8e0f.8810

Upload: candhare

Post on 07-Apr-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/6/2019 8e0f.8810

    1/13

    Applied Catalysis A: General 221 (2001) 253265

    Recent advances in processes and catalystsfor the production of acetic acid

    Noriyuki Yoneda a, Satoru Kusano a,, Makoto Yasui b, Peter Pujado c, Steve Wilcher c

    a Chiyoda Corporation, 3-13 Moriya-cho, Kanagawa-ku, Yokohama 221-0022, Japanb Chiyoda Corporation, 2-12-1 Tsurumichuo, Tsurumi-ku, Yokohama 230-8601, Japan

    c UOP LLC, 25 East Algonquin Road, Des Plaines, IL 60017-5017, USA

    Abstract

    Novel acetic acid processes and catalysts have been introduced, commercialized, and improved continuously since the

    1950s. The objective of the development of new acetic acid processes has been to reduce raw material consumption, energy

    requirements, and investment costs. At present, industrial processes for the production of acetic acid are dominated by

    methanol carbonylation and the oxidation of hydrocarbons such as acetaldehyde, ethylene, n-butane, and naphtha. This paper

    discusses advances in acetic acid processes and catalysts according to the following routes: (1) methanol carbonylation; (2)

    methyl formate isomerization; (3) synthesis gas to acetic acid; (4) vapor phase oxidation of ethylene, and (5) other novel

    technologies. 2001 Published by Elsevier Science B.V.

    Keywords: Acetic acid; Methanol carbonylation; Hydrocarbon oxidation; Reaction mechanisms

    1. Introduction

    Acetic acid is an important commodity chemical

    used in a broad range of applications. As shown in

    Fig. 1, acetic acid is used primarily as a raw ma-

    terial for vinyl acetate monomer (VAM) and acetic

    anhydride synthesis, and as a solvent for purified

    terephthalic acid (PTA) production. The demand for

    acetic acid has increased, especially in southeast Asia,

    where several new PTA plants have been built. Withthe increased demand and installed capacity for PTA

    in southeast Asia, the region has become a major

    producer of polyester (PET) fiber, film, and resin.

    Although the economic crisis in Asia momentarily

    suppressed the demand for acetic acid to less than

    expected levels, in the medium and long terms there

    Corresponding author. Tel.: +81-45-441-9151;

    fax: +81-45-441-1281.

    is potentially a great demand for acetic acid in this

    market.

    The total world capacity of acetic acid has reached

    approximately 7.8 million t in 1998 with BP-Amoco

    and Celanese accounting for more than 50% of

    the worlds capacity [1]. BP-Amoco and Celanese

    have installed capacities of 1.5 million t (19%), and

    2.0 million t (26%), respectively.

    2. Processing routes to acetic acid

    Originally, acetic acid was produced by aerobic fer-

    mentation of ethanol, which is still the major process

    for the production of vinegar. The first major com-

    mercial process for the synthetic production of acetic

    acid was based on the oxidation of acetaldehyde. In

    an early process for the conversion of acetylene to ac-

    etaldehyde introduced in 1916 in Germany and used

    0926-860X/01/$ see front matter 2001 Published by Elsevier Science B.V.

    P I I : S 0 9 2 6 - 8 6 0 X ( 0 1 ) 0 0 8 0 0 - 6

  • 8/6/2019 8e0f.8810

    2/13

    254 N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253265

    Fig. 1. Use of acetic acid.

    in China until recently, an organo-mercury compound

    was used as the catalyst. The toxicity of the mercurycatalyst resulted in significant environmental pollu-

    tion, and as a result, has essentially been phased out.

    As the petrochemical industry developed in the 1950s,

    the raw material for the production of acetaldehyde

    shifted to ethylene. Other processes for the production

    of acetic acid introduced in the 1950s and 1960s were

    based on the oxidation of n-butane or naphtha. The

    major producers of acetic acid via direct oxidation of

    hydrocarbons were Celanese (via n-butane) and BP

    (via naphtha). However, these reactions also produce

    significant amounts of oxidation by-products, as sum-marized in Table 1, and their separation and recovery

    can be very complex and expensive.

    The homogeneous methanol carbonylation route to

    acetic acid that used a homogeneous Ni catalyst was

    first commercialized by BASF in 1955. An improved

    process was later disclosed by BASF in 1960. The

    process used an iodide-promoted CO catalyst and

    operated at elevated temperature (230 C) and pres-

    sure (600 atm). The product yields exhibited by this

    Table 1

    Acetic acid process

    Catalyst Reaction condition

    (C, atm)

    Yield By-product

    Methanol carbonylation Rhodium complex 180220, 3040 MeOH: 99%, CO: 85% None

    Acetaldehyde oxidation Manganese acetate

    or cobalt acetate

    5060, atmospheric

    pressure

    CH3CHO: 95% None

    Ethylene direct oxidation Palladium/heteropolyacid/

    metal

    150160, 80 Ethylene: 87% Acetaldehyde CO2

    Hydrocarbon oxidation

    (n-butane, naphtha)

    Cobalt acetate or

    manganese acetate

    150230, 5060 n-Butane: 50%,

    naphtha: 40%

    Formic acid,

    propionic acid, etc.

    process were 90, and 70% based on methanol and CO

    consumption, respectively, [2]. In 1970, Monsanto

    commercialized an improved homogeneous methanol

    carbonylation process using a methyl-iodide-promotedRh catalyst [36]. Compared to other acetic acid syn-

    thesis routes (ethanol fermentation, and acetaldehyde,n-butane, or naphtha oxidation), homogeneous Rh

    catalyzed methanol carbonylation is an efficient route

    that exhibits high productivity and yields. The pro-

    cess operated at much milder conditions (180220 C,

    3040 atm) than the BASF process and exhibited su-

    perior performance: acetic acid yields were 99 and

    85% based on methanol and CO consumption, re-

    spectively, [7]. Celanese and Daicel further improved

    the Monsanto process during the 1980s by adding

    a lithium or sodium iodide promoter to enable the

    operation in a reduced water environment [815]. At

    lower water concentrations, by-product formation via

    the water gas shift reaction is reduced, thus improving

    raw materials consumption and reducing downstream

    separation costs.

