Добавил:
Upload Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:
перевод книги швед.doc
Скачиваний:
30
Добавлен:
20.11.2018
Размер:
8.39 Mб
Скачать

Xylosec

Mannosec

CKa,d 31.6 71 32 11 23 50 900 0.3 2.06 1.19 1.5 1.8

PSAQa,d 33.6 72 32 12 23 45 100 0.3 2.72 1.52 1.2 3.07

Sie,f6 29.0 79 20 6 28 80 000 0.33 –0.82

CK/Oa,d 17.2 62 21 18 13 33 000 0.45 1.28 1.36 1.57 1.93

PSAQ/Ob,d 18.7 65 20 19 13 39 100 0.41 2.0 1.56 1.22 2.86

Sie,g 17.9 85 17 8 21 76 400 0.54 0.18

a. Conventional kraft pulp (CK) after oxygen bleaching.

b. Polysulfide/anthraquinone pulp (PSAQ) after oxygen bleaching.

c. Carbohydrates in mg 100 mg–1 cellulose.

d. proRL; 6f RL; 7g NaCl RL; d, f and g indicate different purification

procedures.

e. Two stage neutral sulfite pulp (Si).

structures, an increase in the proportion of carboxylic acids, a decrease in the

molar mass, and an increase in the hydrophilicity were observed [62]. The arabinose

content increased and the galactose, as well as the mannose, content

decreased during oxygen bleaching (Tab. 7.10).

Tamminen and Hortling [62] showed that the CK pulp was the most hydrophilic

before oxygen delignification, with the best bleachability. The sulfite pulp lignin

was structurally quite different from the alkaline pulp lignins, and seemed to dissolve

during oxygen delignification without any major oxidation reactions [62].

7.3.2.3 Oxygen (Dioxygen) and its Derivatives

The element oxygen (chemical symbol O) exists in air as a diatomic molecule, O2,

which strictly should be called dioxygen. Over 99.7% of the O2 in the atmosphere

is the isotope oxygen-16 (16O), but there are also traces of oxygen-17 (17O, about

0.04%) and oxygen-18 (18O, about 0.2%) [125].

The solubility of dioxygen and the physical transport of dissolved dioxygen gas

in an aqueous phase are important properties. The model of Broden and Simonson

was used to estimate the solubility of oxygen in equilibrium conditions, as a

function of oxygen pressure, temperature and hydroxide ion concentration [126].

7.3 Oxygen Delignification 641

7.3.2.3.1 Dioxygen: Electronic Structure

The diatomic oxygen molecule O2 has two unpaired electrons, each located in a

different p* antibonding orbital, having the same spin quantum number or, as is

often written, having parallel spins (Fig. 7.24); this is the most stable state – or

ground state – of dioxygen. If dioxygen, which can act as an oxidizing agent,

attempts to oxidize another atom or molecule by accepting a pair of electrons

from it, then both of these electrons must be of antiparallel spin so as to fit in to

the vacant spaces in the p* orbitals (Fig. 7.24). However, a pair of electrons in an

atomic or molecular orbital would not meet this criterion, as they would have

opposite spins in accordance with Pauli’s principle. This imposes a restriction on

electron transfer that tends to make O2 accept its electrons one at a time, and contributes

to explaining why O2 reacts sluggishly with many nonradicals [125].

According to the law that electrons reacting with each other must have antiparallel

spin, ground-state dioxygen (triplet-state) cannot react with atoms or molecules in

the singlet state due to the spin restriction [127].

Fig. 7.24 A simplified version of bonding in the diatomic oxygen

molecule (16 electrons) in the ground state and of the

excited state of dioxygen and his reduction products [125].

7.3.2.3.2 Principals of Dioxygen Activation

Due to the electron-configuration, dioxygen takes up one electron at a time – a

process termed one-electron reduction [125,127]. By stepwise addition of electrons

to the molecular orbital of ground state dioxygen, the reduction products of oxygen

are formed (Fig. 7.25 and Scheme 7.3). On the addition of one electron, superoxide

is formed. A second electron produces peroxide. Two more produces 2 separated

oxides since no bonds connect the atoms (the number of electrons in antibonding

and bonding orbitals are identical). Each of these species can react with

protons to produce species such as HO2

_, H2O2 (hydrogen peroxide) and H2O.

