5 3.64.3 S.7 1 Z~4~373 The distilled C7 aldehyde products of the two hydroformylation runs were combined to provide 1353 g (11.87 moles) aldehyde intermediate.
The composition of this combined aldehyde product with the distribution of the various heptanal isomers i9 provided in Table XXV.
The heptyl formate rich distlllate fractions were also combined to providc 396 g of another intermediate for hydrogenation. This colorless liquid fraction contains about 50Z formate ester. Its boiling range is from 43'to 65C at lOmm. The rest are aldehydes and alcohols and their condensation products. Thus the product is equivalent to about 3 moles.
As such, it is the calculated amount for the hydroformylation of 252 g (3 moles) of hexene. The isomer composition of this formate rich product is comparable to that of the aldehyde in Table XXV.
The aldehyde product consists of about 90X aldehydes, 9X formates and l~ alcohol. Aboue 82.6Z of all the components of the aldehyde fraction were specifically identified. Of the total identified aldehydes, 35.2X is n-hep.tanal. The n/i ratio of aldehydes is 0.54. Most of the iso-aldehydes are monobranched C7-aldehydes. Ony one dibranched aldehyde, 2,4-dimethyl-pentanal, was found. On an average the aldehyde product mixture contained0.63 branches per molecule.
The higher boiling for~atc rich fractlon (E-7170-II) could be only partially analyzed by capillary GC. The high boiling dimer and trimer components were not eluted from che capillary column. According co packed column GC, th~ formate fraction contained small a~ounts of dimers (about 2X), but large amounts of trimers (about 30X). The relative percentage of formate.c among thes0 components is 60X. It is interesting to observe that th~ alkyl carbon backbones of these isomeric heptyl formates correspond eo those of the primary aldehyde products.
After distilling off mose of the aldehydes, the residual reaction mixtures were combined and their fractional distillation continued at 10~
using an 18 in. packed column. Further distlllate fractions containing increasing amoun~s of C7 alcohol and heptyl formate secondary reaction products were obtained betwen 50 and 66C at lOmm. During the distilla-tion, thermal decomposition of ;he formaee esters occurred to an increasing degree as the temperature of the heaeing bath increased to 150C. Thus the last small distillate fraction ~11 g) consiseed of 70X h~ptanols, 12X
dimers and 18X trimers.

.2~ 3 Table XXV
COMPOSITION OF THE C7 ALDEHYDE PRODUCTS AND FOR~ATE BY-PRODUCTS DERIVEDFRO~ A C6 OLEFINIC DISTILLATE FRACTION OF FLUID-COKER NAPHTHA

Compos~tion bv Ca~lllary CC
Aldehydes _Formates ~ of % of Z of Seq. Nams Designatlon Total Listed Listed ~_ ~ ~~Qmpo~entP~--ÇQ~mg:~D~ Cm~, C~p~s. C~pds.
1 2,4-Dimethylpentanai 2,4-Di-Me-CH02.22 13.50 18.10 2 2-~hylpentanal 2-Et-CH0 8.07 9.77 0.68 3 3-Methylpentanal 3-Me-CH0 11.15 13 50 1 99 4 2-~&thylpentanal 2-He-CH0 15.50 18 76 2 53 5-Methylpentanal 5-Me-CH0 6.74 8.71 1.23 6 4-M~ehylpentanal 4-Me-CH0 4.55 5.52 0.90 7 n-Heptanal n-C6-CH0 26.20 31.72 8.98 8 2-Methylhexanol 2-Me-CH20H 0.45 5.39 10.82 9 n-Heptanol n-C6-CH20H 0.31 0.38 12.53 2-Ethylpentyl formate 2-Et formate0.10 0.12 0.67 11 3-Methylhexyl formate 3-Me formate0.51 0.62 2.88 12 2-~ethylhexyl for~ate 2-Me formate2.93 3.55 18.10 13 5-Methylhexyl formate 5-Me formate0.80 0.97 5.72 14 4-Methylhexyl formate 4-Me formateO.46 0.56 3.78 n-Heptyl formate n-C7 formate2.59 3.14 28.90 16 Total aldehydes R-CHO's 74.43 90.13 16.60 17 Total alcohols R-CH20H's 0.76 0.92 23.35 18 Total formates R-CH202CH's7.39 8.95 60.05 19 Sum Totsl 82.58 lOO.OO

aThe h~gh boLling ~dimer and trLmer" components were not eluted from the capiIlary GC colu~n. According to packed column GC there were about 2i ~ b-u~ 301 ~rim-ri pre~en-After the decomposition of the formate ester by-products, the rest of the residue was distilled using a one foot column at O.lmm. About 640 g of a distillate consisting of about 70X dimer and 21Z alcohol was obttained as a clear, pale yellow liquid boiling between 55 and 74C.
Thereafter, a trimer rich fraction (80X) and a tetramer rich fraction (58X) were also obtained as yellow liquid distillates. The "trimer" fraction distilled at 130 to 132C at O.lmm with decomposition.
The final distillation residue was only 116 g, i.e. 2.8Z of the starting reaction mixtures. However, the recovery of pure distillate products was poor, due to the concurrent decomposition of the ormates and heavier by-products. The present results suggest the hydrogenation of the comple~e reaction mixture immediately subsequent to cobalt removal. This should result in much improved recovery of the desired hepanols.
The sulfur compound components of C6 olefinic coker naphtha feed and the hydroformylation feed were also studied primarily to determine sulfur distribution according to boiling point. The total sulfur contents of feed plu~ selected product and by-product fractions is shown by the fol~lowing tabulation:
Boiling Point Main Component Sulfur Ç/mm ~X) ppm 56 - 65 ~760 Hydrocarbon Feed (100) 640 55 - 59 /760 Recovered Hydrocarbon (99) 575 31 /02 Aldehyde (85) and Higher 92 58 - 65 /10 Formate (46) and Lower 123 55 - 57 /0.1 Dimer (82) and Lower 1740 130 - 132 /0.1 Trimer (80) and Lower 3380 - DistillatIon Residue 596 The data indicate that some of the sulfur compounds of the feed are con-verted to high boiling compounds. The aldehyde product has a relaeively low sulfur content.
An investi~ation of the sulfur distribution by GC/MS showed that the thiol components of the feed were lar~ely con~erted to H2S and to high boiling sulfur compounds while the thiophene component remained mostly unconverted. The main sulfur compound impurity in the aldehyde fraction was thiophene, due to poor separation.
Thb sulfur containing compounds in the dimer fractions were thiolheptanoic acid propyl and butyl esters. It is assumed that the propyl ester was derived by the reactisn of the propanethiol component of the fesd 9~

with the C7 aldehyde product 2 C6H13CH0 + C3H7SH ~ C6H13cos~ + C6H13CH0 The corresponding butyl ester could have been derived via butadiene derived from ~hiophene according to the following hypothetical set of reaction equatLons:

S + 2H2 ~ CH2~CH-CH-CH2 + H2S
2 C6H13CHO ~ ~ C6Hl3C2H + C6H13CH2H
C6H13C2H + H2S ~ C6H13CSH ~ H20 C6H13CSH + CH-cH-cH-cH2 -- C6Hl3coscH2cH--cHcH3 l H2 C6Hl3coscH2GH2cH2cH3 The above hypothesis is supported by a model experiment. A 9 to 1 mixture of l-hexene and thiophene was hydroformylated under the pre-viously used conditions Ln the presence of 0.02X cobalt in a highly exothermic reaction between 140 and 185C. GCjMS studies indicated that 5X
of the thiophene was converted to bu~yl thiolheptanoate and propyl thioheptanoate.
The main sulfur containing components of the trimer fraction were found to be diheptyl sulfides. These were presumable derived from the heptanal products as indicated by the following hypothetical sequence of r2actions:

C6H13CH0 ~ [c6Hl3cHs] '- C6H13CH2SH

C6Hl~cH2sH - ~ C6H13CH(OH)Scx2c6Hl3 ~ H2 C6Hl3cH2scH2c6Hl3 Whether the above hypotheses of the course of sulfur compound conversions are right or wrong, the present exa~ple demonstrates tha~, in the present process, the sulfur containing impurities of the feed are partially - log- ~ 73 converted to high boiLing thiol esters and sulfides rather than sulfur com-pounds of the aldehyde boilLng range. Thus aldehydes of low sulfur content can be isolated by frational distillation.
The aldehyde product was hydrogenated as a 2 to 1 mixture with toluene in the presence of 5X water and lOX CoS/MoS based catalyst at about 160C under 3000 psi for 20% hours. Probably as the result of the high temperature employed, significant dimerization occurred. According to packed colu~n GC, the distribution of the oxygenated components of the final reaction mixture was the following: 56% alcohols 39% dimeric aldol alcohols and 5Z trimers, Sulfur specific GC indicated no sulfur in the alcohol range, but H2S and high boiling sulfur compounds in the dimer range.
The crude alcohol product was further diluted with toluene to produce a 1 to .1 mixture. This was washed with lOX aqueous sodium hydroxida and then with water to remove the H2S and other acidic impuri-ties. The resulting organic phase was then fractionally distilled using a 24 plate Oldershaw column. The heptanol product was obtained as a color-less, pleasant smelling liquid between 98 and 103C at 55mm, The dimeric aldol alcohol distilled in the 74 to 99C range at 3mm. The trimer by-product remained as the distillation residue, No sulfur could be detected by GC in the alcohol product, but minor sulfur impurities were noted in the dimer, Tha heart cut heptanol product containing 22X n-normal isomer was converted to semilinear diheptyl phthalate which was evaluated as a plasticizer.

Example 30 ~ydroformylation of C7 Naphtha by 1/1 ~2~CO
with O.2X Cobalt at 13~C under 3000 psl ant the Hytrogsnatlon of the CR Aldehyde Product A broad cut C7 Fluid-coker naphtha was redistilled to provide an olefin enriched oxo-feed, The narrow fraction of a 15~10 distillation, boiling betwen 88 and 94C, was utilized, It contained about 6.5X
2-methylheptene, 30X l-n-heptene, 12X n-hepeane, 4.3X trans-2-heptene, 2,8X
cis-2-heptene. Only small amounts of aromatic hydrocarbon were present:
O,lX benzene and 0,5~ toluene, A sulfur specific GC of this distillate prior to use as a feed indicated thae some of the sulfur containing components were converted to hign boiling compounds, The hydroformylation of the above feed was carried out using a to 1 mlxture of H2/C0 under 3000 psi at 130C with 0.2~ Co catalyst, added as a solution of Co2(CO)g in the feed, in the manner described in the previous example. A maximum rate of pressure drop was observed about an hour after the start of the reaction. The conversion of l-heptene was essentially complete in two hours. The reaction was completed in four hours. The reaction was highly selective to aldehydes. The distribution of the various types of components of the final reaction mixture by packed column GC was the following: 33.7X unconverted C7 hydrocarbons, 59.lX C8 aldehydes, 4.2X C7 alcohols and C7 alkylformates, plus 3X dimers and trimers. Capillary GC provided the following isomer distribution of C7 aldehydes: 43.8Z n-octanal, 11.7X 2-methylheptanal, 8X 3-methylheptanal, 5.8X 2-ethylhexanal and 1.7X 2-propylpentanal. A sulfur specific GC of the reaction mixture indicated the presence of H2S, some volatile sulfur compounds in the C7 feed range and minor non-volatile sulfur compounds in the dimer range. There were no measurable sulfur compounds in the aldehyde range.
The hydroformylation reaction mixture was decobalted by aeration with hot aqueous acetic acid. Thereafter, the cobalt free mixture was hydrogenated without prior removal of the unreacted C7 hydrocarbons and volatile sulfur compounds. Two hydrogenation experiments were performed in the presence of lOX CoS/MoS based catalyst and 5X water under 300 psi (206 atm) pressure. The first experiment was carried oue at 150C. After 20 hours only about half of the aldehydes were reduced. Thus the reaction was completed in 40 hours. The second experiment was carried out at temperatures increasing from 130 to 160C in four hours in the manner described in Exampla 26 and 27. After an additional 16 hours at 160C, the hydrogenstion was complete.
Hydrogenation at 150C resulted in the formation of major amounts of dimers. The ratio of C8 alcohols to the C16 aldol alcohols ln the reaction mixture was 64 to 37. There was also significant C8 paraffin formation as indicated by the 78/22 ratio of C7 and C8 hydrocarbons. In contrast, the better controlled, variable temperature hydrofor~ylation was highly selective: The ratio of C7 alcohols to dimer alcohols was 91 to 9 and the C7 to C8 hydrocarbon ratio 92 to 8. The n-aldehyde reactants were preferably dimerized. Hydrogenation at 150C produced C8 alcohols coneain-ing 36.4X n-octanol, while the more controlled variabls temperature reaction gave a C8 alcohol con~aining 40.3X normal isomer.

~2~ 73 The crude C8 alcohols were washed with lOX aqueous NaOH and ~hen with water. Thereafter, the mixture was fractionally distiled using a 22 plate Oldershaw column. The C8 alcohol product was obtained as a clear, colorless liquid between 81 and 87C at 13mm. It contained 33X n-octanol No sulfur could be detected by GC. The compound was converted to a semilinear dioctyl phthalata plasticizer.

Examples 31 to 34 Hydrofor~ylation of Broat and Narro~ Cu~ C8 ~aphtha Fractions With and U~thous ~rio~ Caustic Trea~ent in the Pro~enc~ of Cobalt ~ith 1/1 H2/CO a~ 300~C and 3000 psi The broad and narrow cut C~ Flexicoker naphtha distillate fractions, described earlier by Table IX and Figures 4 and 4, were utilized as oxo-feeds. Half of each of these feeds were extracted with a 30X KOH
solution in methanol containing 2X water to remove the thiol components.
The caustic treated fractions were then washed with water and hydroformylated under the same conditions as the untreated fractions. The synthesis gas reactant was a 1 to 1 CO/H2 mixture. The reactions were carried out at 130C (266F) at 3000 psi (207 atm).
As a catalyst precursor, Co2(CO)g was added as a 6X toluene solution under reaction conditions. The catalyst addition was mostly in increments providing O.lX cobalt to the reaction mixture. The occurrence of hydroformylation was tested by shutting off the synthesis gas supply and observing the rate of pressure drop. In general, some pressure drop was always observed on catalyst addition, but it was not sustained if the amount of cobalt was insufficient. In such cases, an additional O.lX
cobalt was added every 60 minutes until a sustained reaction resulted.
With sufficlent amounts of cobalt, the pressure drop increased during catalyst preforming and then gradually decreassd as thç olefin reactants were depleted.
The reaction mixtures were periodically sampled and analyzed by packed column and capillary gas chromatography (GC). The pressurs drop and packet column GC data were used to estimate reaction raees, feed conversion and overall selectivity to aldehydes, alcohols plus i`ormates and aldehyde dimers plus trimers. The results obtainad by capillary GC are summarized in TAble XXVI.
Table XXVI shows that the four feeds e~hibited increasing reactivitle in this order: untreated broad, caustic washed broad, -112- ~2~ 73 untreated narrow and caustic washed narrow. For these four feeds the minimum effective concentration of cobalt catalyst was 0.4, 0.4, 0.3 and 0.2X, respectively. Although the effective catalyst concentration was 0.4Z
cobalt for both the untreated and the caustic washed feeds, the caustic washed feed was much more reactive.
The packed GC data of Tabl Q X~I show that the GC percentage of the total oxo-products by the end of the reaction ranged from about 39.3Z
to about 58.6Z. Since the GC response factor for aldehydes is about 1.3, these GC percentage correspond to 45.7 wt. X and 64.8 wt. Z oxo-products, respectively.
The percentages of oxo-products derived from the narrow olefinic feeds are hi~her than those from the broad feeds. This is expected based on the different olefin reactant content of the two types of feeds. It is unexpected that the selectivity to total aldehydes is also higher in case of the narrow feeds. It is believed that the decreased amount of by-products in the reactlon mixtures derived from the narrow feeds are due to the lower percentage of cobalt catalyst used. The selectivity was clearly the highest in the case of the narrow, caustic treated feed where the lesst catalyst was employed.
The composition of the hydroformylation reactlon mixtures was further studied using a combination of capillary gas chromotography and mass spectrometry. The main aldehyde products were identified primarily on the basis of the characteristic MS fragmentation patterns involving the McLafferty rearrangement. The formation of the isomeric Cg aldehyde isomers found from tho C8 olefin isomers of the feed by hydroformylation is outlined by the following reaction schemes:

i73 Table XXVI
The Con~crsion of th~ Olefinic Componen~s of C8 ~ractlons of Fluid-Coker Naphtha by Hydro~or~ylation with 1/1 H2/CO at 130C and 3000 pgi in ths Pre3Qnc~ of Cobalt Catalyst D2rlvot fro~ Co2(CO~a Concentration in the Reaction Run No. Effect- PMixeure by Packed GC. X a and Total ive Co Drop,bUn-Feed Time Time Conc. psireacted Alde- Alco- Di-Used Min.Min~, ~ X ~ eedC hydes ~Q~ ers 120 150 0.2 1.4 93.2 1.9 3.6 1.3 Untreated 180 100 0.3 2.5 93.2 3.1 3.1 0.6 Broad 240 0 0.3 1.2 92.3 3.7 3.3 0.7 300 60 0.4 89.4 6.2 3.6 0.8 450 210 0.414.3 61.228.9 5.2 4.7 530 290 0.~ 57.333.7 5.4 3.6 0.2 2.5 93.2 2.5 3.1 1.2 Broad 120 0 0.3 3.6 92.4 3.6 3.3 0.7 Caustic180 60 0.4 9.1 88.9 6.5 3.9 0.7 Washed240 120 0.453.2 72.219.4 5.7 2.7 300 180 0.4 6.0 59.628.2 5.8 6.4 450 230 0.4 - 50.437.0 7.1 5.5 530 310 0.4 - 49.836.9 8.1 5.2 0 0.2 5.0 95.5 3.0 0.4 1.1 Narrow120 60 0.343.8 55.839.8 1.6 2.8 180 120 0.316.6 48.847.0 2.4 1.8 270 210 0.3 5.0 42.750.4 3.4 3.5 350 290 0.3 - 41.451.5 4.2 2.9 0 0.1 7.0 96.9 2.9 0.1 0.1 ~larrow120 60 0.243.2 86.313.4 0.2 0.1 Causeic180 120 0.240.6 61.636.1 1.6 0.7 270 210 0.211.1 50.146.3 2.0 1.6 350 290 0.2 7.0 46.750.8 1.8 0.7 .
aGC concentrations are recorded as such; no conversion factors were used.
bThe rate of pressure drop was observed while H2/CO supply was shut off fro~ the r~aceor. CBenzene solvent for the catalyst was excluded fro~ the calculation~. dAlso includes formate esters.

f~3 Abreviation C3H7CH2CH2CH~CH-CH2 ~ C3H7CH2C~2CH2CH2CH2CH0 n-C3H7CH2CH2CH;~CHCH3 ~ C3H7CH2CH2CH2CHCH0 2-Me ~ C~3 C3H7cH2cH-cxcH2cH3 ~ C3H7cH2cH2cHcHo 2-Et \~ C2H5 C3H7CH;CHCH2CH2CH3 ~ C3H7CH2CHC~0 2-Pr (CH3)2CHCH2CH2CH2CHCHO2,6-Di-Me (CH3)2CHcH2cH2GH-CHCH3 CH3 (CH3)2CHCH2CH2CHCH02,5-Et,Me ~ I
(cH3)2cHcH2cH-cHcH2cH3 C2H5 (CH3)2CHCH2CHCH0 2,4-Pr,Me The linearities of the C8 aldehyde products derived from the two pairs of C8 feeds are described by the capillary GC data of Table_ XXVII.
As expected, Cg aldehydes containing a higher percentage of normal nonanal wera derived from the narrow feeds than ~rom the broad feeds. The final n-nonanal percentages for the narrow feed derived products are 37.2 and 40X. From the broad feeds final products containing 28% and 23.7Z
n-nonanal ware derived. The difference between the products derived from untreated and caustic washed feeds of the same hydrocarbon composition is due to their different degrees of conversion.
With increasin~ conversion, ehe linearity of the products is de-creasing. At first higher amounts of n-nonanal and 2-methylheptanal (2-Me) are formed from the most reactivç I-n-octene feed component. As more and more of the less reactive internal and branched octenes are hydroformylaeed, the ratio of normal to iso-aldehydes is decreasin~. In case of the broad feeds, the final n/i ratio is 0.39 and 0.31. Due to the higher percentage of l-n-octene in the narrow feed, the final n/i ratio of aldehyd~s there is 0.59 and 0.67.

~2~ 73 --I ,5 O ~ C ~ coo r~
~ ~ ~, ~ o ~ ~ ~ ~ ~ ~ ~ U~ ~o ~ ~ ~ ~ ~
o ¢ ¢ ~
.~ ~U c x o ~ ~ ~~1~ ~ ~ ~ ~ ~ ~ ,., o ~o~ o c u o v R r ~ ~ ~ ~ ~ ~ ~ ~ ~ P~ ,~
¢ IC:~ ~ ~ ~ w r~ ~ U7 o ~ 1~ .. o O O O O O O O O O O O _~ O O O _ ~rl ~ r ~ ~ 50 ~ O ~t O o~ o 3~ r~ ~ :~
u~ C u~ cr. o U ~ ~ C~
^u OOO OOOo OOO_~ OOOO

O C~ I ~ ¦ ~ ~0 It') ~ 1~ ~ ~ N C`~ ~ N ~ l ~ 5 C
g ~~ til C~ N ~ c~ ~ a ~ o CO ~o U C~ o 1~ C~ o ~ ~ ~
c~l o ~ ~ o ~ ~ ~ ~o ~ ~ ~ I~ ~ o u !~ ~DU ~ Q~ ~ ~ I~ ~ U~ ~ ~ Ir~ ~ r~ ~ ~ ~ ~D ~ ~q h 5 ~ o cO u7 ~ ~5), C ~J) O Cl O O CO CO O ~ r~ O O C~ O O _f O .,~
y O QO .~? c~l ~1 ~ ~ t~ 5 C n --I U ~ N ~ J ~ ~ C~- ~ O X
U c~ ~ O O ~ l ~ C _ ~
o ~~ o ~ ~ qJ o c _~ql V c~ r ~ co ~ C oJ L~ o ~ O ~ C ~ ~ O tO ~ 0 h ~ ~ ~
o ~ ~ C o ~ O O C~ ~ ~ O ~ ~ O ~ ~ 5 ~ O ~ ¦ `O ~ CO 1` '1 11i t`~ ~1 CO 1~ ~ ~'- O ~ ,C 4~ C

O O O O O O O O O O O O O O O O O O O , 'U C o U
3 ¦ cJ C~ N N ~ 1 N h C~l h ,C

~ a U ~ ~ I ~ ~ U~ ~0 ~ I ~ O u~ ~ ~ N CO 1`~ Ul O O O S C O

0~ Z ~ U ~ o ~ N 1`'1 ~I C`l ~ U

~ O ~0 ~ I O O O O O O O O o o ~ C .~: C

l.J ~ ~ ~ 3 3 t'~ ~ _C~
h f~8 aJ h ~ ~ C _ O
a~ 3 æ z o 3 ~ ~ _ U
U t~ U
~ ~ h . --Among the minor branched aldehyde components of the reaction mixture, significant quantities of 2-ethylheptanal (2-Et: 3.75 to 5.09X) and 2-propylhexanal (2-Pr) are formed. (The percentage of 2-propylhexanal includes the GC response for an overlapping peak.) These two aldehydes and 2-methylheptanal plus-n-nonanal are all deri~ed from linear octenes.
The two ~ajor cyclic Cg aldehydes (cyclic 1 and 2) were also formed in significant amounts ranging from 16.40 to 18.23Z of the C8 oxo-products. Compared to these, only lesser amounts of identified dibranched aldehydes, 2,~-dimethylheptanal (2,6-Di-Me) and 2-methyl-5-ethylpentanal (2,5-Me,Et) were found. Their combined total ranged from 2.89 to 4.25X.
Among the products derived from the broad C8 feeds, there were also significant quantities of 2-propyl-4-methyl-pentanal (2,4-Me,Pr) and cyclic C8 aldehydes. However, these lower boiling aldehydes not listed in the table since most of their derivation clearly depends on cyclic C7 olefins which are not present in significant quantities in the narrow feed.
The total percentage of some by-products having longer retention times than n-nonanal is also shown in Tabla II. These components are cyclic Cg aldehydes, Cg alcohols and Cg alkyl formates. Their total in the final reaction mixtures derived from the broad cut C8 feeds is 18.9 and 17.2X. The same by-products derived from the narrow C8 feeds amount to only 6.42 and 4.41X, respectively. The greater amounts of broad feed derived by-products are due to the formation of higher amounts of cyclic aldehydas.
To assess the further processability of the hydroformylation reaction mixtures, they were further analyzed by capillary GC using both a flame ionization detector (FID) for organic compounds and a sulfur specific detector (SSD) with quadratic response. As expected two different types of chromatograms were obtained for the reaction mixtures derived from the broad and narrow feeds. Uhether the feed was caustic treated or not did not make a perceptible difference as far as the composition of the reaction mixtura was concerned. The pentanethiol components of the untreated feeds were essen~ially all converted during hydroformylation. The thiophenic sulfur compounds remained essentially unchanged in all cases.
Only trace amounts of sulfur compounds (5-50 ppm range) boiling in the Cg aldehyde range were formed during hydroformylation. Sulfur specific GC showed that thelr formation was concurrent with tha early fact period of hydroformylation. No high boiling sulfur compounds could be ~rd43 '~3 detected in the reaction mixture. It is recalled that the amount of dimers and trimers is minimal under the mild hydroformylation conditions used.
In the following, the composition of the reaction mixtures is illustrated by Fi~ure 14, showing comparable chromatograms based on FID and SSD detection. Figure 14 shows chromatograms of a final, untreated mixture derived from the methanolic potassium hydroxide washed narrow feed.
The lower FID trace of Figure 14 shows, that in the case of the narrow cut feed there is a wide separation between the higher boiling C8 aromatic hydrocarbon feed components and the lower boillng dibranched Cg aldehyde products. It i5 also noted t~at the total aldehyde product selectivity is high. There are only small amounts of high boiling dimer and trimer by-prod~lcts in the reaction mixture.
The upper SSD trace of Figure 14 shows that most of the sulfur is in the region of the unconverted hydrocarbons. The thiophenic sulfur components of the feed remained unchanged. However, some other sulfur components were converted mostly to sulfur compounds of unknown structure.
Minor amounts o H2S are present in the reaction mixture. There are also two very minor higher boiling sulfur compounds present. They have GC
retention times slightly greater than the aldehyde products. The sulfur concentration of these two sulfur compounds is about 20 ppm while the sulfur concentration in the hydrocarbon region is about 2,000 ppm.
All four reaction mixtures were aerated in the presence of refluxing aqueous acetic acid to convert the cobalt compounds to water soluble cobalt acetate. In case of the reaceion mixture derived from the untreated broad cut feed some dark precipitate was formed on standing.
This was not dissolved completely during the cobalt removal procedure.
Thu~, the mixture was filtered to remove it. All the other reaction mixtures were decobalted without complication. The products derived from the narrow feed were easier to process. The methanolic KOH treated feed gave a mixture which was particularly easy to phase separate after caustic washing.
Capillary GC analysis of the cobalt free reaction mixtures indicated no significant change during decobalting. However, the de-cobalted mixtures do not appear to be storage stable. During two months standing at room temperature, the formation of high boiling sulfur compounds was observed by sulfur specific GC analysis of ~he decobalted reaction mixture derived from the untreated broad cut feed. Also, more dimer formation occurrsd during distillation if the decobalted aldehyde was aged. (It is also noted that the distilled aldehyde also tends to dimerize, i.e. aldolize, slowly during long term storage. In addition, the aged aldehyde forms ~ore heavies during hydrogenation.) The di~tillation of the decobalted aldehyde was carried out at a very low pressure to avoid any thermally induced dimerization. At first the uncon~erted C8 hydrocarbons were distllled at about 20-25C under 1.2 mm pressure. Most oi the hydrocarbon fractions from the broad cut feeds and all the hydrocarbon distillates from the narrow cut feed were color-less. Thereafter, the vacuum was decreased to 0.1 ~m and the aldehydes were distilled. The aldehydes derived from the broad cut were fractionated between about 20 and 35C, while those of the narrow cut feed were obtained between 23 and 31C. Several aldehyde fraceions were taken. ~ost of them were colorless but so~e of them were slightly yellow. The alcohols and formate esters were distilled between about 31 and 42C at 0.1 mm. They were also clear, colorless liquids. However, distillation of the formates at higher temperature (i.e. under increased pressure) resulted in de-composition and a slightly yellow distillate. There was no attempt made to obtain pure alcohol and formate fractions.
The results of the distillations of the various hydroformylation mixtures were summarized and compared. The aldehyde plus alcohol and formate distillates were combined in each case and designated as oxo-products. Based on che results the yields of oxo-products and resLdual heavy (dimeric and trimeric) by-products per 1000 g reaction mixture were calculated. The comparative data obtained are tabulated below:
Yields of Oxo-Distillate Products and Residual By-Products From Various Hydroformylation Mixtures, g/1000 g DerivedFrom Different C8 Flexicoker Naphtha Fractions Broad Broad Narrow Narrow Un- Caustic- Un- Caustic-treated treated treated treated , . __~ , ._ _ Aldehydes 397 494 525 534 Di~ers 85 79 72 53 The tabulation shows that the yield of distilled aldehydes per 1000 g reaction mixture varied widsly, from 394 to 534 g. As expected, great~r yields of distilled oxo-products were obtained from the narrow C8 cut feeds of higher olefin content than fro~ the broad C8 fractions. Fro~

the two broad cuts, the caustic treated led to a much higher yield than the untreated because of the higher feed conversion (See Table I).
The Cg aldehyde distillates (9OZ or more) obtained from the different C8 Flexicoker naphtha ~ractions were mostly combined to provide feeds for hydrogenation to produce the corresponding Cg alcohol products.
It is noted that in the case of the products derived from the narrow cut untreated distillates, the normal n-nonanal content was generally lower and the higher boiling components were more prevalent than in the narrow cut feed derived products. All the hydrogenations were carried to in the previously described manner in the presence of 5X water and lOX Co5/MoS based catalyst at 160C under 3000 psi pressure for 20 hours. The same catalyst was used repeatedly in the present tests.
Analysis of the hydrogenated raction mixtures by packed and capillary GC indicated that the Cg aldehydes were reduced to the corresponding Cg alcohols in a selective manner. Alcohol selectivities ranged from 85 to 97~.
According to packed GC, the hydrogenation of the aldehyde derived from the broad C8 fraction occurred with less than 2% dimer formation. In the case of the narrow cut derived aldehydes, dimer formation was about 15X.
Capillary GC indicated that paraffln for~ation was minimal, less than 5%, in all cases. n-Nonane was by far the largest paraffin by-product. Its concentration was leqs than 4X of the alcohol products. In general, the selectivity toward the isomeric nonyl alcohol products, was very high.
Sulfur specific GC shows that most of the low boiling sulfur impurities are dimethylthiophenes, i.e. components of the broad cut C8 feed. There was no sulfur in the alcohol retention time range.
However, all of the ~ixtures showed the presence of some sulfur in the dimer range. Apparently, the minute smounes of aldehyde range sulfur compounds, which were present in the feed, were convert~d during hydrogenation into sul~ur derivatives of low volatility.
The composition of the low boiling components o~ the aldeh~de ~eeds and alcohol product mixtures was studied by capillary GC and compared in Table XXyIlI~ The data of Table XXVIII show that both of tha feeds and the products derived ~rom the broad C8 cut contain less normal, i.e. less C3 straight chain, oxygenates and more higher boiling components than those of the narrow feeds. The concentration of n-nonyl alcohol -120- ~$~73 Table XXVIII
Hydrog~natlon of tha Gg Ald~h~des Deriv~d Via the Hydroformylation of Various C8 ~losicoker Naphtha Fractions Components Composition by Capillary GC X
in Relation (of Reactants and Products Derived to Their ~from Various C8 Fractions TypeRetention oftimes to Broad Broad Narrow Narrow MixtureNormal Un- Caustic Un- Caustic ~31Y3~ g~ reated Treated tre~ted Treated Aldehyde ShorterC 57 52 64 56 Reactant Normala 27 21 33 37 Longerd 16 27 3 7 AlcoholParaffinse 1 4 2 0 ProductShorterf 61 52 60 59 Normala 22 17 29 35 Longerg i6 27 9 6 a)Normal n-nonanal reactant or n-nonyl alcohol product. b)Based on analyses of hydrogenation feeds and reaction mixtures. C)Branched aldehydes of shorter retention time. d)Aldehdyes of longer retention time. e)c~ Paraffin by-products. f)Alcohols of shorter retention times. g)Alcohols of longer retention time, excluding dimers.