    Homogeneous metal catalysts less costly than Rh

    (for example, Ni [16,17,75,76] and Ir [3,1824] with

    other metal additives) have also been investigated.

    The Ir-based process allows operation at reactor water

    levels comparable to those of the improved Celanese

    process and was commercialized by BP Chemicals in1996.

    Until recently, virtually all new acetic acid ca-

    pacity has made use of the homogeneous methanol

    carbonylation technology developed by Monsanto

    and practiced commercially by all major acetic acid

    manufacturers, including BP-Amoco, Celanese, and

    others. As a result, more than 60% of the world acetic

    acid production employs the methanol carbonylation

    methods, as shown in Fig. 2.

  • 8/6/2019 8e0f.8810

    3/13

    N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253265 255

    Fig. 2. Acetic acid process routes.

    Inherent to the homogeneous system, however, are

    drawbacks relating to catalyst solubility limitations

    and the loss of expensive Rh metal due to precipi-

    tation in the separation sections. Accordingly, immo-

    bilization of the Rh complex on a support has been

    the subject of considerable investigation. Chiyoda and

    UOP have jointly developed an improved methanol

    carbonylation process for the production of acetic acid

    based using a heterogeneous Rh catalyst system [25].

    A direct oxidation process for the production of

    acetic acid starting from ethylene was commercialized

    by Denko in 1997. While the raw material, ethylene,

    is more expensive than in the methanol carbonylationroute, the investment cost is reported to be lower and

    competitive for small or medium-size capacity plants.

    Wacker-Chemie plans to commercialize a new acetic

    Fig. 3. Catalytic cycle for rhodium carbonylation.

    acid process based on butylene feedstock. This process

    also employs direct oxidation. Its key features are the

    use of a relatively cheap raffinate-2 feedstock and com-

    petitive economics in medium size plants. Recently,Poulenc and others have disclosed the direct produc-

    tion of acetic acid from ethane; there are no indica-

    tions of impending commercialization for this route.

    Generally, the production cost of commodity chem-

    icals such as acetic acid is dominated by the raw mate-

    rial costs, and methanol carbonylation is still regarded

    as the preferred route to produce acetic acid. Table 1

    summarizes reaction conditions, catalysts and yields

    for the major processes used to produce acetic acid. A

    number of reviews on production of acetic acid have

    been published and are referred [7,29,73,74].

    3. Methanol carbonylation

    3.1. Rhodium catalyzed methanol carbonylation

    The methanol carbonylation process, Mon-

    santo process, is operated under mild conditions

    (180220 C, 3040 atm) and exhibits high selectivity

    to acetic acid based on methanol (99%) and carbon

    monoxide (85%) [7]. While the reaction, as shown

    below, can be carried out in a variety of rhodium(I) or rhodium (III) complexes [6,18], under reaction

    conditions they are almost invariably converted to the

    active catalyst [RhI2(CO)2]1. As shown in Fig. 3,

  • 8/6/2019 8e0f.8810

    4/13

    256 N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253265

    methyl iodide is provided by the reaction of feed

    methanol with hydrogen iodide:

    CH3OH+ CO

    Rh complex

    CH3COOHMethyl iodide is oxidatively added to the rhodium

    dicarbonyldiiodide complex [RhI2(CO)2]1 (A)

    to generate a rhodiummethyl complex (B). This

    rhodiummethyl complex can rapidly undergo a

    methyl migration to a neighboring carbonyl group in

    the acetyl form (CH3CO) and react with CO (C) to

    generate the rhodiumacetyl complex (D). Reductive

    elimination of acetyl iodide (CH3COI) can then lib-

    erate the original rhodium complex (A). Hydration of

    acetyl iodide is very rapid in the presence of excess

    water and will result in the formation of acetic acidand hydrogen iodide to complete the cycle.

    The reaction rate is independent of methanol

    concentration and carbon monoxide pressure. The

    rate-determining step is believed to be the oxidative

    addition of methyl iodide to the rhodium center of the

    rhodium complex (A), and the reaction rate is essen-

    tially of first order in both catalyst and methyl iodide

    concentrations under normal reaction conditions:

    reaction rate [catalyst][CH3I]

    A substantial quantity of water (1415 wt.%) is re-quired to achieve high catalyst activity and also to

    maintain good catalyst stability [8,9,1214]. However,

    as rhodium also catalyzes the water gas shift reaction

    (Fig. 4), the side reaction forming CO2 and H2 from

    CO is significantly affected by water and hydrogen

    iodide concentration in the reaction liquid [26,27].

    Propionic acid is observed as the major liquid

    by-product in this process. This is produced by the

    carbonylation of ethanol that is often present as a

    minor impurity in the methanol feed; however, other

    Fig. 4. Mechanism for water gas shift.

    routes are active since more propionic acid is observed

    than can be accounted for by only this mechanism.