642 7Pulp Bleaching

7.3 Oxygen Delignification 643

0 1 2 3 4

0

100

200

300

400

500

H

2

O

+e-

Activation energy ΔG

0

(kJ)

Reduction equivalent

O2

+e-

O2

-

+2H+

H

2

O

2

+e-

+H+ OH + H2O

+H+

+e-

Fig. 7.25 Energetics of oxygen-reduction [127].

O2 O2 H2O2 H2O + 2 H2O pH 7

- HO

- 0.33 - 0.89 + 0.30 + 2.40

+ 0.59

+ 1.20

+ 1.349

+ 0.29

+ 0.281

+ 0.815

Scheme 7.3 Dioxygen redox potentials at pH 7 [127].

Under the conditions used in dioxygen delignification, with the pH in the range

between 10 and 13, the standard redox potentials of the reactive species are substantially

reduced (Scheme 7.4) due to the lower potential of the ionized form.

O2

+e-, H+

HOOH H2O+ 2 H2O

pKa= 4.8 11.6 11.9

O2

-

H++ H++HOO- O- H+ +

+e-, H+

+e-, H+ +e-, H+

E - 0.33 0.20 - 0.03 1.77 0 at pH 14

Dioxygen

Hydroperoxyl

radical

Hydrogen

peroxide

Hydroxyl

radical

Superoxide

anion radical

Hydroperoxy

anion

Oxyl anion

radical

Oxygen species

Anionic form

45.71

Hydroxide

ion

OH-

HOO HO Water

Scheme 7.4 Dioxygen reductions proceeding in four consecutive

one-electron steps (E0 standard reduction potential)

(according to [128]).

Therefore, in order to initiate a reaction, increases of the temperature and the ionization

of functional groups (ionized phenolic hydroxyl groups on lignin) are necessary

to facilitate the electron transfer to dioxygen and its related species.

7.3.2.3.3 The Reactions of Dioxygen and its Reduction Products

Under alkaline conditions the reaction of dioxygen with an activated lignin model

compound (particularly a phenolate) generates a superoxide anion radical

[3,129,130]. This is generally the rate-determining step of the oxidation requiring

elevated temperatures [131] or the presence of metal ions [132], whereas the superoxide

anion radical can undergo a metal-catalyzed dismutation [133–135] forming

hydroperoxy anion that can further undergo a metal-catalyzed disproportionation

reaction forming a hydroxyl radical. The following equations show some interconversion

reactions that oxygen species can undergo. Most of these are extremely

rapid, but others [Eqs. (4) and (5)] are very slow; indeed, for some reactions, metals

or protons are required for catalysis [136].

_O_2 _OH_→OH__O2 _1_

HOO__HO_→HO___O_2 _2_

HO_ _ HO_→H2O2 _3_

_O_2 __O_2 _H2O→HO_2 _O2_OH_ _4_

_O_2 _H2O2→OH__HO_ _ O2 _5_

HOO__H2O2→HO_ _ _O_2 _H2O _6_

_O_2 _OH_ _ H2O→HOO__OH_ _7_

O2__O_→_O_3 _ __O_ O2_HOO_ _8_

7.3.2.3.4 Autoxidation

The term “autoxidation” comprises the oxidation with dioxygen [137] and a multitude

of free radical reactions catalyzed by transients in the system such as hydroxyl

radicals (autocatalysis) and superoxide anion radicals [1,7,138]. This chain process

with various phases of autoxidation starting with the initiation of the reaction

of dioxygen [131], being the least reactive, with the activated substrate (particularly

a phenolate) requires an elevated temperature [137] and/or the presence of heavy

metals [132], acting as redox catalysts [3], forming a superoxide anion radical and

a substrate radical [Eq. (9)]. As noted, oxygen bleaching must be conducted in an

alkaline environment (pH > 10) and a temperature of about 80–100 °C and beyond

to ensure reasonable rates. At higher pH (alkali charge) and temperature (about

120 °C), hydroperoxides decay homolytically to hydroxyl radicals. Whilst activation

644 7Pulp Bleaching

of the substrate and elevated temperatures are needed to initiate the reaction [Eq.