.

products is generally lower than that of their n-nonanal presursors. This is due to the preferred aldolization of n-nonanal to provide the corresponding dimer. The exact concentrations of dimers could not be determined by capillary due to their limited volatility. The concentration of volatile C8 paraffin by-products was generally very low, 0 tO 4X, as indlcated by the Table XXIII.
The hydrogenation reaction mixtures wer~ washed with lOZ
aqueous sodium hydroxide solution and then water to remove hydrogen -~ulfide and carboxylic acid by-products. The separation of the aqueous and organic phases occurred readily. After a final water wash, the mlxtures were fractionally distilled to recover the alcohol products.
A 24 plate Oldershaw, column was employed to separate the hydrocarbon solvent and by products from alcohol product fractions. The dlmer and trimer by-products wers usually obtained as a distillation residue. The fractional distillations were carried out under reduced pressure using a heating bath of less than 200C to avoid the decomposition of sul~ur containing dimeric by-products.
All the Cg alcohol distillates were colorless clear liquids. The alcohol fractions of the reaction mixture derived from the broad C8 feed fractions, were distilled between 91 and 111C at 18 mm. As expected, the alcohol distillates derived from the narrow C8 feed had a narrower boiling range. They were obtained be~.ween 95 and 107C at 18 mm.
The dimer by-product derived from the narrow C8 feed was also distilled.
It was obtained as a clear colorless liquid between 88 and 98C at 0.05 mm.
The alcohol distillate products derived from each of the four C8 Flexicoker feed~ were analyzed by capillary GC and combined to provide four alcohol products. The distillation residues consisting of undiqtilled alcohols, dimers and trimers were anlayzed using a packed colu~n GC. The yields of products and by-products, obtained from the crude products and by-products, obtained from the crude product mixture after distiling of the hydrocarbons, are shown togather with the yields of combined alcohol distillates in the following tabulation:

122~ 73 Yields of Alcohol Distillate Products and Residual Products, wt Z
(Derived from Different C8 Flexicoker Fractions) .
Broad Broad Narrow Narrow Un- Caustic Un- Caustic TreatedTreated Treated Treated .. .. _ . . _ Alcohol Distillate 90 58 77 6~
Residue 10 42 23 40 Alcohol 9 10 3 10 Dimer 1 31 18 30 Trimer - 1 2 As it is shown by the above data, the yields of alcohol distillates range fro~ 58 to 90Z. The large differences in Cg alcohol yields are apparently due to the different degrees dimer by-product formation. The dimers are dexived from the aldehyde reactants via aldolization hydrogenation. It is noted that there is less dimer formation from the aldehydes derived from ehe untreated C8 feeds, possibly due to the presence of aldolization inhibitors.
For a comparatlve characterization of the composition of the free alcohols based on eheir capillary GC, the conceneraeions of the components, having shorter retention eimes than n-nonyl alcohol, were added up. Similarly, the total percentage of the components having longer retention times was determined. These percentages were ehen compared wieh that of the n-nonyl alcohol componen~. They~are shown for all four alcohol products by the following tahulation:
: :

Composition of Cg Alcohol Distillate Products (Derived from Different C8 Flexicoker Fractions)~

Grouping of C~mponents Broad Narrow According to GCBroad Caustic Narrow Caustic Retention TimesUnTreatedTreated UnTreated Treated Shorter ~4.2 73.7 63.5 67.0 n-Nonanol 26.3 19.1 34.5 31.6 Longer 9.5 7.2 2.0 1.4 The data of the tabulation show that the broad C8 feed derived alcohols have a lower percentage of n-nonyl alcohol component than those based on narrow C8 feeds. The differences betwen the n-nonanol content of distilled alcohol products derived- form untreated and caustic treated C8 feeds appear to be due to difEerences in alcohol recovery. The distillation of higher boiling alcohol components was less complete from the product mixtures derived from the caustic treated feeds.
The above discussed four Cg alcohol distillates did not contain any sulfur detectable by the capillary GC method employed (This method would have detected any single sulfur compound present in 5 ppm concentration or more). Samples of these distillates were submitted for total sulfur analyses. The broad untreated and treated distillat~s were found to contain 22 and 43 ppm sulfur while the corresponding distillates derived from the narrow cut feed contained 13 and 31 ppm, respectively.
Overall, the above analytical results and other observations suggest that most of the sulfur compounds, distillin~ in the dimer by-product range, were formed during the aldehyde to alcohol hydrogenation.
A series of comparati~e odor tests were carried out with the alcohol products. The results showed that these alcohols have odors typical of Cg oxo alcohols in general.
The aicohols were converted to semilinear diheptyl phthalate which was evaluated as a plasticizer.

g~3 Examples 35 and 36 Hydroformylation of A C9 Olefinic Fraction of Naphtha by H2/C0 with Cobalt fit 3000 pSi in the 130-150C Tempera~ure Range The Cg olefinic feed for the present hydroformylation experiments was derived from a C4 to C12 Fluid-coker naphtha by a double 15/10 type distillation. The second distillation started with a broad Clo cut of bp. 145 ~o 155C and produced a narrow cut of bp. 143 to 148 in about 40X yield. The concentrations of the major components of the broad and narrow boiling fractions is shown by the following tabulation:
Cg Olefinic Feeds Boiling Point ~
F Narrow Cut Broad Cut Ide~iflcation C _ A~prox.Bp! (143-148C)Bp. (145-155C) Ethylbenzene 136.2 277 0.35 2.10 2,5-DiMe-Thiophene 136.7 278 p-Xylene 138.3 281 1.82 7.04 m-Xylene 139.1 282 0.64 2.05 o-Xylene 144.4 292 4.93 7.90 l-Nonene 146 295 24.09 21.06 n Nonane 150.8 303 17.43 15.45 2,3,5-TriMe-Thiophene 160.1 320 l-Me, 3-Et-Benzene 161.3 322 0.75 2.17 It is apparent that the second 15/10 distillation was not sufficiently effective to produce the desired narrow l-n-nonene rich fraction in a high yield. However, it was possible to exclude most of the aromatic components from the narrow nonenes fraction by accepting a low distillate yield. Sulfur GC indicated that most of dimethylthiophenes and trimethylthiophenes were removed during the second fractionation with the xylene~ and l-methyl-3-ethylbenzene respectively. The sulfur content was reduced from about 1.5 to 0.2%.
The narrow, olefinic fraction of the Cg Fluid-Coker naphtha (E-7285~ was hydroformylated using a 1/1 mixture f H2 and C0 in the presenca of 0.2Z Co. The cobalt catalyst was introduced as a 13X solution of its precur~or, Co2~CO)g, in toluene. The reaction was carried out at 3000 psi at variable temperatures. The tem~erature was increased during the course of the reaction to convert the various types of olefins at their minimum reaction temperature.
The solution of the Co2(CO)g catalyst precursor was added at 120C. This reaction temperature was maintained for 1 hour.
Thereafter, the reaction temperature was raised to 130C. Similarly, the temperature was increased to 140 and then 150C after 1 and 2 hours, respectively. After a total reaction time of 4 hours, the reaction was discontinued.
The results of the hydroformylation are shown by Table XXI~. For comparison, Table XXIX also lists some of the data obtained in an experiment carried out with 0.1~ catalyst in an identical manner.
The composition of the reaction mlxtures by packed GC
showed that the reactlon was highly selective to aldehydes. After 4 hours in the presence of 0.2X Co, the hydroformylation reaction~was essentially complete. According to GC, the concentration of aldehydes in the reaction mixture reached 45X. Experiments to determine GC response factors with n-decane, n-decanal, n decanol mixtures indicated that the 45X GC response corresponds to about 50% by weight of aldehyde. To reach this aldehyde concentration, a minimum of 44.7X olefins in the feed had to be hydro-formylated. At this point, the reaction mixture still contained only about 5X alcohols plus formate esters and about 3X dimers and trimers.
The isomeric Clo aldehyde distribution by capillary GC
showed that the n-decanal was far the most prevalent oxo product.
n-Decanal, of course, was mostly derived from the most reactive l-n-nonene olefin component of the feed. Thus, its percentage of the total oxo-products wa particularly high (58.2X) during the early phase of the reaction. At the completion of the reaction, the percentage of the n-decanal was 44.8X. The percentage of the second largest Clo aldehyde isomer, 2-methylnonanal was 16.2X. Thus, these two aldehyde isomers which can-be derived from l-n-nonene made up 61X of the oxo products. As expected the other 2-alkyl substituted Clo aldehyd~s derived from linear internal octenes (2-ethyloctanal, 3-propylheptanal and 4-butyloctanal~ were other significant product isomers, in a total concentration of 71.3X.
Thus, the total concentration of oxo products derived from linear olefins is about 84.8X.
Capillary GC also indicated the formation of comparable amounts of isomeric alcohols and their fcrmate esters having alkyl structures corresponding to the isomeric decanal products. These secondary 3 ~ o 3 ¦ ~ ~ ~
~ ~y ~o ¢ U ~ o o o X U
o ~ '~
~o ~ ,~ ~
.~¢ ~oo ,~ "

~- ~ t~ ~ 0~ A 'g 8 ~ ~ " o ,. o u~
~ ~j o 3 3 C ~æ: rC a~ D O
~ ~ ~ ~ z ~ = 3 O N '~ C ¦ ~ ~ ~1 ~

-¢~ o I O~

O ~ O o 0 ¦ ~o ~0 o~ ~ ~ ~ ,~ r ~
~ ~ cC ~ ~ ~
o 3 ~~ ~c ~ ~ o o ~ E~ '~
~- ~ ~ l u ~ 7 @ U~ C~ OOOO OO
~,o ~~ ox C~

C E~ ~ N

by-products were rich in the normal isomers, particularly the alkyl formates.
The decobalted combined hydroformylation reaction mixtures w~re fractionally distilled in vacuo uslng a two foot packed column. At first 1160.5 g ~34.7X) of unreacted hydrocarbon components were removed at close to room temperature under 1 mm pressure. Thereafter, the remaining 2184.5 g (65.3X) mixture of oxygenated products was fractionated. Most of the isomeric Clo aldehyde products were received between 43 and 49C at 0.5 mm. me total amount of the aldehyde distlllates was 1197.5 g (35.8X of ~he reaction mixture). No attempt was made to separate the Clo alcohol and Clo alkyl formate products. They were d1stilled between 40 and 55C at 0.05 ~m and received as 466.5 g (14X) of a colorless to light yellow liquid. Thus, the combined weight percentage of alcohols and alcohol precursors in the reaction mixture was 49.8. The C20 dimer products of aldehyde condensation were largely distilled be~ween 118 and 122C at 0.05 mm. About 296 g (11.8X) of these dimers were obtained as a pale yellow liquid, Finally, 102.5 g (3.2X) C30 trimers were also obtained largely at abut 215C at 0.5 mm. as a clear yellow distillate. The last of the trimers wers distilled with some decomposition. The distillation residue was 22 g (0.5%) of the reaction mixture.
To prepare the desired semilinear Clo alcohol, a combined Clo aldehyde feed of the following composition was used.
2-Butylhexanal 1.0 n-Decanal 23.4 2-Propylhexanal 3.3 n-Decanol 2.1 2-Ethyloctanal 3.9 n-Decyl Formate 4.2 2-Methylnonanal 10.1 29.7 18.3 The percentage of the n-decanal in this feed (23.4X) is low due to its preferential condensation and further side reactions of n-decanal during distillation. Sulfur GC of this aldehyde showed no sulfur compounds in the aldehyde range. Trimethylthiophenes were present in about 40 ppm concentration indicatLng an imperfect separation of feed hydrocarbons from product aldehydes by distillation.
The above Clo aldehyds feed was hydrogenated in ~he presence of 5Z water 10.7 wt.X CoS/MoS catalyst at 150C (302F) under 3000 psi (307 atm) for 40 hours. The desired al~ehyde to alcohol conversion was complete. Combined GC/MS analyses indicated the absence of aldehydes.
The percentage of n-decyl alcohol in the crude product 29.6Z. This percentage corresponds to the combined concentration of n decanal, n-decyl formate plus n-decyl alcohol in the aldehyde feed.
Sulfur GC of the crude alcohol indicated that the trimethylthiophenes were the main components (61X). However, minor amounts of sulfur (39%, about 14 pp~) was also present in the alcohol range. These latter sulfur compounds were apparently formed from the low molecular weight sulfur compounds during hydrogenation.
Most of the crude Clo alcohol hydrogenation product (1356 g) was fractionally distilled using a 24 plate Oldershaw column. An early product fraction (47 g) containing aromatic hydrocarbons and trimethylthiophenes was obtained between 21 and 110C at l9 mm.
After the above fraction, six colorless, clear isomeric decyl alcohol distillate fractions were obtained. Their amounts, boiling ranges, linearity and total sulfur content are listed in the following:
Bp. _ _cohols,_Z _ Fraction AmountC/mm Decyl- Highera Approx.
No. g n- i- S, ppm V 222 111-116/19 6.7 93.3 50 VI 222 116-117/19 14.1 84.7 2.2 20 VII 423 117-121/19 41.9 49.0 9.1 40 VIII 196 121/19 58.3 41.717.3 30 XI 3~ 58/0.1 2.0 83.083.0 The higher alcohols having retention times lon~er than n-decyl alcohol are probably dibranched undecanols.
The above tabulation shows that ~ractions enriched in the linear alcohol can be obtained. Fractions VI, VII and VIII were selected as heart cuts ~or conversion to semilinear didecyl phthalate esters, which are evaluated ~s plastlcizers.
Altogether 1230 g (86Z) of clear, colorless alcohol products were recovered by distillation between 111 and 121C at l9m~.
About 45 g (3.4 wt. X of the feed) of a clear yallow distillate was recovered ~n the broad dimer range. Apparently, some aldehyde condensation occurred during hydrogenation. Most of the dimer distilled between 165 to 172C at 0.1 mm. Packed GC of the distillation residue (18 g) indicated that its volatile components were in the trimer range. The aldehyde fraction had no sulfur detectable by GC. However, the dimeric aldol alcohol contained about 1.2X sulfur.

Examples 37 and 38 Hydro~ormylation o~ Clo Naphtha B7 3/2 H2/CO
wi~h 0.2 and lX C~balt at 130C ~nd 3~00 psi The preYiously described Clo fraction of the Fluid coker naphtha was hydroformylated as a 1/1 mixture with hexane at 130C by an about 60/40 mixture of H2/CO at 3000 psi, using the high pressure procedure. The catalyst precursor was dicobalt octacarbonyl.
In the first experiment, the cobalt complex catalyst used was equivalent to 0.2X cobalt, i.e., 34mM. The reaction mixture was period-ically sampled and analyzed by capillary GG. The progress of the reaction was followed by determining both the l-decene reactant consumed and the aldehyde product. The main aldehyde produts were the n-aldehyde and 2-methyl substituted aldehyde derived from l-decene. The data obtained are tabulated in the following:
_ Reaction Time. Min _10 30 60 120 l-Octene Converted, X 12 54 100 100 Ma~or Aldehydes Formed, Z 7 51 93 105 Total Aldehydes Formed, X 143 203 n/i Ratio of Ma~or Aldehydes 3.35 3.39 3.15 It is apparent from the data that the l-n-decene was converted at first. However, by the end of the 2 hour reaction period a significant reactlon o~ the isomeric decenes also occurred. The final ratio of the two ma~or aldehydes formed was 3.15. No significant secondary reaction took place. Alcohol formation was negligible. High boiling by-products were virtually absent.
In ~he second experiment, the same reaction was carried out in the presence of 1~ cobalt. This resulted in a very fast reaction. In 10 minutes, the l-decene component was completely converted. The amount of the two ma~or aldehydes formed was 105~ of the theoretical quantity -130~ L~ ,3 derivable from l-decene. The total aldehydes formed were 212X of this calculated value. The n/i ratio of the two major aldehyde products was 2.71.
The second experiment was also run for 2 hours. During the second hour much hydrogenation occurred. By the end of the second hour, essentially all the primary aldehyde products were converted to the corresponding alcohols.