    The rhodium catalyst system can generate acetalde-

    hyde, and it is proposed that this acetaldehyde isreduced by hydrogen in the system to give ethanol

    which subsequently yields propionic acid. One possi-

    ble precursor for the generation of acetaldehyde is the

    rhodiumacetyl species, as shown in the following

    mechanism [28]:

    [RhI3(CO)(COCH3)]+HI

    [RhI4(CO)]+ CH3CHO

    [RhI4(CO)] RhI3 + I

    + CO

    Reaction of this species with hydrogen iodide would

    yield acetaldehyde and [RhI4(CO)]1. The latter

    species is well known in this system and is postulated

    as the principal cause of catalyst loss by precipitation

    of inactive rhodium tri-iodide [28].

    Acetaldehyde undergoes self-condensation or aldol

    condensation and yields butenal and higher aldehy-

    des. These can undergo further reactions to alcohols

    and carboxylic acids as summarized in the network of

    Fig. 5 [28]. It would be expected that the homologa-

    tion observed would result in unsaturates and iodides

    having an even number of carbon atoms, and long

    chain carboxylic acids with an odd number of carbonatoms. Particular problems are encountered with the

    C6 species present. The boiling points of the unsatu-

    rated compounds, including hexanal and some of its

    isomers, are very similar to that of acetic acid. Further-

    more, hexyl iodide is observed to form a constant boil-

    ing azeotropic mixture with acetic acid. The presence

    of the unsaturates, even at low parts per million con-

    centrations can cause problems with product stability.

    Separation of pure acetic acid product from the re-

    action medium presents few problems. In this process,

    however, the expensive Rh metal can be lost due to itsprecipitation and vaporization in the flash column. A

    schematic of a conventional methanol carbonylation

    plant configuration is shown in Fig. 6 [28]. Rhodium

    catalyst is separated from the product acetic acid by

    conducting a simple flash; the catalyst remains in the

    liquor and can be recycled to the reactor. The sepa-

    ration of light compounds, such as methyl iodide and

    methyl acetate, may be carried out in the first distil-

    lation column. This column is followed by a drying

    column and then a column for the removal of heavy

  • 8/6/2019 8e0f.8810

    5/13

    N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253265 257

    Fig. 5. Network of liquid by-products.

    by-products. Energy usage in this fractionation train

    can be high, depending on the concentration of waterand impurities such as propionic acid, heavy unsatu-

    rates, and hexyl iodide present.

    In the Monsanto process, because of the high water

    concentration in the reactor (1415 wt.%), the separa-

    tion of water from the acetic acid product is a major

    energy consumer and can limit the unit capacity. In

    addition, excess water causes carbon monoxide yield

    loss due to the water gas shift reaction, and increases

    the formation of by-products such as propionic acid,

    thus lowering the acetic acid quality. Considerable sav-

    ings in operating costs can be realized by operatingat low water concentration if a way can be found to

    compensate for the consequent decrease in the reac-

    tion rate and catalyst stability [29]. As a result, the

    rhodium complex stability at low water concentrations

    has been extensively investigated.

    Fig. 6. Schematic of a acetic acid plant configuration.

    Group I metal iodides, especially lithium iodide in

    combination with methyl iodide, were identified earlyas a good agent for enhancing the stability of the

    rhodium catalyst at low reactor water concentrations

    (45 wt.%), and also for decreasing liquid by-product

    formation [1214]. Further work in this area revealed

    that the addition of a substantial quantity (1620 wt.%)

    of group I metal iodides also enhanced the reactor

    productivity even at quite low water concentrations

    (2 wt.%) [811]. These features reportedly allow exist-

    ing plants to expand their capacity for little incremen-

    tal capital cost. The improved methanol carbonylation

    process, low water process, effected by addinggroup I metal iodides to the Monsanto process was first

    commercialized in the 1980s by Celanese and Daicel.

    In this process, it is proposed that the addition of

    a significant quantity of group I metal causes the Rh

    complex to be more coordinated by CH3COO and

  • 8/6/2019 8e0f.8810

    6/13

    258 N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253265

    Fig. 7. Reaction acceleration mechanism by iodide salt.

    increases the rate of oxidative insertion of methyl io-

    dide (the rate determining step), thus promoting theprimary carbonylation reaction (Fig. 7). As Figs. 7 and

    8 shows the effect of the addition of lithium iodide

    on the reaction rate, the overall carbonylation rate in-

    crease is presumably due in part to the formation of

    a strong nucleophilic five-coordinate dianionic inter-

    mediate [Rh(CO2)2I2L]2 (L = I, OAc) which

    is more active toward oxidative addition of methyl io-

    dide [811,29].

    The main advantages of the low water process rela-

    tive to the conventional Monsanto process are reduced

    raw materials consumption, increased productivity,

    lower utility requirements, and lower capital costs

    per unit of product. However, low water operation

    Fig. 8. Effect of Li salt addition. Reaction condition:

    [CH3] = 1.0 M, [MeOAc] = 0.3M, [H2O] = 1.0 M, temperature

    = 190 C, total pressure = 400 psig.

    with alkali-iodide promoters results in a higher iodide

    environment, and higher residual iodide in the final

    product. High iodide concentration in acetic acid leads

    to catalyst poisoning problems in some downstreamapplications, such as in the manufacture of VAM. To

    overcome the problems associated with high iodide

    concentration in the final product, treatment by active

    carbon [30], hydrogenation [31,32], and extra distilla-

    tion [33,34] have been proposed. Celanese disclosed

    the silver-guard process for the removal of very low

    levels of iodide impurities from acetic acid in their

    patent [35]. The use of silver metal on an ion exchange

    resin such as Amberlyst-15 reduces the iodide level

    to below 1 ppb, as opposed to 20 ppb more normally

    achieved by conventional methods. One particular

    advantage of this system is the ability to effectively

    remove the halide impurity in a single step, thus avoid-

    ing the need for additional distillation and recovery.