(9)], dioxygen, on the other hand, reacts very rapidly with any substrate radical to

the corresponding peroxyl radical [Eq. (10), propagation]. The recombination and

termination respectively is accomplished by coupling of two radicals and does not

require activation by ionization, as is needed for electron transfer.

Eq. (9): Initiation [4]

R__O2→R_ _ _O_2

RH _ HO_→R_ _ H2O

RH _ _O_2 →R_ _ HOO_

RH _ O2→R_ _ HOO_

Eq. (10): Propagation [4]

R_ _ O2→ROO_

ROO_ _ RΗ→ROOH _ R_

ROO_→R_ _ _O_2

R_ _ RΗ→RH _ R_

Eq. (11): Recombination – Termination

R_ _ HO_→ROH

R_ _ _O_2 →ROO_

R_ _ R_→R__R

ROO_ _ ROO_→ROOR _ O2

ROO_ _ R_→ROOR

The initiation step above occurs mostly at C atoms which can produce the most

stable free radicals (allylic, benzylic position, and 3 > 2 >> 1c arbons).Hence, unsaturated

fatty acids are extra-reactive at themethyleneCthat separates the double bonds.

7.3.2.3.5 Singlet O2 – Excited State

Singlet dioxygen, with a lifetime of about 0.06 s, can be generated from triplet

dioxygen by photoexcitation [127,139]. Alternatively, it can be made from triplet

oxygen through collision with an excited molecule (photosensitizer), which relaxes

to the ground state after a radiationless transfer of energy to triplet oxygen to form

reactive singlet oxygen [Eq. (12)] [125,140,141]. Furthermore, singlet oxygen is

generated by the reaction between a hydroxyl radical and a hydroperyl radical [Eq.

(13)], and between HOCl and peroxide [Eq. (14)] [142], the oxidation of the superoxide

anion radicals with heavy metals [Eq. (15)] or ozone [Eq. (16)], and in consequence

of the decay of polyoxides.

7.3 Oxygen Delignification 645

O2

hm

Photosensitizer _____

O2

_1Dg_ _12_

HO_ _ HOO_→O2

1Dg _ __H2O _13_

HOOH _ OCl_→Cl__H2O _ O2

1Dg _ _ _14_

_O_2 _Fe3_→Fe2_ _O2

1Dg _ _ _15_

O3 _ _O_2 →O_3 _O2

_1Dg_ _16_

Alkenes (double-bonds) react with oxygen to form hydroperoxides, potentially

through an epoxide intermediate, and dienes reacts with oxygen in a Diels–Alderlike

reaction to form endoperoxides.

7.3.2.3.6 Superoxide Anion Radical

This is generated during the initiation step of the autoxidation [Eq. (9)], and undergoes

several interconversion reactions with other dioxygen-derived species

[Eqs. (1), (4), (5), and (7)]. In the presence of metal ions, the superoxide anion radical

can be oxidized to dioxygen [Eq. (17)] or reduced to the hydroperoxy anion

[Eq. (18)] [125]. The reduction of Fe3+ by the superoxide anion can accelerate the

Fenton reaction, giving a superoxide-assisted Fenton reaction [Eq. (19)] [125].

_O_2 _Fe3_→Fe2_ _O21_5 _ 108M_1s_1 _17_

_O_2 _Fe2_ _H_→Fe3_ _HO_2 1 _ 107M_1s_1 _18_

Fe2_H2O2→Fe3___OH _ OH_

Fe3__ _O_2 →Fe2__O2

H2O2__O_2

Fe

catalyst __ _OH _ OH__O2

_19_

Due to its low oxidation potential, the superoxide anion radical is highly selective.