Examples 39 and 40 ~ydro~or~yla~ion of C8 Naphtha by 3/2 and 1/1 H2/CO with Cobalt at 130C and 3000 psi The C8 fraction of the previously described naphtha was hydro-formylated in hexane in the presence of 0.2Z cobalt at 130C and 3000 psi in two experiments. The H2/CO reactant ratio was about 60/40 in the first experiment while an equimolar mixture of synthesis gas was used in the second.
Qualitatively, the reaction of octenes in this example was similar to that of decenes as described in the previous examples. However, the reaction rates were generally lower. A summary of data obtained in the first experiment with 60/40 H2/C0 is provided by the following tabulation:

_ Reation~Time. Minutes 30 _ 60 120 l-Octsne Converted, X 6 14 29 100 ~ajor Aldehyde Formed, X 3 10 21 92 The reaction had an induction period during the first hour. However, the conversion of l-n-octene and some of the isomeric octenes was rapid during the second hour. The total amount of aldehydes formed was 144X of the theoretical amount produced from l-n-octene. Nevertheless, due to the low reaction temperature, no aldehyde hydrogenation to alcohol occurred. The n/i ratio of ehe two major products was 2.78, definitely lower than in the analogous experiment of the previous example.
The second experiment of this example was carried out under the same process conditions, but using a l/l rather than 3/2 mixture of H2 and CO reactant. The results of the two experiments were very similar; the H2/CO reactant ratio had no apparent major effect at this temperature. The second experiment using 1/1 H2/CO appeared to have a slighely longer $73 induction period. Howsver, during the second hour of the reaction, a rapid conversion took place. By the end of the second hour, all the l-n-octene was converted. The reaction was continued for a third hour. Additional conversion of the other isomers occurred. After three hours reaction time, the total amount of aldehydes formed was 187X of the theoretical yield calculated for the l-n-octene component of the feed. On the same basis, the yield of the total aldehydes formed in 2 hours was 125X.

Examples 41-42 Hydroformylation of Cg Naphtha by H2/CO ~i~tureg of Va~ying Ratio~ with Dlcobalt Oceacarbonyl at 150C and 3000 p~i C8 naphtha fraction was hydroformylated in hexane solution as usual in the presence of 0.2X ~obalt provided as dicobalt octacarbonyl.
Compared to the previous example, the only significant difference was the use of a higher temperaturs, 150C. Three experiments were run with different initial and/or final H2/CO ratios.
In the first experiment, where a 3/2 ratio of H2/CO was used all through the reaction, a severe inhibition of hydroformylation was observed.
After 1 and 2 hours reaction time, the amounts of reacted l-n-octene were only 20 and 27X, respectively. As expected, the significant products were n-nonanal and 2-methyl octanal. Their ratio was 3.48.
In the second experiment with an initially equimolar H2/CO
reactant, a much faster reaction was observed. About 20X of the l-n-octene component reacted in 10 minutes according to GC; all ~he l-octene reacted in 30 minutes. In 60 minutes, much of the linear octenes and 2-methyl heptene-l wera also converted. The product data obtained on GC analyses of product ~amples were the following.

-~@i9s12D_~L~e__~in _ 30 _ 60 120 Two Ma~or Aldehydes Formed, X 59 92 84 Total Aldehydes Formed, X 82 182 201 n/i Raeio of Major Aldehydes 2.59 2.41 1.92 The data indicate that significant amounts of olefin isomerization occurred during hydroPormylation. Dur~ng the first part of the reaction, the ma~or l-n-octene component was partly isomerized to the thermodynamically favored linear octene Thus, no l-octene was shown in ~ ~ 6 3 the reaction mixture after 30 minutes, even though only 59~ of the products derivable from l-n-oc~ene were formed. Most of the hydrofor~ylation took place during the subseq~-ent 30 minutes. An apparent side reaction during the second hour was the hydrogenation of aldehyde products to the corresponding alcohols. By the end of the reaction, llZ of the total n-octanal formed was converted to n-octanol. However, the hydroformylation of internal octenes during the same period more than made up for the loss of total aldehydes via hydro~enation. During the second half of the hydrogenation period, the yield of the total aldehydes for~ed increased from 182X to 201X of the calculated yield for l-n-octene. At the end of the reaction, less than half of ehe aldehydes were derived from l-n-octene.
As the amount of aldehydes formed from isomeric octenes rather than l-n-octene increased with time, the n/i raeio of the two main aldehyde products dropped from 2.59 to 1.92. The apparent increase of 2-methyloctanal formed in part was due to the overlap of GC peaks.
However, additional amounts were formed from 2-octene.
It is noted that although the initial H2/CO mixture used to pressure the reaction vessel was equimolar, the feed gas durin~, the reaction had a H2/CO ratio of about 60/40. Since the liquid reaction mixture was sampled four times with considerable gas loss, by the end of the reaction the H2/CO ratio Increased to 60/40. It is felt that the initially low value of H2/CO was critical in overcoming reaction inhlbition.
In the third experiment, the H2/CO reactant ratio of both the initial and run synth~sis gas was equimolar. However, the maintenance of the low H2/CO ratio r~sulted in decreased reaction rates when compared to the prevlou~ experiment.
The a~ounts of l-octene converted after lO, 30, 60 and 120 ~inutes, were 30, 38, 79 and 100%, respectively. The yields of the t~o major products, n-octanal plus 2-methyl heptanal, after 60 and 120 minutes were 44 and 86X, respectively, based on l-n-octenf~. During the same last two periods, the yield of the total aldehydes for~ed was ~l and 170%. The n/i ratio of the two ma~or products was 2.70 and 2.48, respectively. By the end of the reaction, 3.5X of the n-oceanal was hydrogenated to n-octanol. Overall, ths GC data obtained showed that although l-octene conversion started immediately, the final extent of hydroformylaeion was lower than in ehe previous example. High CO partial pressure was important $ L~: ~ 7 3 in overcomin~ the initial inhibition, but the H2 partial pressure was insufficient to assure a high hydroformylation rate.

Example 43 Hydroformylation of C8 Naphtha by 3/2 H2/CO wlth Dicobalt Octaearbonyl at 150C and 4500 p8i A hexan0 solution of C8 naphtha was hydroformylated as usual in the presence of 0.2X cobalt by 3/2 H2/CO at 150C and 4500 psi. The conditions were identical to those of the first experiment of the previous example, except the pressure was increased In the present experiment from 3000 to 4500 psi. This resulted in a drastically reduced initiation period and a much more complete conversion of the olefin components during the two hour reaction period.
In ten minutes, 19% of the l-n-octene was converted and n-octanal was formed in amounts corresponding to llX of the starting l-n-octene reactant. Thereafter, a rapid reaction took place. In 30 minutes, essentially all the l-n-octene and the 2-methyl heptene-l were converted.
GC analyses provided the following data on the products formed.

Reaction ti~ Min.

Two Major Aldehydes Formed, X 95 120 105 Total Aldehydes Formed, X 149 247 291 n/i Ratio of Ma~or Aldehydes 2.9 2.7 2.5 n-Octanal Converted to n-Octanol, Z 10 16 It is particularly noted, that after the initial conversion of the l-n-octene in 30 minutes, the total yield of aldehydes increased from 149 to 291% of the calculated yield for l-n-octene. This increase is due to the conversion of internal olefins. It should also be noted that the final n~i ratio of the two majo~ aldehyde products was fairly high (2.5), considering the high conversion of intPrnal olefins.
During the second hour of the reaction period, there was a ~light decrease of ~he two ma~or aldehydes in the mixture. This is apparently due to hydrogenation of aldehydes to alcohols. A comparison of the GC signal intensities indicated that about 16X of the n-octanal formed was converted to n-octanol.

-134- 1~7~

Thus the results show that at the increased pressure of the 3/2 H2/CO mixture, the concentration of C0 is sufficient to overcome the sulfur inhibition. The high partial pressure of hydrogen results in a high reaction rate of both l-n-olefins and internal olefins.

Example 44 Hydroformylation f Clo Naphtha By H2/CO Mixtures of Vsrying Ratios ant With Varying Concentrations of Cobalt and Sep~ration of Cll Aldehyde Products The Clo fraction used was a high boiling naphtha fraction. The l-decene content of this fraction by GC was about 16X. Based on an NMR
analysis, the type distribution of the decene components was the following:

RCH-CH~ RCH~C~ R2C CH2 R2C-CHR Decadiene I II III IV Conjugaeed 43% 22X 14X 9X 12%
Ass~ing that l-decene is the only type I olefin present, the total percentage of olefinic unsaturates was 37X.
About 1900g portions of a Clo Fluid-coker naphtha fraction similar to the one previously described were hydroformylated without any significant amount of added solvent in a 1 gallon reactor. The cobalt catalyst was added as an approxi~ately lOX solution of dicobalt octacarbonyl in toluene. The resulting essentially non-diluted feeds contained increased concentra~ions of both olefin reactants and sulfur inhibitors. As such, they required greater amounts of cobalt for effective ca~talysis.
There were two experimenes carried out using dicobalt octacarbonyl as a catalyst precursor at 130C under 3000 psi pressure, with 1/1 H2/C0 and 3/2 H2/C0 reactant gas, respectlvely. The initial amount of the catalyst employed was equivalent to 0.2% cobalt in both cases. This amount of catalyst did not lead to any significant hydroformylation in S
hours in either case. Thereafter~ an additional O.lX and 0.2X, respectively, of cobalt were added after cooling the mixtur~ and starting the reactions again.

-135- ~2~73 When the first experiment was resumed in the presence of a total of 0.3X cobalt, hydroformylation occurred at a moderate ra~e. All the l-n-decene was consumed in 120 minutes. The total reaction time was 5 hours. GC analysis of the final reaction mixture indicated a total aldehyde product yield of about 253~, based on the amount of l-n-decene in the feed. The n/i ratio of the two ma~or aldehydes was about 2.7. Ths percentage of these aldehydes in the total aldehyde mixture was 41Z. In the case of the second experiment (Example 30~, with a total of 0.4%
cobalt, the hydrofor~ylation was fast. All the l-decene was converted within 10 minutes. This reaction was continued with the increased amounts of cobalt for 3 hours.
Overall, the two experiments gave similar results, and indicated that the initial small amoun~s of cobalt catalyst were deactivated, but the inhibitors were thus consumed. Thus, the added amounts of cobalt showed high activity which was little dependent on the H2/CO ratios employed.
The composition of the combined final reaction mixtures is shown by capillary GC and packed column GC's in Figures 8 and 9, respec~ively.
Fi~ure 15 shows a typical reaction mixture containing major amounts of n-paraffin and n-aldehyde. Clearly, recognizable isomeric aldehyde products are also shown. These 2-alkyl substituted aldehydes are apparently derived from thc various linear olefin isomers of the feed.
Their structure was established in GC/MS studies on the basis of the characteristic ions ~ormed on electron impact ionization. As it is indicated by the spectrum, decreasing amounts methyl, ethyl, propyl and butyl branched aldehydes are present.
Fi~ur~ 16 shows the packed column GC of the same reaction mixture. This GC shows less separation of the individual components, but extends the analysis to the high boiling aldehyde di~er and trimer by-products. It indicated that they amount only to about 2.9X of the total reaction mixture.
For a more detailed study of the products, it was decided to distill the reaction mixtures. The t~o products were combined. The cobalt was removed as cobalt acetate by hot aqueous acetic acid plus air treat ment. The organic phase (976g) was then fractionally distillsd in high vacuo using a one foot packed column. The unreacted Clo hydrocarbons were distilled at room temperature at 0.1mm and were collected in a cold erap, (491g, 50 wtX~. Therea~ter, the Clo aldehydes were dlstilled. Durin~ the distillation, some thermal decomposition of the residual liquid (probably -136~ $ ~3 of the formate by-products) took place. As a consequence, the vac~-um dropped to 0.5mm. How~ver, while the bath temperature was slowly increased to 100C, the decompositlon has subsided, the vacuu~ improved and the Cll aldehyde products were distilled between about 50 and 60C at 0.lmm and received as colorless liquids (371g, 38 wtX). The residual liquid dimers and trimers were 112g, 12 wtX. Packed CC indicated that about 2/3 of this residue was consisted of very high boiling compounds, probably trimers. A
large percentage of ~hese heavy by-products was formed upon heating the mixture during fractional distillation.
The distlllation results indicate that the total oxygenated product corresponds to ths yield calculated for a feed at 45% olefin content assuming complete conversion. The isolated aldehyde content is less, it corresponds to an effective utilization of about 36% of the total feed.
Capillary GC of the distillate product showed that the two major aldehyde products are derived via the hydroformylation of l-n-decene:
C8H17CH-CH2 C/H2~ C8H17CH2CH2CHO + CgH17CHCHO

These two ma~or products, n-undecanal and 2-~ethyldecanal, constltute 49X
of the aldehydes. Their ratio is 2.23. Other minor aldehydes were also identified by GC/MS.
Based on the above detailed analyses, it was calculated that the total oxygenated products contain 0.65 branch per molecule.
The Cll aldehyde products were reduced to the corresponding C
alcohols which were converted to semilinear diundecyl phthalates. The latter were evaluated as plasticizers.

Examples 45-47 Hydro~or~ylation of Atmo~ph8rically and Vacuum Distilled Clo Naphtha Fractlons ~ith Cobalt A series of three hydroformylation experiments was carried oue w~th three different Clo naphtha fractions in a manner described in the prevlous two examples to determine the effect of the conditions of the fractional distillation of the naphtha feed on the reaceivity. Infor~ation about the feeds and hydroformyla~ion results i5 sum~arized in Tabla_XXX.
The first fraction employed as a ieed was an atmospherically -137- ~ $~3 C I ~ ~ ~
~, ~ ~ _ _ ~Ql G' O et ~ --O
C ~ C I ~ o ~ ~
- ~ .l ._ 33 ; ~1 ~' ~-- ~ . ~a O ~, L
~æ ~ ~ c a ~i3c E ¦ O O O O O O ~77 0 ~ v~
o c~
C ~ ~ l OO~ 3 .~ I -~a ~ O a : ~ a~ ~1~ co _ ~ ~ o cr o~ _ o o~ ~ O ~
~_ ~ e t,~ 1~ _ O O ~ D O ~ N r~ 0 r~ ~ ~ _ X
_~ ~ ~ ~ D W C~ O ~ ~ e ~O ,0 ~ .
O E co O O o o o O o O o O o o o LL ~ ~ ~_ ~ a~ V~ 1-+ + -+~ ~

~a ~ a ~ o ~ ~ ~
O ~ ~, ¦ o c e o ~ ~ ~ ~ E ~

o o ~ ~ ~ c .'~ ~ E a;l G
a ~ -- ~ ~ E s ~) aa g c z E v~ o~n ~t 0~ E ~ ~ _ o ~ .C ~ ~ ~ c ~ --E 2~ ~

L.J ~ 1' "