    3.2. Nickel catalyzed methanol carbonylation

    Recent studies have shown that nickel catalysts

    can operate under mild conditions (190 C, 70 atm)

    with the addition of methyl iodide as a co-promoter

    [16]. The activity of nickel catalyst systems can be

    increased and the volatility of nickel carbonyl com-

    pounds lowered by the introduction of stabilizers suchas phosphines, alkali metals, tin, and molybdenum

    [16,17,25,75,76]. The active catalysts are thought to

    be Ni(0) complexes. For phosphine-promoted cata-

    lyst, Ni(PR3)2 is considered an active form of catalyst

    and, in addition, Ni(CO)4 was observed in all cases,

    and its concentration was reduced by strongly coordi-

    nating ligands and enhanced by weakly coordinating

    ligands [76]. Recent work on nickel catalyst systems

    shows that reaction rates and selectivities can ap-

    proach those achieved in the rhodium catalyst system.

    Although nickel catalysts have the advantage of beingmuch cheaper than rhodium, and are easy to stabi-

    lize at low reactor water concentrations, [Ni(CO)4] is

    known to be a very toxic and volatile compound. To

    date, commercialization has not proceeded.

    3.3. Iridium catalyzed methanol carbonylation

    The potential use of iridium instead of rhodium was

    identified as part of the early work done by Monsanto

    [3,18,25], however, the reaction rate exhibited by the

  • 8/6/2019 8e0f.8810

    7/13

    N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253265 259

    rhodium catalyst system was superior to that of irid-

    ium. Recently, it was disclosed that an improved irid-

    ium catalyst, in combination with a promoter metal

    such as ruthenium, has higher activity and results inlower product impurity levels than reported in pre-

    vious iridium systems [19]. The production of acetic

    acid using the iridium catalyst system has been com-

    mercialized by BP-Amoco in two world scale plants to

    date, and has received wide publicity as the Cativa

    process. Although much iridium is required to achieve

    an activity comparable to the rhodium catalyst-based

    processes, the catalyst system is able to operate at re-

    duced water levels (less than 8 wt.% for the Cativa

    process versus 1415 wt.% for the conventional Mon-

    santo process). Thus, lower by-product formation and

    improved carbon monoxide efficiency are achieved,

    and steam consumption is decreased. Until the early

    1990s, the difference in the prices of rhodium (US$

    500/oz) and iridium (US$ 60/oz) was the driving force

    for replacing rhodium with iridium. However, current

    price increases for iridium (US$ 450/oz) negate the

    advantage in catalyst price.

    The unique differences between the rhodium cat-

    alytic cycle and that of iridium in methanol car-

    bonylation have been investigated [36]. The anionic

    iridium cycle shown in Fig. 9, is similar to that shown

    earlier for rhodium. Model studies have demonstrated

    Fig. 9. Catalytic cycle for iridium carbonylation.

    that the oxidative addition of methyl iodide to the

    iridium center is of the order of 150 times faster

    than the equivalent reaction with rhodium [36]. This

    represents a possible improvement in the availablereaction rates, as methyl iodide addition is not the

    rate determining step. The slowest step in this cycle is

    the insertion of carbon monoxide to form the iridium

    acetyl species, that involves the elimination of ionic

    iodides and the coordination of an additional carbon

    monoxide ligand. This would suggest the following

    expression. The dependence on ionic iodide:

    reaction rate [catalyst][CO]

    [I]

    suggests that high reaction rates should be achiev-

    able by operating at low iodide concentrations. It

    also suggests that the inclusion of species capable of

    assisting in the abstraction of iodide should promote

    the rate-limiting step. The patent would suggest that

    ruthenium, or rhenium are the preferred promoters

    [20,21]. In effect, a proprietary blend of promoters

    has been found to increase reaction rate. The above

    expression does not imply any effect from the water

    present in the matrix, but water is found to have a

    significant effect on rate [22].

    In the improved iridium system, low water concen-

    tration in the reactor results in the formation of fewer

  • 8/6/2019 8e0f.8810

    8/13

    260 N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253265

    by-products such as propionic acid than in the original

    Monsanto rhodium system, and no addition of lithium

    iodide is required. Consequently, the iridium catalyst

    system is also characterized by the formation of fewerhigher alkyl-iodide species than in the conventional

    low water process.

    3.4. Heterogeneous rhodium catalyzed

    methanol carbonylation

    In order to overcome the limitations of the ho-

    mogeneous catalyst system (e.g. Rh precipitation

    and catalyst solubility limitations), the immobiliza-

    tion of the Rh complex on a support has been the

    subject of considerable investigation. Active carbon

    was investigated as a possible support and proposed

    for vapor-phase operation [7,37,38]. However, the

    reaction rate was 1/10001/10 that of Monsantos ho-

    mogeneous process and selectivity was also poorer.

    Inorganic oxides and zeolites were also investigated

    for use in vapor-phase operation [39,40]. For exam-

    ple, attaching the Rhphosphine ligand complex to

    alumina by silylation was attempted [41,42]. The

    resultant reaction rates for these catalysts were also

    found to be poor relative to those observed for the

    homogeneous system. To increase catalyst activity

    for operation in the liquid phase, ion exchange resinsbased on cross-linked polystyrene and incorporating

    pendant phosphines, or vinyl pyridine copolymers

    have been evaluated [4345]. Although the activity of

    these catalysts in the liquid phase was comparable to

    Monsantos homogeneous catalyst, there were prob-

    lems with rhodium metal leaching from the resins

    and the decomposition of the resins during opera-

    tion at elevated temperature. Vinyl pyridine resin was

    known to be more robust and more tolerant of oper-

    ation at elevated temperature relative to polystyrene

    resins. It was disclosed that catalysts using pyridineresins exhibited high tolerance to operation at ele-

    vated temperature and pressure, and higher reaction

    rate than Monsantos rhodium system [46]. Further-

    more, Chiyoda introduced novel pyridine resins and

    catalysts that exhibited high activity, long catalyst

    life, and no significant rhodium loss [4749]. Based

    on this heterogeneous Rh catalyst, Chiyoda and

    UOP have jointly developed an improved methanol

    carbonylation process, called the acetica process,

    for the production of acetic acid. Until the recent

    development of a commercial heterogeneous Rh cata-

    lyst system by Chiyoda, no successful demonstration

    of such a catalyst had been known [7].