It is a very strong Bronsted base capable of accepting a hydrogen from weak

acidic structures, and preferentially reacting with dihydroxy structures through

deprotonation followed by dehydration.

7.3.2.3.7 Hydrogen Peroxide

In contrast to dioxygen, which contains multiple bonds between the O atoms,

hydrogen peroxide has only one bond, which can be easily broken. Remember,

bonds can be broken in a heterolytic manner (both electrons in a bond go to one

of the atoms), or in a homolytic fashion, in which one electron goes to each atom.

646 7Pulp Bleaching

During dioxygen delignification, hydrogen peroxide [143] and the hydroperoxide

anion respectively evolve in situ from:

_ nucleophilic substitution reactions

O

OOH

O

OH

+ OH -

+ H2O2

and

_ disproportionation reactions 2_O_2 _H2O _

H_ HOO__HO__O2

2_O_2 _2H_ HOOH _ O2

Some properties and reactions of hydrogen peroxide include the following:

_ Acid/Base: H2O2

pKa1

_H_ _ HO_2

pKa2

_H__ O2_ 2

(pKa1 = 11.8; pKa2 > 16–18 [127] or 30 [128]; see Scheme 7.4)

_ Reaction with Fe2+: The Fenton reaction: HOOH + Fe2+ →Fe3+ + HO· + HO–. In

this reaction, a homolytic cleavage of the O–O bond occurs, generating OH– and

the hydroxyl radical (OH·), which will react with any molecule it encounters.

The hydroxyl radical may be formed via an oxoiron(IV) intermediate [144]. The

peroxide can also be effective as an oxidant, and in a transition metal-induced

cleavage of the H–OO bond the hydroperoxyl radical (HOO_) is formed:

HOOH + Fe3+ →Fe2+HOO· + H+

_ Thermal or photochemical homolytic cleavage of hydrogen peroxide:

HOO_ _ HOOH Energy _ H2 _ _HO _ _H_

7.3.2.3.8 Hydroxyl Free Radical

As mentioned earlier, this species is extremely reactive [136]. It will react with any

molecule it encounters, and does so immediately. It can abstract a H atom, leaving

another free radical. The anionic form, the oxyl anion radical (see Scheme 7.4),

displays properties that are distinctly different from those of the hydroxyl radical.

In contrast to the latter, the oxyl radical reacts predominantly by hydrogen abstraction

and is therefore probably less selective than the hydroxyl radical [145].

Note: The terms hydroxyl free radical and hydroxyl radical are used synonymously.

Care must be taken using the term hydroxyl ion, which is the synonym

for the hydroxide ion (OH–).

7.3.2.3.9 Electrophilic–Nucleophilic Reactions

As noted, delignification during bleaching is initiated by electrophilic reactions,

which may be followed by nucleophilic processes [6–9]. The reactive oxygen species

(ROS) are listed in Tab. 7.11, according to their electrophilic–nucleophilic

character. Under the conditions of oxygen- alkali bleaching, the hydroperoxyl radical

is deprotonated to produce the superoxide anion radical. About half of the hydroxyl

radical is present as its base the oxyl anion radical (Tab. 7.11; see also

Scheme 7.4), and about half of the hydroperoxy anion is present as hydrogen peroxide

(Scheme 7.4).

7.3 Oxygen Delignification 647

Tab. 7.11 Reactive oxygen species (ROS) listed according to

their electrophilic – nucleophilic character.

Electrophiles

Triplet dioxygen 3O2

Hydroperoxyl radical HOO_

____

pKa_4_8

_O_2 Superoxide anion radical

Hydroxyl radical HOO_

____

pKa_11_9

_O_2 Oxyl anion radical

Nucleophiles

Hydroperoxy anion HOO–

Singlet dioxygen 1O2

The sites of electrophilic and nucleophilic attacks in lignins are shown in

Fig. 7.26. The p-system of the aromatic ring can be overlapped by the lone electron

pairs on the oxygen atom in para-hydroxy and the para-alkoxy groups, creating

centers of high electron density (Fig. 7.26), that can be attacked by electrophiles.