7~3 distilled Clo cut between 342 and 350F (172-177C). According to capillary ~C, it contained 10.9X l-n-decene and 13.9X n-decane. About 55.5X of the components of this cut had longer retention times than n-decane. These components included indene.
The second fraction was obtained at reduced pressure under 240mm.
It contained 17.0Z l-n-d~cene and 15.0Z n-decane plus 42.7~ of higher boiling components.
The third fraction was derived from an atmospherically distilled Clo fraction by redlstilling it in vacuo at 50~m. This vacuum distilled fraction mainly consisted of compounds boiling in the range of l-n-decene, n-decane or lower. The n-decene and n-decane contents were 19.5 and 16.5X
respectively. Only 23.lX of this fraction had GC retention times greater than that of n-decane.
The above described, somewhat different, three Clo fractions were used as hydroformyla~ion feeds in the presence of 0.lX and then an additional 0.1X Co catalyst, both added as Co2(C0)g. Each run was carried out using 1/1 H2/CO as reactant gas under 3000 psi at 130C (266'F). The reaction mixtures were sampled at intervals and analyzed by packed and capillary GC columns. The results are summarized in Table XXX.
The GC composition data of the threa Clo reaction mixtures hydroformylated in the presence of 0.1~ cobalt in athe a series of experiments in Table XXX) show that no significant hydroformylation occurred in 360 minutes. There was some initial reaction as indicated by a small pressure drop and minor aldehyde formation during the first cen minutes. However, the reaction soon virtually stopped. It is apparent that the cobalt carbonyl was deactivated by the inhibitors present in the Clo coker dlstillate feed.
After the unsuccessful attempts of reacting the three Clo fraetions in the presence of 0.lX cobalt, an additional 0.1~ cobalt was added to the reaction mixtures. This resulted in effective hydroformylation in all three cases (in the b series of experiments).
However, the hydroformylation rates were somewhat dependent on the particular Clo feed as described in the following.
The atmospherically distilled Clo naphtha was the least reactive.
Even after the addltion of the incremental cobalt the reaction was slow to start and sluggish as it indicated by the minor amounts of products formed in an hour. The vacuum distilled naphtha fraction was significantly more reactive. When the additional amount of cobalt was added, major amounts of -139- ~ 73 aldehyde products (29Z) were formed within an hour. The reaction was essentially complete in 3 hours. The atmospheric Clo naphtha cut which was redistilled in vacuo was somewhat more reactive. However, the vacuum distilled naphtha was more active than the atmospheric naphtha redistilled in vacuo. This seems to indicate that the inhibitors formed during atmospheric distillation are not removed on redistillation in vacuo.
The data oi the table also show that there was very little dimer by-product formation in all cases. The amount of dimers formed during these reactions was less than 3X of the main aldehyde products. Although the amounts of trimers fo~med were not determined in this series of experiments, it is noted that as a rule considerably less trimer is formed than dlmer.
Analyses by capillary GC show that, as expected, the two main products of these hydroformylations were n-undecanal and 2-meehyldecanal, derived from l-n-decene. As it is shown by Table XXX, the n/i ratio of these two main products in the final reaction mixture was in the 2.9 to 3.7 range. There were, of course, other minor branched aldehydes present.
These were derived fron internal and branched olefins. The amount of the completely linear aldehyde, n-decanal, in the final reaction mixtures ranges from 31.1 to 38.3X. This variation clearly reflects the different percentages of l-n-decene present in these feeds. Similarly, as a consequence of the varying feed composition, the combined amounts of n-undecanal and 2-methyldecanal (n+i) changed from 41.7X to 51.lX. The rest of the product largely consisted of other monobranched 2-alkyl substituted Cll aldehydes such as 2-ethylnonanal, 2-propyloctanal and 2-butylheptanal. These monobranched aldehydes were apparently derived from isomeric linear internal decenes.
In general, comparisons of samples, taken from the reaction mixtures at different intervals, indicate that the l-n-decene component reacted at first, as axpected. Consequently, the products of partially reacted feeds were m~inly consisting of n-undecanal and 2-methyldecanal. As the reaction proceeded, and the internal and branched olefinic components were also converted, various branched aldehydes were formed and the relative amounts of the two major products derived from l n-decene decreased.
Only minimal amounts of the aldehyde hydroformylation products were reduced by hydrogen to the corresponding ~lcohols. The only 9~ 3 identifiable alcohol by-product was n-undecanol. Its amount was below 1 of the Cll aldehyde products.
The three final reaction mixtures obtained were brown, as usual.
Some of the brown color of the mixture derived from the atmospherically distilled feed persis~ed after the removal of the cobalt by the usual aqueous acetic acid, air treatment. However, the brown color of the mixture derived from the vacuum distilled feeds changed to dark yellow upon cobalt removal.
The cobalt free reaction mixtures were fractionally distilled, using a 2 ft. packed colu~n in vacuo, at pressures in the range of 0.1 - 50 O.2mm. The unconverted feed components were distilled as colorless liquids with a yellow tint at ambient temperatures (20 to 30C) using a dry ice cooled recelver. The aldehyde products were obtained as lighe yellow liquids between 47 and 57C at O.lmm pressure.
Due to the relatively low distillation and heating bath temperatures (100-135C bath), relatively little aldehyde dimerization and trimerization occurred during distillation. For example, in the experiment using vacuum redistilled feed, 1700g of the crude reaction mixture was distilled to obtsin 570g product and 51g distillation residue. GC analysis indicated that this residue contained 31Z product, 43X dimer and 26X
trimer. Thus, the combined dimer and trimer product was 35.2 g i.e., about 6X of the main product.
The aldehyde distillate products of the three runs were combined.
The combined product contained 37.lX n-undecanal, 10.4X 2-methyldecanal, about 8.6X of other 2-alkyl substituted monobranched aldehydes, about 28.7X
of aldehy~es having retention times longer than that of n-decanal. These latter compounds include doubly branched and possibly C12 aldehydes. The amount of n-undecanol is minimal, about O.2Z.

Hydrofor~ylation cf C~-C 5 Fluid Coker Light Gas Oil Fractlons ~th Co~alt (E~amples 4~-64 ,~
The previously described Cg to Cls light coker gas oil and its distillate fractions were hydroformylated without prior treating in the presence of cobalt at high pressure.
The hydroformylation of the non-fractionated Cg to C16 light gas oil was studied with cobalt in the presence and in the absence of added phosphine ligand. Thereafter, the hydroformylation of narrow single carbon distillate fractions from Cll to Cls was investiga~ed in the presenca of cobalt st 3000 psi. In general, it was found that the gas oil fractions were more reactive than the naphtha fractions, particularly when distilled in vacuo. The reaction rates were directly related to the temperature, in the 110 to 170C range. The n/i ratio of the aldehyde products was inversely related to the reaction temperature. The isomeric aldehyde products were isolated from the reaction mixtures by fractional distillAtion in vacuo. The two major types of products were n-aldehydes and the corresponding 2-methyl aldehydes. The aldehydes products were reduced to the corresponding alcohols, in the presence of a sulfur resistant Co/Mo catalyst.

Example 48 Hydroformylation of Cg-Cls ~hole Coker Light Gas Oil ~ith Cobalt st 150C and 4500 psi The previously described Cg-Cls light gas oil was hydroformylated withou~ solvent by a 1:1 mixture of H2/CO. A toIuene solution of Co2(CO)g was introduced at 120C temperature and 3000 psi pressure into the reaction mixture to provide a cobalt concentration of 0.4X. When no reaction occurred, the conditions were changed to 150C and 4500 psi. After a 30 minute induction period, a rapid hydroformylation reaction occurred. This agrees with the hypothe~is that there are equilibrl~ among the various sulfur substituted cobalt carbonyl complexes. Dapendent upon the types and amounts of sulfur compounds present in the feed, sufficiently high concentrations of CO are required to avoid the formation of inactive carbonyl-frea complexes.
After a total reaction period of 3 hours, the reaction was discontinued. Th~ capillary GC of the resulting mixture is shown by Fi~ure 17, It is apparent from the figure that the prominent l-n-olefin peaks of the gas oil feed are absent after hydroformylation. The l-n-olefins were converted mainly ~o n-aldehydes which show up as prominent peaks in the high retention region of the GC. ~he relative intensities of the Cll eO
C16 aldehyde peaks are about the same as those of the parent Clo to Cls olefins. The l-n-olefins of the feed appear to be of similar reactivity without regard eo their carbon number. This is in contrast to the behavior of branched higher olefins whose reactivity i5 rapidly decreasing with increasing carbon number.

-142~ 73 Examples 49-51 Hydroformylation of At~osph~rically a~d V~c~um Distilled Cll Naphtha ~nd Gas Oil Fractions with Cobalt A series of three hydroformylation experiments was carried out with a Cll fraction of naphtha and the combined Cll light gas oil fractions of a Fluid-coker distLllate in the manner described in Examples 41 to 45.
The experim2nts were designed to determine the effect of the conditions of the distillation of the gas oil feed on reactivity. Information about the feeds used and the hydroformylation results obtained is summarized in Table ~1. Some of the details are described in the following.
A narrow cut Cll naphtha fraction boiling between 63 to 71C (146 to 150F) under 238mm pressure was used in Example 49. In Example 50, the previously described combined Cll fractions of light Fluid-coker gas oil, were employed. These fractions were obtained between 185 to 196C (365 to 385F) at atmospheric pressure. Part of tha same Cll fraction of light coker gas oil was redistilled without fractionation at 50mm pressure. This redistillation of the orange Cll fractions gave a yellow distillate, used as a hydroformylation feed in Example 51.
Each of the above Cll feeds was hydroformylated Ln the presence of O.lZ Co, added as Co2(CO)g. Each run was carried out using 1/1 H2/CO
under 3000 psi at 130C (266F). The reaction mixtures were sampled at intervals and analyzed by packed column and capillary GC.
The GC composition data of the reaction mixtures of Table XIX
show that all the Cll fractions could be hydroformylated under tha above conditions but at different rates. The vacuum distilled naphtha fraction was more reactive than the atmospherically distilled gas oil fraction (Examples 49 and 50, Seq. Nos. 1 and 2). The gas oil redistilled in vacuo was the most reactive Cll fraction of all (Example 51).
It is clear from the comparative reactivities observed that distillation in vacuo rather than at aemospheric pressure resulted in increased rsactivity. While the present invention is independent of the explanation of thes~ findings. We hypothesize that atmospheric distilla-tion at high temperatur~ results in the thermal decomposition of some of the ~hiol components to H2S plus olefin. Some of tba H2S formed may dissolve in the atmospheric distlllate and inhibit the hydroformylation proc0ss.

_ O ~ ...... . - . - ... ,c ~ .C: C C O O O O O O O O O O C
O ~ ~ ~ C L
~ ~ ~ O ~ c v a ~, ~ + al o c_ _ L
~1 ~ " ~ ~
o ''~ ~a ~ ~ o ~o co ~ C ~ I` I ~ "1 ~~
~ c ~ ~ a- o u ~ ~ .o c o L~ ,~ Ln ~ ~
~ ~ ~ ~ o~
o~O~ c~ v 3~--,; ~. L a~ o _ L C V~
~ 8 E ~ t:u ~ ~
~ NL E tu ~ L, _~ e:l X ~11 O C~ o E a1 ~ ~ ~ r~ O ~ O U I~ 0 ¢ ~ ._ ~ o ~ u ~ -~ L
E~ ~ ~ ~ 3 c L
L~ 8 ~ u u~
U ~ L ~ CO ~ CO ~ O N ~.0 r~ o ~o L ~ E 5- c ~_ L ~
~ ~ E ~1 ~ o `~5 o o o o o o o o o o ~ CL o ~ .C
_ ~ O ~ ca ~ c E
~ 3 ~ L~ .~
O '~
O :~ 'E7a E ':7 ~ c _ ~ _ ~ ~ a~ ~ ~ L

. ~ ~ ~ o ~ ~ _ O O _ O ~1~ ~ ~ ~ O._ ~ U~ ~ _ U7 '_ ~ O O.
Z CL_ V ~ O ~
~ ~ ~ 0 ~ _ ~ '- G~
~ _~_ C ~'~ ._ ~ C~J_ O ~
e a1 o.c x ~--._ _ V ~ O
L 13 O a~ O ~ 41 e C~l o o a~ ~ z ~ ~C ~ ~: a:
~ el -144- ~ 73 The analyses of the total reaction mixture by packed column GC
show that, concurrent with the decrease of the percentage of the Cll feeds, mostly C12 aldehydes formed. There is very little aldehyde dimer and trimer formation; only about 3X of the main aldehyde products Analyses by capillary GC show that the two main products are n-dodecanal and 2-methylundecanal derived from the l-n-undecene component of the feeds. As it is shown by the table, the ratio of these two products in the final reaction mixtures is in the 2.7 to 3.1 range. There are of course, other branched aldehydes present. These are derived fro~ internal and branched olefins. Thus, the amount of the completely linear aldehyde, l-n-dodecanal, is ranging from 37.7 to 39.4Z of the total C12 oxygenated products. l-n-Dodecanal and 2-methylundecanal together represent 48.2 to 51.9%. The rest of the product contains major amounts of other, monobranched 2-alkyl substituted C12 aldehydes such as 2-ethyldecanal, 2-propylnonanal, 2-bueyloctanal and 2-pentylheptanal. These monobranched aldehydes were apparently derived from isomeric linear internal undecenes.
Only minimal amounts of the aldehyde hydroformylation products were reduced by hydrogen to give the corresponding alcohols. The only identifiable alcohol by-products were n-dodecanol and 2-methyl-undecanol. Their combined concentration was only l to 3X of that of the total aldehydes.
The three reaction mixtures obtained in the above described three examples of Cll coker distillate hydroformylation were worked up in a manner similar to that described in Examples 41 to 45.
It was noted that the reaction mixtures derived from the vacuum distilled Cll feeds were of definitely lighter brown color than that from ths atmospheric distillate feed. The removal of cobalt by the usual aqueous aceeic acid, air treatm2nt reduced the color of all the mixtures.
However, the difference between the now generally lighter colored mixtures persisted. All the mixtures were clear, free of any precipitate.
The cobalt free reaction mixtures were fractionally distilled using a 2 ft. packed column, at about 0.1mm pressure. The unconverted feed components were distilled at close to amblent temperatures (20 to 30C).
The aldebyde product was obtained between 57 to 67C. Both distillates were light yellow, clear liquids. Due to the relatively low distillation temperature of the aldehyde products, relatively little aldehyde dimerization and trimerization occurred during distillation. The residual dimers were only about 2.5X of the total oxygenated products formed. The -145- ~2~ 3 trimers were less than lZ although it is noted that their accurate determination by GC was not possible.
Examples 52-55 Hydro~ormylation of C12 Gas Oil with Cobalt in the 110 to 150C Temperature Range A series of four hydroformylation experiments was carried out with a previously described, vacuu~ distilled combinsd C12 fraction of gas oil in a manner described in Examples 44 and 45 to determine the effect of temperature on reaction rate and selectivity. Each run wac carried out using 1/1 H2/CO at 3000 psi. The reaction temperatures employed were 110, 120, 130 and 150C. The reaction mixtures were sampled at intervals and analyzed by packed and capillary GC as usual. The results are summarized in Table XXX U-The results of the table show that the C12 fraction was morereactive than the lower boiling fractions produced by the same Fluid-coker unit. About O.lX cobalt was found effective in the first three examples of the present series, while 0.2 to 0.4X cobalt was required in the previous axperiments.
As the temperature was increased from 100 to 130C in Examples 52, 53 and 54, the reaction rate significantly increased, At 150C in Example 4, only 0.05Z cobalt was used. Hydroformylation occurred, nevertheless, indlcatlng increased activiey. The composition of the final reaction mixtures indicated that in the hydroformylations 130 and 150DC, at about 1/3 of the feed was converted to aldehydes.
It was found that selectivity of hydroformylation to produce a high n/i ratio of the two major aldehyde products decreased with increasin~
temperature. Also, more aldehyde d$mer by-product and alcohol hydrogenation products were formed at 150C than at lower teMperatures.
For the selective production of aldehydes with good 012fin conversions, temperatures in the order of 130C are preferable. The daea indicate that, in general, the l-n-dodecene is selectively hydrofor~ylated at first, producing a high ratio of n-tridecanal and 2-methyldodecanal.
Thereafter, the linear internal olefin components are con~erted to various 2-alkyl substituted aldehydes. Concurrently, hydroformylation of the minor branched olefins also occurs to give some further branched aldehydes. Thus, with increa~lng conversion, product llnearity decreases. For e~ample, at -146- ~99~73 , _ _ o~ ~
_ . o oo ~ o o o V) , C
E
r~ O-- ~
+ ~ ~ o o ~1 C I~ ~ r~ ~D r--Ln 1~ In _ ~
L ~ ~ ~ 3 ~_ ~ C_~ ~, O V~
~;;5 C~ ~ ~ _ O el~ O ,~
e7 c ~ a~ I ~ C
C U'~ ~ ~ O cn a ~ ~ o CL 3 l ,~ O ~
~ _ ~ .. O . .. .. ~ a~ ~ =
~$ ~ ~ ~ ~ ~ ~ ~ "~
~ ~ ~ ~ aJ C
--4................................. , ,~, ~ ~ o U~ ~ o O E O o OO o r~_ oc~
Y ~ c a ._ _ L
.~ ~ t 1~ O G~ ~
J X 7~ ~JJ ~ ~V:~ r.~ O ~ ~ r.~ N U ~
0 4~ ~ ~ C q) ai ~
. J~ X C~ ~ o .~ "., ,- t ~ ~ ,, a. E~ e rD U ~ c ~_ ~ o ~ i~ o~ o ~ ~ ~ ~
~ ~3 , ~ cn ,~
Vl_ ~ _ ., ~ a~ o o oo o o o o o o o ~ c C-- E c ~ O 0 ~ D ra 1--a: --~ --~ -- --~ c E U
~ ................................. 3 ~ o ~
~- ~ E
~ ~4 _ ~ c g ~ v O O o o o ~ a~ c U~ O c _ ~ ~ L
O _ ~ ~ 00~0 0 _ al ô ~ ô c~
~3 ~ r ~ ~ U U _ O O
V C ~ E ~ ~' N O c n~ O I ~ ~ ~o~
~ ~ 0 Q~ IL q- ~ O ~ , E a o c~: O ~ J ~ ~ ~ a.l c ~ O
c ~? e ~ ~
~1 ~ o ~ ~ o ~ o O _ . _ _ o~ o_ .~