    The heterogeneous catalyst commercialized for theacetica process consists of Rh complexed on a novel

    poly-vinyl pyridine resin [50], which is tolerant of

    elevated temperatures and pressures. Under reaction

    conditions, the Rh is converted to its catalytically

    active anion form [Rh(CO)2I2]1. Furthermore, the

    nitrogen atoms of the resin pyridine groups become

    positively charged after quaternization with methyl

    iodide. Thus, the strong ionic association between the

    pyridine nitrogen groups and the Rh complex causes

    the immobilization (Fig. 10). The concentration of Rh

    on the solid phase is determined by the ion exchange

    equilibrium. Because equilibrium strongly favors the

    solid phase, virtually all the Rh in the reaction mixture

    is immobilized.

    In the acetica process, the methanol carbonyla-

    tion reaction is conducted at moderate temperature

    (160200 C) and pressure (3060 atm) and at low

    water concentration without any additives present.

    Catalyst stability has been demonstrated in both once-

    through and continuous-recycle pilot plant testing at

    process conditions, low water content, and no Rh or

    resin makeup. The catalyst exhibited no deactivation

    after continuous operation for more than 7000 h [50].With homogeneous methanol carbonylation routes,

    acetic acid productivity is directly proportional to

    catalyst concentration in the reaction liquid, and as

    a result, acetic acid production is restricted by the

    solubility of the active metal. Limited success has

    been achieved in improving catalyst solubility in

    these systems by increasing the reaction-mixture wa-

    ter concentration or by adding iodide salt stabilizers

    [8,9,1214]. Both additives, however, result in in-

    creased recycle and separation costs, higher corrosion

    rates, and difficulty in product purification.With the heterogeneous catalyst system, catalyst

    solubility limitations no longer govern reactor capac-

    ity since catalyst concentrations several times greater

    than those achievable in the homogeneous systems

    are possible. Immobilization also significantly re-

    duces the loss of expensive Rh metal because the

    catalyst is confined to the reactor rather than circulat-

    ing downstream, where reduced pressures may cause

    precipitation of rhodium and vaporization losses of

    metal carbonyl compounds. The lower water content

  • 8/6/2019 8e0f.8810

    9/13

    N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253265 261

    Fig. 10. Rhodium immobilization.

    of 37 wt.% typical of the acetica process results

    in reduced production of CO2, and hydrogenated

    by-products via the water gas shift reaction. Also,

    because of the lower water content, less hydrogen

    iodide is present in the system, and consequently the

    process environment is less corrosive.

    While the continuously stirred tank reactors (CSTR)

    used in the conventional homogeneous processes

    can be limited by gas solution rates to liquid and

    are often prone to mechanical problems, the bubble

    column, or gas lift reactor employed with the hetero-

    geneous catalyst process does not suffer from such

    problems and limitations. The acetica three-phase

    gas lift reactor has no moving parts or mechanical

    Fig. 11. Bubble column reactor and acetica process flow.

    seals and was designed to maximize the performance

    of its unique heterogeneous catalyst system without

    any rotating equipment (Fig. 11). Methanol and CO

    feeds are introduced at the reactor bottom, where the

    compressed CO gas is distributed through a sparger.

    Both of these feeds, along with the recycle liquid

    and catalyst, flow up the reactor riser, where the

    CO is consumed in the reaction. The process flow,

    which is similar to that of a conventional homoge-

    neous process is shown in Fig. 11. In cases where the

    acetic acid product will be used for VAM production,

    novel iodide removal technology is available to re-

    duce the iodide in the acetic acid product to less than

    3 ppb [51].

  • 8/6/2019 8e0f.8810

    10/13

    262 N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253265

    4. Methyl formate isomerization

    It has been proposed that acetic acid can be pro-

    duced by isomerization of methyl formate in the pres-ence of a homogeneous rhodium catalyst together with

    other metal additives [52,53]. Heterogeneous rhodium

    catalysts supported on poly-vinyl pyridine resin have

    also been proposed for this application [54]. This cata-

    lyst has the same chemical morphology as a methanol

    carbonylation catalyst. Methyl formate is produced by

    dehydrogenation of methanol [55] or by methanol car-

    bonylation under high pressure in the presence of:

    HCOOCH3 CH3COOH

    copper oxide and alkali catalyst. It is noted thatacetic acid production via methanol dehydrogena-

    tion followed by methyl formate isomerization re-

    quires only methanol and no carbon monoxide

    plant:CH3OHHCHOHCOOCH3CH3COOH-

    Acetic acid can be produced from only methanol us-

    ing a Ru-Sn catalyst according to the following steps

    [56,57]. Ru-Sn bimetallic complexes are proposed to

    be the active species.

    5. Synthesis gas route to acetic acid

    A nearby synthesis gas plant to produce CO is nor-

    mally required to provide feed to an acetic acid plant.

    On the contrary, an efficient integrated synthesis

    gas and methanol synthesis plant and acetic acid plant

    are available by combination of current technology at

    the natural gas source. This integrated process could

    achieve a significant capital cost reduction relative to

    the conventional flow scheme.