High electron density (d-) also appears at the Cb atom of aliphatic double bonds

O-

R2 OCH3

HC R1

arylalkane unit

R1 = OH, OAr or OAlk

arylpropene unit

O

R2 OCH3

CH2

CH

δ- δ-

δ-

H2C R1

δ-

O

R2 OCH3

C

C

H2C R1

R δ-

O-

R3

- δ-

α-carbonyl group containing

R = OAr, Ar or Alk

O

R2 OCH3

CH

δδ

δ

arylalkane unit

R1 = OH, OAr or OAlk

arylpropene unit

quinone-methide intermediate

O

R2 OCH3

CH

HC

CH2

O

R2 OCH3

C

C

CH2

R

O

R3

α-carbonyl group containing

R = OAr, Ar or Alk

δ

δδ

δ

δ

δ

δ

δ

C C C C O

δδ

C

C O-

C

C O

-

δ-

δ-

ELECTROPHILIC

NUCLEOPHILIC

O

R2 OCH3

CH R1

δ- - δ-

δ-

-

O-

R2 OCH3

CH

CH

H2C R1

Fig. 7.26 Sites of electrophilic (d-) and nucleophilic (d+)

attacks in lignin (adapted from Ref. [2]).

648 7Pulp Bleaching

conjugated to the aromatic ring. By elimination of an a– (see Section 4.2.4, Chemistry

of kraft pulping, Scheme 3) or, in conjugated structures, a c-substituent, a

quinone-methide intermediate is formed from the arylalkane unit (Fig. 7.26),

which involves the loss of two electrons, resulting in the generation of centers of

low electron density (d+) that constitute the sites of attack by nucleophiles [2].

7.3.2.4 A Principal Reaction Schema for Oxygen Delignification

Over 30 years of research into the oxidation of lignin and lignin model compounds

with dioxygen has now elapsed, and has provided insights into the reactions

involved in the degradation, and their mechanisms. Based on the reaction

products formed from the degradation of lignin and lignin model compounds

with dioxygen and with ROS generated during bleaching, a number of mechanisms

have been proposed. Several excellent reviews have been produced on the

mechanisms involved in lignin degradation [1,2,4,6,7,72,101,122,138,146,147] and

the reactive species present in these reactions [3,9,90,129,130], including their

selectivity. The latter remains of interest [148,149], especially in connection with

protective systems and additives [94,150–156]. In addition, an excellent book on

oxygen delignification chemistry was published a few years ago [157]. It is impossible

to cover all of these mechanisms in detail within this chapter; thus, a general

summary with selected mechanisms will be provided.

Oxygen delignification is actually based on the competitive reactions of oxygen

or ROS within pulp lignin and carbohydrates [94]. Lignin removal under alkalioxygen

conditions is accompanied by a kinetically less favorable oxidation of carbohydrates,

whereas the oxidation of the carbohydrates becomes a more favorable

process when the kappa number decreases [94]. The reaction of phenolic compounds

with oxygen produces ROS, namely the hydroxyl radical (_OH), which

can degrade nonphenolic (model) compounds.

As shown previously (see Scheme 7.1), the initial step in oxygen-alkali bleaching

is the formation of the phenoxyl radical as a consequence of an electrophilic

attack by oxygen (Scheme 7.5A). Moreover, the hydroxyl radical formed during

oxygen treatment [Eqs. (5), (6), and (19)] is also capable of generating a phenoxyl

radical (Scheme 7.5B) being reduced to the hydroxide anion (OH–).