CL ~ Q~ L
O ~ ~
CL ~ Vl v~
e o ~ ~ ~u~ c~ E ~ O
X u~ n ~_ ~ X :~E L

-147- ~2~73 130C in Example 54, the percentage of n-tridecanal decreases from 55.6 to 44.lX a~ the percentage of unconverted feed drops from 73 to 66~.
Example 53 additionally shows a low temperature ~eneration of the active catalyst species from a cobalt carboxylate rather than dicobalt octacarbonyl. In this Example, the use of cobalt naphthenate at 120C
resulted in approximately the same conversion as that of Co2(CO)g at 110C
in Example S2.
The four reaction mixtures of these four examples were worked up to isolate the products in a manner similar to that of Examples 44 and 45.
All the hydroformylation product mixtures were clear dark brown liquids, free from precipitates. They were readily decobalted with aqueous acetic acid plus air treatment in the usual manner. The cobalt free mixtures were lighter brown. They were worked up separately.
Fractional distillation of the cobalt free mixtures yielded almost colorless distillate fractions of unconverted components and colorless to light yellow C13 aldehyde products. The aldehyde products were distllled using a 1-l/2 ft column between about 70 and 80C under about O.lmm pressure with an oil bath of 130-160C. Durln~ the slow distillation of about 8 hours, significant additional unsaturated aldehyde dimer formation occurred. l'his was the ma~or factor in determining the isolated product yields. If alcohols are the desired products, hydrogenation of the decobalted reaction mixture prior to fractional distillation is preferred.
The dl~tillate aldehyde products of the four examples were combined to provide sufficient amounts for subsequent hydrogenation.
According to capillary GC, the combined product contained 40X
n-tri-decanal, 14.4% 2-methyldodecanal and 17.6X of 2-alkyl substituted aldehyde~ plu~ minor amounts of alcohols in the order of 2X.
The detailed structure of the isomeric aldehydes is illustrated by Fi~ure_l~ which shows the aldehyda region of the capillary gas chromatogram of a reaction mixture. Based on GC/MS studies, the figure indicates that besides the ma~or n-tridecanal, the 2-methyl and higher 2-alkyl branchcd isomeric aldehydes are present in decreasin~ a~ounts.
Mass spectrometric studies also showed that 2-methyldodecanal 3-methyl-dotscanal are present in comparable amounts.
Example 56 Hydroformylation-Acstslization of C12 Light &a~ Oil with Methanol ~ rho Pre~qnc~ O.lX Cobalt in tha 120-150-C Ran8o A C12 Fluid-coker naphtha fraction of bp. 207 ta 217C was hydrofor~ylated in a one to three molar mixture with methanol at 3000 psi (207 atm), in the presence of O.lZ cobalt added as a toluene solution of Co2COg at 130C. The reaction took off immediately, and proceeded at a faster rate than without the added methanol. Nevertheless, to complete the reaction of branched olefin components, the temperature was raised to 140C
after 2 hours and to 150C after a total of 4 hours. The reaction was discontinued after a total of 6 hours. GC analysis of the reaction mixture after standing at room temperature showed a highly selective formation of the dimethyl acetal derivatives of the C13 aldehyde products and negligible dimer formation.
The reaceion mixture was diluted with aqueous methanol to separate the cobalt and then was distilled in vacuo. The dimethyl acetal of the tridecanal hydroformylation product was distillPd using a 2 ft.
packed column and obtained as a clear colorless liquid between 80 and 85C
at 0.05 mm. Capillary CC/MS indicated that the isomer distribution was similar to that observed in the absence of methanol.

Examples 57-60 Hydroformylation of C13 Gas Oil with Cobalt in the 130 to 170C Tcmpersture Range A series of four hydroformylation experiments were carried out with a previously described, vacuum dis~illed combined C13 fraction of gas oil in a manner described in Examples 41 to 45. The reaction conditions were the same as those in the previous example. The experiments were to determine the effect of increased reaction temperature up to 170C. The results are summarized in Table XXXIII.
The data of the table show that the rate of the reaction increased right up to 170C. This is in contrast to the hydroformylation behavior found in studies of the C~ naphtha fraceion.
As it is indicated by these data, reaction temperatures below 150C were ad~antageous for the selective production of aldehydes (Examples 57 and 58). The percentage of dimer and trimer by-products increased with the temperature. At 170C, major amounts of alcohols were formed ~Example 60).

1294~3 I O
~ , o _ ~ o ~ ~ ~1: _~
.", ~ o ~1 O .+ D O ~U~ ~ O
~a ~ o c u~
~;~ ~ _ ~o ~o ~o U~ U
o ~Ll_ O _ ~ O oO U~ r_ et ':~ ~ O U
L~ ~~Y _~
~_ V) ~ ~ _ .
o U ~ ~o o E ~ o ,, Oa~~ .t,~ 00 00 0_ ~" O

L
.~X ~ O O ~ 30 E~ ~ _ ~ ~ ~ r _ U~ ~ CO ~ ~ ~ ~
, ~ ~ ~ ~ ~ ~ ~ ~ ~ ~0 -C
CC ~ ~
c 0~I ~ cooa ~ oln D~ C ~
~~ O~ I~ C
'I ~ ~-~0 C O
~ S -- ~ LL. ~O $ O f'- ~ ~
C o C~ ~0 O ~ ~_ ~
o a~0~ ~0 0 ~O ~G ro _ _ _ ~ lCy oJ ~

X I~ ~ ~ ~D

-150~

It was also observed that the percentage of the n-aldehyde component of the total aldehyde product decreased with the temperature.
Thus, the data show that reduced reaction temperatures result in ircreased product linearity and decreased by-product formation. It should be noted, though, that the sharply decreased n-aldehyde content of the 170C reaction mixture is largely due to hydrogenation to n-alcohol. At 170C, aldehyde formation is essentially complete in 60 minutes. Thereafter, the prevalent reaction is aldehyde hydrogenation to alcohols.
All the hydroformylation product mixtures were clear brown liquids, free from precipitates. They were readily decobalted with aqueous acetic acid plus air treatment in the usual manner. Some additional dimerization of the aldehyde product occurred during distillation at O.lmm using a 2 ft packed column and a heating bath of about 135C. The aldehydes distilled between 75 and 85C at O.lmm.
It was interesting to observe during the distillation of the reaction mixtures, that the color of both the unconverted componen~s and the aldehyde products were dependent on the reaction temperature. The mixture from the 130C reaction yielded yellow distillates of both unconverted gas oil components and aldehyde products. The mixtures of the 140 and lS0C reactions gave colorless hydrocarbon distillates but yellow aldehyde products. The 170C reaction mixture yielded colorless distillates of both hydrocarbon and aldehyde fractions.
The above observations indicate that during hydroformylation, double bond hydrogenation and, probably, desulfurization via hydrogenation become increasingly significant side reactions with increasing reaction temperatures. It is felt, though, that these hydrogenations are better carried out during the subsequent hydro~enation of the reaction mixture which provides the usually desired higher alcohol product.
The distilled aldehyde products all contained tetradecanal and 2-methyltridecanal as the major components. As it was also found in the previous examples, other 2-alkyl substituted C14 aldehydes, when combined, constituted the third group of product components. It was shown by GC~S
studies that the 2-alkyl substituents of these aIdehydes ranged from C2 to C6 n-alkyl.

-151~ 73 Examples 61-63 Hydroformylation of C14 Ga~ Oil w~th Cobalt in the 110 to 130C Temperature Range A series of three hydroformylation experiments were carried out with a previously described, vacuum distilled combined C14 fraction of gas oil in a ~anner described in Examples 44 and 45. The rPaction conditions were the same as in Examples 52 to 54, however, the amount of cobalt catalyst used was increased from 0.1 to 0.3X. The results are shown in ~!
The data indicate that the reaction rate was the smallest, but product linearity was the greatest, at 110C, the low temperature of Example 61. Conversely, at 130C, i.e., the high temperature of Example 49, the reaction rate was the greatest but product linearity was the smallest. Since the reaction temperatures were relatively low in all three examples, there was no significant aldehyde dimer and trimer formation. The amount of alcohol hydrogenation by-products also remained low, around 3~ of the aldehydes.
The product linearity is best indicated by the percentage of the n-aldehyde (and n-alcohol) in the total oxygenated products. At the end of the hydroformylation, this value was 45.2X at 110C, 42.2X at 120C and 40.8X at 130~C. The percentage of the l-n-olefin derived n-aldehyde was inversely dependent on the hydroformylation o the less reactive internal and branched olefins which provide branched aldehydes. Thus, the n-aldehyde percentage was inversely proportional to the total olefin conversion.
Tha n/i ratio of the two main aldehyde products, n-pentadecanal to 2-methyl-tetradecanal, was more independent of olefin conversion since both of these products can be derived from the reactive l-n-olefin component, l-n-tetradecene. (2-Methyl-tetradecanal can be also derived from 2-tetradecene). This n/i ratio was largely dependent on the temperature. It was inversely proportional to it as it is indicated by the data of ~he table.
The data of these and the previous examples suggest thac a preferred method of hydroformylation is c~rried out at variable temperatures wherein the l-n-olefin component is substantially converted at 130C or b~low, and the other olefins are malnly reacted at temperatures exceeding 130C up to 170C. Such a variable temperature operation can bs o~
._ ~ ",I ' ' ' ' o CL ~ o~ o .
L ~ L O C I~ ~ r~ r~ ~ I~
_ ~ E _ ~ ~ ~ ~ ~ U'~
_ O ~ 0 ~ O 1~ 0 U~ ~ ~ U ~
O ~! -~IJQ
_ ,~ I r ~ C~ L
~- ~ c ~ o 01~ Cl. I ~r~ ,~,~, ~ ~'_~
~ _ C~ O

r~ ~ ~ c r~
~~ L ~ E ~N u~ O r~ a~ O _ _ ~> L
X~1 ~ C~1 E ~ c~ J O O ~ O O .-U8 ~ O _ -c lLv) X C-~ ~ cO ~ -- ~ '- ~ C
_ ~ ~ Oa~ ~ 2_ a) ~ ~ ~ ~ c ~ ~ ~
~ O ~ ~ C O L
~ 3 ~ ~ _ ~
O ~ .CI ~ ~ L E
d' , ~ a.1 o 30 æ ~ aLI ar~ O
0 0 c C ~ ~-- ~ ~ æ

0 ~ o ¦ ~ O ~ _ C ~IJ

,C ,, U ~ ~ _ '~ X ~ a~ ~
LL ~ ~ ~ ~

'73 carried out in reactor system comprising reactors operating at different temperatures.
All the hydroformylation product mixtures were decobalted with aqueous acetic acid plus air treatment in the usual mannçr, and then fractionally distilled in vacuo. The Cls aldehyde product was obtained as a clear yellow liquid distillate boiling between 95 and 111C at O.lmm.
Using a relatively low temperature bath of 120-140~C, relatively little, about 5X, of the aldehyde was converted into dimers and trimers during distillat~on.
Analyses of the distilled Cls aldehyde product showed that it was essentially free from hydrocarbon impurLties. Combined GC/~S studies lndicated the presence of about 47X n-pentadecanal, 15.5X 2-methyl-tetradecanal and 16X 2-(C2 to C6 alkyl) substituted aldehydes. A distinct dibranched C16 aldehyde was also found in the ~ixeure in about 7.9X
concentration. Minor amounts (0.5X) of n-pentadecanol were alsa presene.

Example 64 Hydroformylation of a C14 Fraction o L~ght Gas Oil by H2/CO
~ith Cobalt under 3000 ps1 Pra~sure at Variable Temp~ratures and the Hydrogenation of the Cls Aldehyde Product The C14 olefinic feed for the present hydroformylation was separated from a light Fluid-coker gas oil by a double 15/10 type distillation. It was hydroformylated either in the presence of 0.2 or O.lZ. As a catalyst precursor, Co2~CO)g was used. It was introduced at 120C as an approximately 6Z solution in isomeric xylenes. The temperature was increased from 120 to 150C during the course of the reaction to convert the various types of olefins at their minimu~ reaction temperature.
The results of both hydroformylation experiments,using 0.2 and O.lX cobalt, respectively, are shown by Ta~e X~- Good olefin conversion was achieved at both catalyst concentrations. The maxim~ aldehyde content of the reaction mixtures was about 30X. However, the n-aldehyde selectivities appeared to be slightly higher at 0.1~ Co.
The decrease in pressure drop with the reaction ~ime indicated and the composition of the reactlon mixture by packed GC showed that the reaction was essentially complete in 4 hours. As expected, the reaction was fastcr at the higher catalyst concen~ration. The final reaction mixtures still contained only minimum amounts of by-products; in the range of 2 to 3% of dimers.