    Applying this concept, Haldor Topsoe proposed

    an integrated process that includes the synthesis ofmethanol and dimethyl ether (DME) in a first catalytic

    reaction stage and the subsequent carbonylation of

    methanol and DME into acetic acid [58,59]. Although

    the reaction pressure required for methanol synthesis

    is higher than the pressure used in acetic acid syn-

    thesis, the combination of methanol synthesis with

    dimethyl ether synthesis can reduce the pressure of

    the first reaction step. The catalyst consists of a mix-

    ture of the catalyst for methanol synthesis (Cu-Zn-Al

    oxide, etc.) and a dehydration catalyst (H-ZSM-5,

    etc.). The reaction is carried out at approximately

    220 C and 40 atm:

    CO+ 2H2 CH3OH

    2CH3OH CH3OCH3 +H2O

    H2O+ CO CO2 +H2

    In the acetic acid synthesis step, carbonylation of

    DME and methanol to acetic acid is carried out by

    the rhodium carbonyl complex catalyst with carbon

    monoxide being supplied from the synthesis gas

    process unit:

    CH3OH+ CO CH3COOH

    CH3OCH3 + 2CO+H2O 2CH3COOH

    Carbonylation reaction conditions of 170250 C and

    2550 atm, can be used to obtain acceptable reaction

    rates in the liquid phase.

    6. Vapor phase oxidation of ethylene

    The two-step oxidation process for the production

    of acetic acid, starting from ethylene through acetalde-

    hyde, was first commercialized in 1960:

    CH2=CH2 +12 O2 CH3CHO

    CH3CHO+12 O2 CH3COOH

    This route involves the liquid phase oxidation of ac-

    etaldehyde using air and typically a manganese ac-

    etate catalyst operating at 5060 C. The reaction is

    based on a free radical mechanism. Although this pro-

    cess features high yield (approximately 90%) and a

    relatively low capital investment cost, it suffers from

    high acetaldehyde feedstock cost and a very corrosive

    catalyst system. Many plants utilizing this technology

    have been shut down over the last 20 years.

    There is also an older process that entails liquid

    phase free radical oxidation of n-butane or naphtha

    in the C4C8 range. These reactions produce a wide

    spectrum of oxidation by-products such as formic acid

    and propionic acid:

    CH3CH2CH2CH3+O2CH3COOH+ by-products

    The direct production of acetic acid from ethylene via

    an acetaldehyde intermediate is a desirable synthesis

  • 8/6/2019 8e0f.8810

    11/13

    N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253265 263

    Scheme 1. Hydration route.

    route that has yet to be developed. Much work has been

    undertaken to develop a simpler, single stage process

    for producing acetic acid directly from ethylene:

    CH2=CH2 +O2catalyst CH3COOH

    Various groups have carried out extensive research and

    development in the area of direct vapor phase oxida-tion of ethylene to acetic acid. Catalyst systems con-

    sisting of palladium chloride and V2O5 supported on

    Al2O3 [60], and combinations of Pd (2%) and H3PO4(25%) on SiO2, Pd-V2O5-Sb2O3 on Al2O3 [61], or

    Pd (1%) on V2O5 [62] have been proposed. These

    catalysts have acetic acid selectivities in the range of

    6090% based on ethylene. The routes have been pro-

    posed according to the different catalyst systems in

    Schemes 1 and 2.

    Denko has developed a direct oxidation process for

    the production of acetic acid based on the hydration

    route [6365] and has commercialized this technol-

    ogy in late 1997. The catalyst consists of either two

    or three components. The first component is palla-

    dium supported on a carrier, preferably in the range of

    0.12% range. The second component is a heteropoly

    acid and their salts, preferably phosphotungstic acid

    salts of lithium, sodium, and copper. The third com-

    ponent is copper, silver, tin, lead, antimony, bismuth,

    selenium, or tellurium.

    The reaction takes place in a fixed bed reactor at

    operating temperatures and pressures of 150160 C,

    and up to 8 atm, respectively. The gases fed to the re-actor are ethylene, oxygen, steam, and nitrogen that is

    used as a diluent. The presence of steam is required to

    Scheme 2. Partial oxidation route.

    enhance the activity and selectivity of the process for

    the production of acetic acid. The selectivity to acetic

    acid is approximately 86%, since it appears that ac-

    etaldehyde and carbon dioxide are necessarily formedin this type of process.

    7. Other proposed technologies for the

    production of acetic acid

    7.1. Ethane oxidation

    In the 1980s, an acetic acid route from ethane was

    introduced. Two reaction mechanisms based on:

    CH3CH3 +O2catalyst CH3COOH+ by-product

    different catalyst systems were proposed: (1) partial

    oxidation of the methyl group, and (2) ethane oxi-

    dation to ethylene followed by ethylene hydration to

    ethanol, or ethylene to acetaldehyde.

    A patent refers to the production of acetic acid

    by reacting ethane, ethylene, or mixtures of ethane

    and ethylene with oxygen over a catalyst containing

    molybdenum, vanadium, and one other metal (Z) in

    the general formula MoxVyZz [66]. In one example,

    the patent describes the gas phase oxidation of a 1/10mixture of ethane and ethylene at 255 C over a vana-

    dium catalyst containing lesser amounts of molybde-

    num, niobium, antimony, and calcium supported on

    an LZ-105 molecular sieve to yield 63% selectivity

    to acetic acid, and 14% selectivity to ethylene at 3%

    ethane conversion. In the combined ethane/ethylene

    feed case, the hydration catalyst further catalyzes the

    hydration of ethylene to ethanol, which is then con-

    verted to acetic acid (Scheme 1). The oxidation cat-

    alyst catalyzes the reaction of ethylene to acetic acid

    and other oxidation products that are converted toacetic acid (Scheme 2).