A principal reaction schema for oxygen delignification [3,6,7,138] starts with the

generation of hydroperoxides, which are key intermediates in the oxidation of lignins

and carbohydrates. They can be formed either by electrophilic or nucleophilic

reactions:

_ Formation of hydroperoxides [8,138]

_ Fragmentation of hydroperoxides (homolytic – forming radicals;

or heterolytic – forming hydrogen peroxide, singlet oxygen)

[8,138]

_ Involvement of the radicals in the bleaching process [3,9,129,130]

7.3 Oxygen Delignification 649

OH

OCH3

CH

CH

CH2OH

δ- δ-

δ-

δ-

O-

OCH3

CH

CH

CH2OH

+ OH-, - H2O

O

OCH3

CH

CH

CH2OH

O2 O2

-

1 2a 3

A

O-

OCH3

CH

CH

CH2OH

O

OCH3

CH

CH

CH2OH

OH

2b 3

OH

OCH3

CH

CH

CH2OH

1

OH-

B

+ OH-, - H2O

+ HO

Scheme 7.5 Formation of the phenoxyl radical by oxygen (A)

and the hydroxyl radical (B).

O

R1 OCH3

C

C R

O

R1 OCH3

C

C R

O

R1 OCH3

C

C R

O

R1 OCH3

C

C R

O

R1 OCH3

C

C R

O -O

O

R1 OCH3

C

C R

O

OCH3

R1

C

C R

O O-

4 5a 6 7 8 9

O

R1 OCH3

C

C R

O -O

O-

R1 OCH3

C

C R

O

O

O

R1 OCH3

C

C R

OH

O

4 5b 10 11 12 13

14 15 16 17 18 19

-O

OCH3

R1

C

C R

O

O

O

OCH3

R1

C

C R

O O-

HO

OCH3

R1

C

C R

O

O

+ H2O, - OH-

O

R1 OCH3

C

C R

O

R1 OCH3

C

O C R -O

O

R1 OCH3

C

O C R -O

O-

R1 OCH3

C

O C R

O

OH

R1 OCH3

C

O C R

O

+ O2

-

+ O2

-

+ O2

-

+ H2O, - OH-

+ H2O, - OH-

R = H, OAr, Ar or Alk

Scheme 7.6 Formation of hydroperoxide intermediates in

alkaline media followed by an intramolecular nucleophilic

attack of the hydroperoxide anions (adapted from Ref. [6]).

650 7Pulp Bleaching

C

C

C

C

O

R

H

C

C

C

C

O

R

+ OH - (-)

C

C

C

C

O

R

+ O2, - HOO

+ HOO , - HOOH

+ HO , - HOH

R = H, OH, organic moiety

+ HOO

+ O2

C

C

C

C

O

R

OOH

20 21 22 23

O-

O-

+ O2, + H+

OOH

O

O-

O

O

+ HOO-

24 25 26

Scheme 7.7 Formation of hydroperoxides in the autoxidation

of enolisateable and enediol structures, and the formation of

the hydroperoxy anion (adapted from Refs. [4,6]).

The abstraction of an electron from phenolate anions by oxygen (or the hydroxyl

radical) (Scheme 7.5) yields phenoxyl radicals (Scheme 7.6, 4 and 14) and the

mesomeric cyclohexadienonyl radicals (5a and 5b) or “quinone methide” radicals

(15). The superoxide anion radicals then form hydroperoxide intermediates (6 and

10) with the mesomeric cyclohexadienonyl radicals or the b-radical (15). A nucleophilic

attack by the peroxide anions on the carbonyl carbon (11) or a vinylogous

carbon of the cyclohexadienone- (7) or quinone methide (17) moieties yields the

corresponding dioxetane intermediates (8, 12 and 18). Intermediate 8 finally form

an oxirane structure (9). The rearrangement of 12 results in an opening of the peroxide

ring and heterolytic cleavage of the carbon–carbon bond, giving a “muconic

acid” ester (13), and 18 is fragmented by scission of the Ca–Cb bond of the former

conjugated double-bond forming the corresponding aldehydes (19) and/or a

ketone, depending on the nature of R.