-156.- ~ 73 -- -- c a ¦ o ID ~
O ~ ~. G O " ¦ ~ , N ~-- ~

~ ~ " u ~ ~ o ~ ~ I .~
,0 ~ ~ æ ~1 l ~ ' g _ . .13 C ~ C ~ ~ S J t~~ N N ~ ~
~ O w ~ O I IO O S, O ~r~ N ,~ ~ r. -- C
C _ O L
L ~. O C~ O ~ o~ ~ o ~
O S~ 'O I ~ . ~ o O
1~ o ~ c r I N N _1 N N N ~ L

~~ g ~ L N~ ~ ~ _ N ~D Qa ~ ~t C
5a J o 1~ a N Od ~ ~> N ~ ~ r ~c ~ ~ ~ ~ 3 E--~ c ~_ L 8 O O ~ ~ Ul O~ ~ ~ i~ V~ O _ C
O ~- L ~ ~ 11~1 ~ N N ~ ~ q C ~1~ C

L _ Cl Q O O ~ O ~ 1 ~ O
O ~ _ o ;~ ¦ 5 ~ ~ ~ ~ ie ~

~ ~ ~ ~ o ~
_ ~ S ~ O _ _ N N _ _ N ~-- a e Q 1_ ,0o n ¦o ~ O o o N ~- D
o " O ~ ~cn ~ ~ o ~ oo o o o o .,n IC e_ 11~ N N NN ~ o ~ O IZ~ _ ~
~ ~ e~ 3 ~ ~o ~ ~
o ~ o ~_ ~ _ ~ o ~ro o ~ ~ o 10 ~ i o ~ ~ e n c~ o ~ ~ L
~ e c~~ eI o o o o o oo o o e o o ~ o 4~ V~ ~ o o ~9 ~ O o O 06:~ o . ~ ~ o . _ ~ ~ ~ ~ ~ ~ L
o c o ~ 1~ e ~ s~ v I o i -- ~

-155- ~.4~ 73 Capillary GC showed tht the n-pentadecanal was the most prevalent Cls aldehyde isomer forred. It was, of course, mostly derived from the main and most reactive olefin component of the feed l-n-tetradecene. Thus, its percentages, 56.0 to 48.9X, were particularly high durin~ the early stages of the reaction. At the completion of the reaction, the n-pentadecanal concentration was 33.1 to 36.1~ of the isomeric pentadecanals.
The monobranched pentadecanals derived from and were the largest group of branched isomers. The percentage of the largest branched Cls aldehyde isomer 2-methyl-tetradecanal ranged from 12.0 to 13.5. The second largest isomer, 2-ethyl-tridecanal was present in concentrations ranging from 5.7 to S.8X. The other monobranched Cls aldehydes derived from n-pentadecenes were also present in concentrations ranging from 1.8 to 3.8~. As expected, these minor aldehyde isomers had n-propyl, n-butyl, n-pentyl and n~hexyl branches in the 2-position. The largest of these minors, 2-propyl-tridecanal was present in the 2.1 to 3.9Z concentration range. It is noted that at low feed conversion ehe recorded relative concentrations of these minor isomers in the table are low because of the limitation of the GC method of determination.
The selectivity of the hydroformylatlon of the l-pentadecene component is characterized by the ratio of n-pentadecanal to 2-methyltetradecanal (n/Me). This ratio is decreasing from a top value of about 3.4 down to 2.7 with increasing reaction temperature and olefin conversion. The overall Cls aldehyde linearity is described by the ratio of the normal isomer to the sum of all the iso-, i.e. branched, aldehydes.
This ratio also drops from 1.27 to 0.49.
Capillary GC also showed the presence of significant amounts isomeric Cls alcohols and Cls formates in the reaction mixture. Their GC
peaks partially overlapped, ~ut n-pentadecanol and n-pentadecyl formate could be distinguished. The combined amounts of alcohols and formates in the final reaction mixture ranged from 14.5 to 17.0X of the eotal oxygenated products. The ratios of the n-alcohol to the n-alkyl formate were from 2.4 to 3.5 Sulfur GC of the reaction mixture indicated that most of the sulfur compound components are in the retention time region of the C14 hydrocarbon feed. Relatively very small ~mounts of sulfur were found in thc aldehyde region.

The hydroformylation reaction mixtures were decobalted with aerated aqueous acetic acid as usual. Then they were hydrogenated in the presence of 10% of a CoS/MoS based catalyst and 5~ water at 3000 psi (306 atm) in the 150 to 170C temperature range. The reduction of the aldehydes was co~plete in 20 hours. Sulfur GC indicated that most of the sulfur compounds in the feed region remained unrhanged during hydrogenation.
The hydrogenation of the aldehyde components of the reaction mixture was studied in some detail. Hydrogenations were carried out under comparative conditions at 150, 155 and 160C under 3000 psi pressure in a 1 lieer stirred autoclave with reaction mixtures described above Samples taken after 2, 5 and 20 hours were analyzed for aldehyde and alcohol content by capillary GC. The results are shown by Table XXXVT.
The data of the table show that hydrogenation occurs at a moderate rate at all three temperatures. The conversion of the n-pentadecanal to n-pentadecanol is 96X or more. The conversion of all the isomeric aldehydes is about 85 to 90Z as a minimum. (The total conversion could not be exactly deter~ined because of GC peak overlap between aldehydes and alcohols. This overlap of minor alcohol components was dlsregarded and all the components having shorter retention times than n-pentadecanal were counted as aldehydes.) It is noted that the data of Table XXXVI show a slightly decreased aldehyde to alcohol conversion with increased temperature. This is apparently due to a slight decrease in catalyst activity in the three experiments of increasing temperature. It is interesting eo note that the amount of n-pentadecane secondary product derived from n-pentadecanol increased with the temperature from a level below detection to an amount equal to 1.3X o~ that of n-pentadecanol. It appears that the n-alcohol to n-paraffin conversion is more temperature dependent than the n-aldehyde tQ
n-alcohol conversion.
The hydrogenated reaction mixtures from the hydroformylation of C14 Fluid-coker light gas oil frac~ion were combined and worked up to isolate the isomeric Cls alcohol products. The combined mixture was then washed with fifty volume percen tof a lOX aqueous sodium hydroxide solution to remove the H2S and any carboxylic acid by-products. This wash resulted in e~ulsion formation. The emulsion phase was largely broken by the addition of xylene. The organic phase was then washed with water and dried over anhydrous MgSO4. The xylene solvent was then mostly removed by film Table XXXVI
~drogenation of th~ Cls Ald0hyt~ Co~ponont3 of the Hydrofor~yl~ion Reactlon ~ture Derivod fro~
C14 Fluit-Coksr Nsphtha fro~ Bllling~A) Convsrsion to C15 Alcohol Products Based on Alcohol Product to Aldehyde Precur~or Ratio . _ ~ .. . . ._ .
At 150-C At 155~C Ae 160C
(E-7630) (E-7633) (E-7636) Action Alcohol X Alcohol X Alcohol X
Ti~eb Wor- To- Nor- To- Nor- To-Hrs.malC tald malC tald malC tald ~9~ 90 9~ 88 96 85 a) Hydrogenation with 6 wtZ CoS/MoS catalyst on alumina support in the presence of 5X water under 3000 ps L pressure b) From the time the mlxture reached reaction temperature c) The percentage of n-peneadecana]. converted to n-pentadecsnol d) The percentage of Cls aldehydes converted to Cls alcohols based on oxygenated co~pounds up to and including n-pentadecanal.

evaporation in vacuo and the residual liquid was fractionally distilled in vacuo.
Fractional distillation reco~ered the unreacted C14 hydrocarbons as a clear colorless liquid between 61 and 72C at about 0.1 mm. The oxo-product residua was about 1 kg. On further distillation the Cls alcohol product was distilled betwen 112 and 129C at 0.1 ~m. About 70X of the oxo-products were obtained as the C1s alcohol distillate, a colorless clear liquid. Another 17X was recei~ed as a yellow liquid distillate mixture between 129 and 245C. This mixture contained C1s alcohols and dim0rs in a 1 to 2 r~tio. The residue was about 12X.
During the distillation of ehe Cls alcohol, it was noted that the n-pentadecanol crystallized on cooling from the higher boiling more linear fraction~. Linear detergent alcohols can be apparently isolated from the present alcohol mixtures by crystallization.

Example 65 Hydroformylation of Cls Gas Oil wlth O.lX Cobalt at 140C
~ The previou~ly described, vacuum distilled combined C1s fraction of gas oil was hydroformylated in a manner described in Exa~ples 41 to 45 at 140C under the conditions of Examples 49 to 52. The results are s = arized in Table,~
Table XXXVII
Hyt~oformylation of C1s Ole~inic Fraction o~
G~s 011 from a Fluld Cokor in tho P~osonco of O.lX Cobalt C~taly~t Dorivod from Co2(CO)g uith 1/1 ~2/CO at 3000 p5i Reaction Mixture Time CQmponents~ X Two Ma~or Min Un-Aldehydesb Products fA~o~ n/i,RatioC
973 3.14 180 8911 2.87 360 7129 2.76 aDetQrmined on packed column GC.
bMostly aldehydes.
Cn-Hexadecanal to 2-methylpenLadecanal -159~ 7~

The data of the table show that at the low concentration of catalyst used, thare was a long induction period. After 1 hour reaction eime, less than 3~ of aldehydes were formed. In three hours, product formation was still minimal. The maximum rate of hydroformylation was reached after 4 hours as indicated by the rate of synthesis gas consumption. A complete conversion of the l-n-pentadecene feed component was obtained in 5 hours. After 6 hours, the amount of products in the reaction mixture was 29X and gas consumption was low. Thus, the reaction was discontinued.
Analyses of the reaction products showed high selectivity to aldehydes. The amount of alcohols and dimers each was about lX in the final reaction mixture. The main reaction produc~s were n-hexadecanal and 2-methylpentadecanal in an n/i ratio of 2.76. These two products amounted to 73.5X of all the C16 aldehyde products. Most of the rest were 2-alkyl substituted C16 aldehydes.
The final reaction mixture was decobalted as usual and fractionally distilled at 0.lmm to separate the C16 aldehyde product. The aldehyde was obtained as a clear yellow liquid distillate, boiling between 115 and 125C at 0.1mm using a heating bath of 150 to 160C. During fractional distillation, significant aldehyde dimer and trimer formation occurred. Only 70X of the C16 aldehyde present in the reaction mixture was recovsred by distillation.

Examples 66-70 Hydro~enation of the Cll-Cls Aldehydes Derived from Coker Distillates to Produce the Corresponding Alcohols The combined distilled Cll to Cls aldehyde products were hydrogenated in the presence of a sulfur insensitive cobalt/molybdenum based hydrogenation catalyst in the manner previously described in the Experimental procedures. After about 24 hours hydrogenation at 232~C under 300 p9i pressure, the reaction mixtures were analyzed by GC/MS for aldehyde conversion. (In the case of the Cls aldehyde, the reaction time was 48 hours.) It was found that the aldehydes were completely converted. The products were mostly the corresponding alcohols. However, some conversion to paraffins also occurred, possibly via the main alcohol products.

CnH2+1CH0 H23- CnH2n+lCH20H 2 ~ CnH2n+lCH3 n 10-14 The product distributions obtained in the Examples are listed in the following:

Ex~mple Carbon No. Product Distribution Z
Number of P~oduct Alcohl Paraffi~

6g 13 88 12 6g 14 89 11 An examination of the isomer distribution of the paraifin by-products by GC/MS showed a higher ratio o normal to iso paraffins than the n/i of the parent aldehydes. This indicated that the n-aldehydes and n-alcohols were preferably hydrogenated to paraffins. Consequently, the percentages of the n-alcohols, and the n/i ratios of n-alcohols to 2-methyl substituted alcohols, somewhat were lower than the n-aldehyde percentages and the aldehyde n/i ratios of the feeds. Since the hydro~enation to paraffins was a minor side reaction, the order of decreasing concentrations of alcohol types (normal, 2-methyl substituted, 2-ethyl and higher alkyl substituted alcohols) remained tbe same as that of the aldehyde feeds.
The reaction mixtures of the hydrogenations were fractionally distilled to separate the alcohol products from the paraffin by-products.
Both were obtained as colorless liquid distillates of the following approximate boiling ranges:

9~

Boiling Ran~e__C~mm..
Carbon AlcoholParaffin.
Number ProductBv-Product 11 135-1~6-2097-132/20 GC/MS studies indicated that the alcohols had qualitatively the same Lsomer distribution as the parent aldehydes. The n-alcohols and the 2-methyl branched alcohols were the main components. GC/MS showed that the paraffins were derived from the aldehyde feed without structural isomerization. The paraffin forming side reaction occurred a~ the highest rate in case of the linear aldehyde component of the feed as indicated by the predominant formation of the n-paraffin.

Semilinear DilLkyl Phthal8te PLaæticizers The Cs to Cls semilinear alcohols of the present invention can be converted to the corresponding dialkyl phthalate esters, via known methods.
The alcohols are reacted with phthalic anhydride, preferably in the presence of a non-oxidizing acid catalyst such as p-toluene sulfonic acid or an alkyl titanate. The rasulting phthalate esters have a unique combination of plasticizer properties as illustrated by the following examples.

Example 71 Semilinear Diundecyl Phthalate Plasticizer The semilinear undecyl phthalate, DUP-F, and a linear undecyl phthalate, Jayflex DUP, were compounded with a Geon 30 polyvinylchloride and additives in the following proportions: parts per 100 wt. parts of PVC
Geon 30, 100; phthalate plasticizer 50, Calcined Clay, 10; Dythal XL, 7;
Stearic acid, 0.2. The physical properties of the resulting plasticized ~ compositions ware then tested. The data obtained are the following:

:

-162 1~9~3 DUP-F Jayflex DUP
Hardness Shore D 36 39 100X Modulus, psi 1890 2190 Tensile Strength, p9i 3040 3110 Elongation, X 325 302 The highsr hardness and modulus of the DUP-F composition indicate decreased plasticizer effectiveness. To obtain similar physical properties, a hi~her amount of DUP-F i3 to be used. Since the plasticizer has a lower cost by volu~e than PVC, reduced plasticizer effectiveness decreases the cos~ of plasticized PVC.
DUI-F and Jayflex DUP were also compared in a plastisol tese in the following formulation: Geon 121, 100; Plasticizer, 70; Mark 7101, 2.
After aging the plastisolQ at 100F (38C) the comparative Brookfield viscosity data were:
DUP-F J-DU~
Cp5 after 2 hours at 3 rpm 4870 1970 30 rpm 36850 15750 cps after 24 hours at 3 rpm 6120 2060 30 rpm 41800 17250 To determine processability, a hot bench gelation test and dynamic mechanical analyse~ (at 10C/min and 1 rad./min and 1% strain) were carried out. The comparative results were:
~ J~DUe Gel point, C 247 264 Gel onset, C 73 81 Gel complete, C144 144 Fu ion co~plete, C 196 196 These data indicste a more facile processing for the Flexicoker alcohol bas2d, DUP-F plasticizer.
The color stability on heating at 350F (177C) was the same for the DUP-F and J-DUP composition. Only the low temperature properties of the semilin~ar DUP-F were inferior to those of the linear J-DUP. In this respec~, the properties of the semilinear DUP-F are in between those of the corre ponding branched and linear ester compositions.

~Z,g~3 Example 72 Semilin~r Didodecyl Phthslate Plssticizer The semilinear dodecyl phthalate, DDP-F, was compared as a plasticizer with branched ditridecyl phthalate, J-DTDP, and branched undecyl decyl phthalate, J-UDP. Plasticized PVC compositions were formulated as follows: Geon~30, 100; Plasticizer, 62; Tribass EXL (lead silicate sulfate stabilizer~; CaC03, 15; Stearic acid, 0.25; BPA
antioxid~nt, 1. After milling at 350F (177C) and molding at 360~F
(182C) the followin~ properties were found:

DD~ p~ UDP
Shore A Ha~ness, 7 day 84.7 84 84 Shore D ~ardne3s, 7 day35.5 38 37 Ori~lnal Phv~ica~s, 0.040"
Tensile Strength, psi 2483 2555 2584 100X Modulus, psi 1749 1856 1868 Elongation, Z 307 293 280 Aged Pbysicals, 0.040" 7 days Retained Tensile Stren~th, ~ 103 107 130 Retained 100X Modulus X 139 146 Too brittle Retsined Elongation, i 64 56 Too brittle Weight Loss, X 4.8 9.7 14.6 Clash Ber~, Tf, 0.070n, C -31 -23.3 -27 ~ , 0.070n, C -28 -21.9 -22 Pad V~lu e~Q~sçlyitv 0 30 1.68 1.94 0.040~, 90-C, ohm-cm x lon These results lndicate that the semilinear didodecyl phthalate is a fine plaseicizer. Its reduced weight loss and lower Clash Berg and Bell Brittlenes~ te~peratures indicate volativity and low temperatur~
charactoristics superior to related branched phthalate esters.

~ TR~ A Rl~

Example 73 Semllinear Ditridecyl Phthalate Plastlcizer The semilinear ditridecyl phthalate plasticizer, DTP-F, and a commercial branched ditridecyl phthalate plasticizer, DTDP, were compared in a PVC formulation described in the previous example with the following results:

DTP-F DTDP

Shore A Hardness 92 91 lOOX Modulus, psi 1740 1770 Tensile Strength, psi 2420 2500 Elongation, X 285 304 Aged Phy$ical. 7 days/136C
Retained Tensile, X 102 105 Retained Elongation, X 77 56 Weight Loss, X 2.8 9.6 Low ~ s Clash Berg, Tf, C -30 -23 Bell Brittleness, Tb, C -25 -22 Pad Volume_~esistlvity 40 mil, 90C, ohm-cm x 10" 2.2 9.5 The data indicate that DTP-F is a fine plasticizer with superior elongation retention and permsnence at high temperature and better low temperature properties th~n DTDP.