    In another catalyst system, rhenium or a combi-

    nation of rhenium and tungsten are introduced to

    replace the molybdenum in the dehydrogenation cat-

    alyst [67]. Tests showed that complete substitution

    of molybdenum by rhenium (RexVyZz) is beneficial

    in the reaction of ethane to ethylene, whereas partial

    substitution can increase the selectivity to acetic acid.

    Tests were not performed on ethylene feed, but tests

    on ethane (21% ethane, 3.8% oxygen, and 75.2%

  • 8/6/2019 8e0f.8810

    12/13

    264 N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253265

    nitrogen) resulted in acetic acid selectivity as high as

    78% at an ethane conversion of 14.3%.

    More recently in 1998, another oxidative process

    and catalyst for the production of acetic acid fromethane, or ethylene was disclosed [68,69]. A new

    molybdenum vanadate catalyst system promoted with

    Nb, Sb, Ca, and Pd allows the gas phase oxidation of

    ethane and/or ethylene to acetic acid, with high yield

    and higher selectivity under milder operating condi-

    tions than previously achieved. The patent discloses

    the production of acetic acid with 86% selectivity and

    11% ethane conversion per pass, at a temperature and

    pressure of 250280 C and 15 atm, respectively.

    In 1999, a catalyst for the co-production of ethylene

    and acetic acid from ethane was disclosed [70]. It con-

    sists of phosphorus-modified molybdenum-niobium

    vanadate of formula Mo2.5V1.0Nb0.32Px in which the

    optimum range for the phosphorus (x) is 0.010.06:

    CH3CH3 +O2catalyst CH2==CH2 + CH3COOH+H2O

    Ethane and air (15:85 (v/v)) at 260 C and 200 psig

    (1100/h GHSV) reacted over the above catalyst system

    (x = 0.042) to produce acetic acid and ethylene with

    selectivities of 49.9, and 10.5%, respectively, at 53.3%

    conversion. At phosphorus levels greater than 0.06%,

    there is a marked increase in ethylene production with

    a corresponding decline in acetic acid.

    Recently, many attempts have been disclosed re-

    garding the use of ethane as feedstock. Although

    ethane is a relatively inexpensive and attractive raw

    material for producing acetic acid, the oxidation pro-

    cesses produce a variety of co-products, the disposi-

    tion of which needs to be considered in any business

    plan.

    7.2. Methane carbonylation

    Novel methods for producing acetic acid directly

    from methane under relatively mild conditions have

    been reported. It was first disclosed that acetic acid

    can be produced from methane and carbon monox-

    ide in the presence of: Pd(OCOCH3)2/Cu(OCOCH3)2/

    K2S2O8/CF3COOH [71].

    Secondly, it was reported that the mixture of

    methane, carbon monoxide and oxygen formed acetic

    acid in the presence of rhodium trichloride dissolved

    in water [72]:

    CH4 + CO+1

    2 O2

    RhCl3

    CH3COOH

    This reaction proceeds in an aqueous medium at a

    temperature of approximately 100 C and gives a high

    yield of acetic acid. The reaction rates are reported

    to be too slow for an economically viable industrial

    process, but this novel process route has the potential

    to reduce the cost of acetic acid production.

    8. Conclusions

    Acetic acid represents a commodity chemicalgrowing at 3.54.5% per year from a significant

    and large base capacity. Significant developments in

    both process and catalyst technology have supported

    the growth in this market since the 1950s when the

    first commercial synthetic process was introduced.

    Methanol carbonylation has emerged as the domi-

    nant route to this product and currently over 60% of

    the world acetic acid is produced using this route.

    However, significant catalyst innovation has occurred

    even within this production route resulting in greatly

    improved yield, and selectivity at milder operatingconditions and lower cost of production. The lucra-

    tive nature of this market and the need for the major

    producers to continually protect their market position

    and investments is expected to drive further inno-

    vation within methanol carbonylation and the other

    promising technology options looming on the horizon

    that have been discussed in this paper.

    References

    [1] Asian Chemical News, 16 November 1998.

    [2] N. Kutenpov, W. Himmele, H. Hohenshutz, Hydrocarbon

    Proc. 45 (1966) 141.

    [3] F.E. Paulik, J.R. Roth, J. Am. Chem. Soc., Chem. Commun.

    (1968) 1578.

    [4] H.D. Grove, Hydrocarbon Proc. 51 (1972) 76.

    [5] US Patent 3769329 (1973) to Monsanto.

    [6] R.T. Eby, T.C. Singleton, Appl. Indian Catal. 1 (1983) 275.

    [7] M.J. Howard, M.D. Jones, M.S. Roberts, S.A. Taylor, Catal.

    Today 18 (1993) 325.

    [8] M.A. Murphy, B.L. Smith, G.P. Torrence, A. Agulio, J.

    Organomet. Chem. 303 (1986) 257.

  • 8/6/2019 8e0f.8810

    13/13

    N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253265 265

    [9] B.L. Smith, G.P. Torrence, M.A. Murphy, A. Agulio, J. Mol.

    Catal. 39 (1987) 115.

    [10] European Patent Application 0161874 (1985).

    [11] Japanese Patent Koukai 60-239434 (1985) to Celanese.

    [12] Japanese Patent Koukoku 04-69136 (1992) to Daicel.[13] Japanese Patent Koukoku 03-38256 (1991) to Daicel.

    [14] Japanese Patent Koukoku 07-23337 (1995) to Daicel.

    [15] R.F. Heck, J. Am. Chem. Soc. 85 (1963) 2013.

    [16] N. Ritzkalla, Industrial chemicals via C1 processes, Am.

    Chem. Soc. (1987) 61.

    [17] US Patent 4133963 (1979) to Eastman.

    [18] D. Foster, Adv. Organomet. Chem. 17 (1979) 255.

    [19] Japanese Patent Koukai 06-340573 (1994) to Poulenc.

    [20] European Patent Publication 643034 (1994) to BP.

    [21] European Patent Publication 728726 (1994) to BP.

    [22] European Patent Publication 752406 (1995) to BP.

    [23] Kagakukougyou Nippou, 12 July 1996.

    [24] T.W. Dekleva, D. Forster, Adv. Catal. 34 (1986) 81.

    [25] European Chemical News, 26 May1 June 1997.

    [26] D. Forster, T.C. Singleton, J. Mol. Catal. 17 (1982) 299.

    [27] D. Forster, T.W. Dekleva, J. Chem. Edu. 63 (3) (1986) 204.

    [28] D.J. Watson, Catalysis of Organic reactions, Marcel Dekker,

    New York, 1998, p. 369.

    [29] M. Gauss, A. Seidel, P. Torrence, P. Heymanns, Applied

    Homogeneous: Catalysis with Organometallic Compounds,

    Vol. 1, VHC, New York, 1996, p. 104.

    [30] Japanese Patent Koukai 7-503489 (1995) to Bohan Iron.

    [31] Japanese Patent Koukoku 5-28212 (1993) to Hoechst.

    [32] Japanese Patent Koukai 62-72636 (1987) to Hoechst.

    [33] Japanese Patent Koukai 46-5367 (1971) to Monsanto.

    [34] Japanese Patent Koukoku 60-54294 (1985) to Monsanto.

    [35] US Patent 4615806 (1986) to Celanese.[36] P.M. Maitlis, A. Haynes, G.J. Sunley, M.J. Howard, J. Chem.

    Soc., Dalton Trans. (1996) 2187.

    [37] R.G. Schultz, P.D. Montgromerry, J. Catal. 13 (1969) 105.

    [38] US Patent 3717670 (1973) to Monsanto.

    [39] A. Krzywicki, M. Marczewski, J. Mol. Catal. 6 (1979)

    431.

    [40] M.S. Scurrell, R.F. Howe, J. Mol. Catal. 7 (1980) 535.

    [41] US Patent 4657884 (1987) to Hoechst.

    [42] US Patent 4776987 (1988) to Hoechst.

    [43] R.S. Drago, E.D. Nyberg, A.E. Amma, A. Zombeck, Inorg.

    Chem. 20 (1981) 641.

    [44] M.S. Jarrel, B.C. Gates, J. Catal. 40 (1975) 255.

    [45] J. Hjortkjaer, Y. Chen, B. Heinrich, Appl. Catal. 67 (1991)

    269.

    [46] European Patent 277824 (1988) to Reilly.

    [47] US Patent 5334755 (1994) to Chiyoda.

    [48] US Patent 5364963 (1994) to Chiyoda.[49] US Patent 5576458 (1996) to Chiyoda.

    [50] N. Yoneda, T. Minami, J. Weiszmann, B. Spehlmann,

    Science and technology in catalysis, in: Proceedings of the

    Third Tokyo Conference on Advanced Catalytic Science and

    Technology, 1998, p. 93.

    [51] US Patent 5962735 (1999) to UOP.

    [52] D.J. Schreck, D.C. Busby, R.W. Wegman, J. Mol. Catal. 47

    (1988) 117.

    [53] G. Bud, H.U. Hog, J. Mol. Catal. A: Chem. 95 (1995) 1134.

    [54] Japanese Patent Koukai 7-60130 (1995) to Chiyoda.

    [55] M. Komatsu, M. Yoneoka, Kagaku Kougyo 43 (1988) 1134.

    [56] S. Shinoda, T. Yamakawa, J. Chem. Soc., Chem. Comun.

    (1990) 1151.

    [57] T. Yamakawa, P. Tsai, S. Shinoda, Appl. Catal. A L1 (1992)

    92.

    [58] European Patent Application 801050 (1997).

    [59] US Patent 5728871 (1998) to Topsoe.

    [60] British Patent 1020068 (1965) to Halcon.

    [61] Japanese Patent Koukai 47-13221 (1971) to National

    Distillers.

    [62] FR 1568742 (1967) IFP.

    [63] Japanese Patent Koukai 7-89896 (1995) to Denko.

    [64] Japanese Patent Koukai 9-67298 (1995) to Denko.

    [65] K. Sano, H. Uchida, Shokubai 41 (4) (1999) 290.

    [66] European Patent Application 294845 (1988) to UCC.

    [67] European Patent Application 407091 (1991) to BP.

    [68] European Chemical News, 29 March4 April 1999, p. 28.[69] World Patent 9805619 (1998) to Hoechst.

    [70] US Patent 5907056 (1999) to Sabic.

    [71] T. Nishiguch, et al., Chem. Lett. 20 (1992) 1141.

    [72] Chem. Week 20 (1994).

    [73] V.H. Agreda, J.R. Zoeller, Acetic Acid and Its Derivatives,

    Marcel Dekker, New York, 1993, Chapters 16, p. 3.

    [74] F.S.Wagner, Kirk-Othmer Encyclopedia of Chemical

    Technology, 4th Edition, Vol. 1, 1991, p. 121.

    [75] W.R. Moser, B.J. Marshik-Guerts, S.J. Okrasinski, J. Mol.

    Catal. A: Chem. 143 (1999) 57.

    [76] W.R. Moser, B.J. Marshik-Guerts, S.J. Okrasinski, J. Mol.

    Catal. A: Chem. 143 (1999) 71.