Semilinear Sur~actants The semilinear alcohols of the present invention are converted to novel surfactants via known methods. These methods are described in the appropriate volumes and references therein of the Surfactant Science Series, edited by M. J. Schick and F. M. Fowkes and published by ~arcel Dekker, Inc., New York. Volume 7, Part 1 in 1976 by W. M. Linfield covers "Anionic Surfactantsn. "Cationic Surfactants" are discussed in Volu~e 4 of 1970 by E. Jungermann. "Nonionic Surfactants" by M. J. Schick are in Volume 1 of 1966.

X' Example 74 Heptaetho~ylated Semilinear Alcohol Surfactants Semilinear C13 and C14 alcohols of the present invention were ethoxylated in the presence of sodium hydroxyde as a base catalyst to provide nonionic surfactants with an average of 9 ethoxy groups per molecule.

C13H27(0CH2cH2)9oH and C13H27(0CH2cH2)9oH

In contrast to the sluggish and incomplete ethoxylation of branched aldol alcohols, these ethoxylations proceeded readily to completlon.
r .~ These surfactants were ~hen compared with a similarly ethoxylated linear C12 to Cls alcohol, Neodol 25-7.
The surface tension, in dynes per cm at 78F (25C), of aqueous solutions of these surfactants at various concentrations according to the ASTM D-1331 test method were the following:
O.OOO~X O.001% 0.2lZ

C14-F07 61 . 38 35 Neodol 25-7 5 34 30 These data indicated similar surface tension reductions at the practical concentratlons~
The cloud points of the lX aqueous solutions of these semilinear and linear surfact~nts by ASTM-D2024 also similar.
Cloud_Poin~
C F

Neodol 250746 116 The ambient temperature wetting times of O.lX aqueous solutions of the semilinear surfactants were superior to those of the linear surfactants; according to the Draves test (ASTM-D2281):

~RP~ m~

Wettin~LTime, Seconds Below Above C~ou Cloud C13-F07 8.5 7.4 10,0 C14-F07 6.7 11.8 12.3 Neodol 25-7 14.0 11.0 13.5 Aqueous solutions of the semilinear surfactants had a definitely reduced foam stability according ~o the Ross-Miles test:
Foam H~ hhL, ~ 8_L~Z E_¦b~LL--_0.1~ l.OX
Init~l 5 Mi~. Initial $ Min, Neodol 250-712 6 15 12 This is advantageous in many nonfoaming applications.

Example 75 Nonsethoxylated Se~ilinear Alcohol Surfactants Semilinear C13 and C14 alcohols were eehoxylated to provide surfactants of incr~ased hydrophilic character with an average of 9 ethoxy groups. These surfactants C13-FO9 and C14-FO9 were compared with a simllarly ethoxylaeed linear alcohol, Neodol 25-9, in ehe tests of the previous example. The results are shown by the following tabulations:
~'~
O,0001% ~Q~l~ ~l%
C13-FOg 52 40 33 Neodol 25-9 :S6 34 32 Cloud Point t 1%
' ~eodol 25-9 74 165 -167- ~ 73 Uettin~ ~ime. Seconds Below Above _ CloudCloud Neodol 25-7 19 Foam HeiRh~cm at 122F (50C~
. _ 1,~ 1 . ox Initial 5 M~ Initial 5 Min.
C13-~07 15 3 21 6 Neodol 250-7 15 12 17 13 It is apparent from the test results that the surfactant properties of the novel semilinear alcohols can be advantageously changed depende~t on their carbon number and degree of ethoxylation. Since the presen~ semLlinear alcohols can be readily produced with even and uneven carbon numbers of choice, they can often provide surfactants of optimum properties without added cost, Both the semilinear alcohols and their ethoxylated derivatives wera readily sulfated and sulonated to provide anionic surfactants of similarly attractive surfactant properties.
This invention has been described and illustrated by means of specific embodiments and examples; however, it must be understood that numerous changes and modifications may be made within the invention without d-p r~ing fr les spirlt Ind s~opn ~s deElned Ln che clalms w~ich follow.

Claims (26)

1. A hydroformylation process comprising reacting an olefinic cracked petroleum distillate feed in the naphtha range, produced from petroleum residua by high temperature thermal cracking, wherein said distillate is a narrow fraction which as a result of fractional distillation contains mostly olefins of the same carbon number, more than 30X
of said olefins being of Type I, and a substantially reduced thiophenic sulfur concentration with carbon monoxide and hydrogen at temperature between about 50 and 250°C and pressures in the range of 50 and 6000 psi (3.4 and 408 atm) in the presence of a Group VIII transition metal carbonyl complex catalyst in effective amounts to produce aldehydes and/or alcohols of semilinear character having an average of less than one alkyl branch per molecule.
2. The process of Claim 1 wherein the distillate feed is selected from the group consisting of olefinic C6, C7, C8 and C9 fractions.
3. The process of Claim 1 wherein the reaction is carried out at temperature between about 100 and 180°C and pressures between 2500 and 6000 psi in the presence of a cobalt carbonyl complex catalyst.
4. A hydroformylation-acetalization process comprising reacting an olefinic cracked petroleum distillate feed, produced from petroleum residua by high temperature thermal cracking, and containing 1-n-olefins as the major type of olefin components, and organic sulfur compounds in concentrations exceeding 0.1% sulfur, (1) at first with carbon monoxide and hydrogen at temperatures between about 50 and 250°C and pressures in the range of 50 and 6000 psi (3.4 and 408 atm) in the presence of a Group VIII transition metal complex catalyst in effective amounts to produce aldehydes of a semilinear character having an average of less than one alkyl branch per molecule, (2) then with a C1 to C6 alcohol at temperatures between 15 and 250°C and pressures between 0 and 5000 psig (0 and 340 atm) during or after said hydroformylation to produce from the aldehydes the corresponding acetals of a semilinear character having an average of less than one alkyl branch per molecule.
5. A hydroformylation-hydrogenation process comprising reacting an olefinic cracked petroleum distillate feed, produced from petroleum residua by high temperature thermal cracking, and containing l-n-olefins as the major type of olefin components, and organic sulfur compounds in concentrations exceeding 0.1% sulfur (1) at first with carbon monoxide and hydrogen at temperatures between about 50 and 250°C and pressures in the range of 50 and 6000 psi in the presence of a Group VIII transition metal carbonyl complex catalyst in effective amounts to produce aldehydes of a semilinear character having an average of less than one alkyl branch per molecule, (2) then with molecular hydrogen at temperatures between 100 and 250°C and pressures between 200 psi and 5000 psi (13.6 and 340 atm) in the presence of a metal or metal sulfide catalyst in effective amounts to produce the corresponding alcohols of a semilinear character having an average of less than one alkyl branch per molecule.
6. A semilinear, isomeric C5 to C21 aldehyde mixture having less than one branch per molecule comprising 15 to 50 weight % of normal aldehyde, 3 to 20% of 3-methyl branched aldehyde, 3 to 20% of 2-methyl branched aldehyde and 3 to 20% of 2-ethyl and higher 2-n-alkyl branched aldehydes.
7. The mixture of Claim 6 , wherein said normal aldehyde and 3-methyl plus 2-methyl branched aldehydes constitute more than 40% of the mixture.
8. The mixture of Claim 6, wherein said semilinear aldehyde mixture contains 5 to 15 carbon atom per molecule and possesses alkyl moieties making the mixture a suitable intermediate for the preparation of ester plasticizers having advantageous low temperature properties.
9. A semilinear isomeric C7 co C21 aldehyde mixture having less than one branch per molecule comprising 15 to 50% of normal aldehyde, 3 to 20% of 3-methyl branched aldehyde, 3 to 20% of 2-methyl branched aldehyde, and 3 to 20% of 2-ethyl and higher n-alkyl branched aldehydes.
10. The mixture of Claim 8, wherein said semilinear aldehyde Mixture contains 10 to 21 carbon atoms per molecule and possesses alkyl moieties making the mixture a suitable intermediate for surfactants having appropriate biodegradability.
11. A semilinear, isomeric C11 aldehyde mixture having less than one branch per molecule and comprising 15 to 50% of normal undecanal, 3 to 20% of 3-methyldecanal and 3 to 20% of 2-methyldecanal, said C11 aldehydes together constituting more than 40% of the total.
12. A semilinear, isomeric C13 aldehyde mixture having less than one branch per molecule comprising 15 to 50% of normal tridecanal, 3 to 20%
of 3-methyldodecanal and 3 to 20% of 2-methyldecanal, said C13 aldehydes together constituting more than 40% of the total.
13. A semilinear isomeric C5 to C21 primary alcohol mixture having less than one branch per molecule comprising 15 to 50% of normal alcohol, 3 to 20% of 3-methyl branched alcohol and 3 to 20% of 2-methyl branched alcohol.
14. The mixture of Claim 13, wherein said normal alcohol and 3-methyl plus 2-methyl branched alcohols constitute more than 40% of the mixture.
15. The mixture of Claim 13, wherein said semilinear alcohol mixture contains 5 to 15 carbon atoms and possesses alkyl moieties making the mixture a suitable reactant for the preparation of ester plasticizers having advantageous low temperature properties.
16. A semilinear, isomeric C7 to C21 alcohol mixture having less than one branch per molecule, comprising 15 to 50% n-alcohol, 3 to 20% of 3-methyl branched alcohol, 3 to 20% of 2-methyl branched alcohol and 3 to 20% of 2-ethyl and higher n-alkyl branched alcohols.
17. The mixture of Claim 16 wherein said semilinear alcohol mixture contains 10 to 21 carbon atoms per molecule and possesses alkyl moieties making the mixture a suitable intermediate for surfactants having appropriate biodegradability.
18. A semilinear isomeric primary C9 alcohol mixture having less than one branch per molecule comprising 15 to 50% of normal nonanol, 3 to 20% 3-methyloctanol and 3 to 20% 2-methyloctanol said C9 alcohols together constituting more than 40% of the total.
19. A semilinear isomeric primary C7 alcohol mixture having less than one branch per molecule comprising 15 to 50% of normal heptanol, 3 to
20% 3-methylhexanol and 3 to 20% 2-methylhexanol, said C7 alcohols together constituting more than 40% of the total.
20. Plasticizer esters consisting of neutral alkyl esters of mono- di- and tribasic carboxylic acids and phosphorus acids wherein said alkyl groups each have less than one alkyl branch and comprise 15 to 50% of normal alkyl, 3 to 20% of 3-methyl branched alkyl and 3 to 20% of 2-methyl branched alkyl groups and together they represent more than 40% of the total.
21. Dialkyl phthalate plasticizer esters of advantageous low temperature properties when employed in plasticized thermioplastic resins C5 to C15 alkyl groups with less than one alkyl branch on the average and comprising 15 to 50% normal alkyl, 3 to 20% 3-methyl branched alkyl, 3 to 20% 2-methyl branched alkyl, said alkyl groups together representing 40% or more of the total.
22. A ditridecyl phthalate of advantageous low temperature properties having tridecyl groups with an average of less than one alkyl branch and comprising 15 to 50% normal tridecyl, 3 to 20% 3-methyldodecyl and 3 to 20% 2-methyldodecyl groups, said tridecyl groups together representing 40% or more of the total.
23. A semilinear isomeric C6 to C12 alkyl acetate ester having alkyl groups with an average of less than one alkyl branch and comprising 15 to SO% normal alkyl, 3 to 20% 3-methyl branched alkyl and 3 to 20%
2-methyl branched alkyl groups said alkyl groups together representing 40%
or more of the total.
24. A semilinear surfactant consisting from the nonionic surfactant group of ethoxylated and/or propoxylated alcohols, the anionic surfactant group of alkylsulfates, ethoxylated and/or propoxylated alkylsulfates or alkanesulfonates; the cationic surfactant groups of alkylamines, ethoxylated and/or propoxylated alkylamines, alkyloxypropylamines, ethoxylated and/or propoxylated alkyloxypropylamines and quaternary ammonium salts of said amines, wherein the isomeric C8 to C21 alkyl groups of said surfactants each have on an average less than one branch and comprise 15 to 50% normal alkyl, 3 to 20% 3-methyl branched alkyl, 3 to 20% 2-methyl branched alkyl and 3 to 20% 2-ethyl and higher n-alkyl groups said alkyl groups together representing more than 50% of the total.
25. A biodegradable semilinear, isomeric ethoxylated C8 to C21 higher alcohol surfactant containing from 1 to 30 ethoxy units wherein the alkyl groups have an average of less than one branch per molecule and comprise 15 to 50% normal alkyl, 3 to 20% 3-methyl branched alkyl, 3 to 20%
2-methyl branched alkyl and 3 to 20% ethyl and higher n-alkyl groups said alkyl groups together representing more than 50% of the total.
26. A semilinear, isomeric ethoxylated tridecyl alcohol containing from 1 to 30 ethoxy units the tridecyl groups have an average of less than one branch per molecule, comprising 15 to 50% normal tridecyl, 3 to 20% methyldodecyl, 3 to 20% 2-methyldodecyl and 3 to 20% 2-ethyl and higher n-alkyl branched tridecyl groups said isomeric tridecyl groups together representing more than 50% of the total.
CA000549993A 1987-10-05 1987-10-22 Process for the hydroformylation of sulfur-containing thermally cracked petroleum residua and novel products thereof Expired - Lifetime CA1294973C (en)

Applications Claiming Priority (2)

Application Number Priority Date Filing Date Title
US10517187A 1987-10-05 1987-10-05
US105,171 1987-10-05

Publications (1)

Publication Number Publication Date
CA1294973C true CA1294973C (en) 1992-01-28

Family

ID=22304438

Family Applications (1)

Application Number Title Priority Date Filing Date
CA000549993A Expired - Lifetime CA1294973C (en) 1987-10-05 1987-10-22 Process for the hydroformylation of sulfur-containing thermally cracked petroleum residua and novel products thereof

Country Status (1)

Country Link
CA (1) CA1294973C (en)

Cited By (1)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
CN111470539A (en) * 2019-10-20 2020-07-31 安徽大学 Preparation method of high-efficiency catalytic benzylamine oxidative coupling catalyst

Cited By (1)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
CN111470539A (en) * 2019-10-20 2020-07-31 安徽大学 Preparation method of high-efficiency catalytic benzylamine oxidative coupling catalyst

Similar Documents

Publication Publication Date Title
US4711968A (en) Process for the hydrofomylation of sulfur-containing thermally cracked petroleum residua
US5072057A (en) Process for the hydroformylation of sulfur-containing thermally cracked petroleum residue and novel products thereof
US5030774A (en) Process for the hydroformylation of sulfur-containing thermally cracked petroleum residue and novel products thereof
KR830001167B1 (en) Hydroformylation Method for Making n-Valer Aldehyde
US6049011A (en) Hydroformylation process
US2595096A (en) Synthesis of alcohol from olefins, carbon monoxide, and hydrogen
CA1258873A (en) Process for the production of 2-ethyl-hexanol
US7582802B2 (en) Propylene hydroformylation
CN102056872B (en) Method for separating 1-butene from C4-containing hydrocarbon streams by hydroformylation
US5463147A (en) Decyl alcohol mixtures, phthalic esters obtainable therefrom and their use as plasticizers
CA1273367A (en) Hydroformylation of olefins
NO311425B1 (en) Process for producing oxygenated products from an olefinic feed
EP0804398B1 (en) Hydroformylation process
US2671119A (en) Hydrogenation of oxo alcohol bottoms
US4922028A (en) Process for the hydroformylation of sulfur-containing thermally cracked petroleum residue and novel products thereof
US6914162B2 (en) Method for producing hydroformylation products of olefins with 2 to 8 carbon atoms
CA2131337A1 (en) Process for the hydroformylation of 1,3-butadiene
CA1294973C (en) Process for the hydroformylation of sulfur-containing thermally cracked petroleum residua and novel products thereof
EP0185477B1 (en) Production of ether/ether-alcohol compositions
US4723047A (en) Process for the preparation of nonadecanediols
US2686206A (en) Oxo process
EP0959063A2 (en) Process for producing alcohols
CA2138222C (en) Process for preparing aldehydes substituted in the .alpha. position by an alkyl radical
US2623074A (en) Production of alcohols and gasoline by the oxo process
EP0080449A1 (en) Synthesis of isoprene from linear butenes

Legal Events

Date Code Title Description
MKEX Expiry