The hydroperoxy intermediates formed during the autoxidation of phenolic

(Schemes 7.6 and 7.8) and enolic (Scheme 7.7) structures in lignin and carbohydrates

can be displaced by the hydroxide ion via a SN2 reaction (2829), or the

bond can be cleaved heterolytically giving the hydroperoxy anion, which is

described elsewhere. Homolytic decomposition of hydroperoxy intermediates produces

phenoxy (31) and hydroperoxy (Scheme 7.8) radicals. The latter can be

reduced to the hydroperoxy anion.

The hydroxyl radical reacts with the main components of wood, and attacks

preferentially electron-rich aromatic and olefinic moieties in lignin. It also reacts

with aliphatic side chains in lignin and carbohydrates, but at a lower rate. Depending

on the pH, the hydroxyl radical is converted to its conjugate base, the oxyl

anion radical (see Scheme 7.4). The oxyl anion radical does not react with electron-

rich structures, but rather with aliphatic side chains in lignin and carbohydrates.

The first step in all reactions of the hydroxyl radical with aromatic substrates

(Scheme 7.9, 32) is a rapid addition to the p-electron system of the aromatic

ring forming a short-lived charge-transfer adduct (33) that decays under

alkaline conditions to give isomeric hydroxycyclohexadienyl radicals (34 and 37).

7.3 Oxygen Delignification 651

O-

R1 OCH3

O

R1 OCH3

+ O2, + H+

OOH

O

R1 OCH3

OH

+ OH -

+

- HOO

O

R1 OCH3

- HOO

+ e-

HOO-

27 28 29 30

31

HOO-

Scheme 7.8 Formation of hydroperoxides in the autoxidation

of phenolic structures, and the formation of the hydroperoxy

anion (from Ref. [6]).

OR

OCH3

R

OH

33

OR

OCH3

R

32

OH

OR

OCH3

R1

34

H

HO

OR

OCH3

R1

37

or OH

Scheme 7.9 Formation of hydroxycyclohexadienyl radicals

(adapted from Ref. [6]).

The hydroxycyclohexadienyl radical can be oxidized by addition of oxygen

(Scheme 7.10) followed by alkali-promoted elimination of the superoxide anion

radical forming a cation radical (35, 38 and 42) and elimination of a proton (rearomatization)

leading to hydroxylation (Scheme 7.10, path A, 36) or, in combination

with elimination of methanol and cleavage of an alkyl-aryl ether bond, to dealkoxylation

(path B) with formation of ortho-quinonoid structures (39). From conjugated

structures (Scheme 7.10, path C), “quinone methide” intermediates (41,

42 and 43) are formed, giving glycolic structures (44) by adding hydroxide ions

that undergo oxidative cleavage of the glycolic C–C bond (44) [6].

The hydroxyl radical adducts (Scheme 7.11, 45) can undergo disproportionation

reactions from which the same oxidation products (46 and 47) arise, together with

the corresponding reduction products (48 and 49) [6].

652 7Pulp Bleaching

OR

OCH3

R1

OH

33

OR

OCH3

R1

34

OR

OCH3

R1

OH

33

R = H or alkyl

H

HO

+ O2

- O2

-

OR

OCH3

R1

35

H

HO

+

- H+

OR

OCH3

R1

36

HO

OR

OCH3

R1

37

+ O2

- O2

-

OH

OR

OCH3

R1

38

OH

+

- CH3OH, - H+

( - ROH )

O

O

R1

39

O-

OCH3

C

OH

40

C

O-

OCH3

C

41

C OH

O-

OCH3

C

42

C OH

+ O2

- O2

-

+

O

OCH3

C

43

C OH

+ OH -

O-

OCH3

C

44

C

OH

OH cleavage

A

B

C

Scheme 7.10 Reactions of the hydroxyl radical adducts of aromatic

and ring-conjugated structures (adapted from Ref. [6]).

O-

OCH3

R1

OH

45

O-

OCH3

R1

HO

46

O

O

R1

47

O-

OCH3

R1

48

O-

OH

R1

49

O-

OCH3

R1

OH

45

Disproportionation

+

Scheme 7.11 Disproportionation of hydroxyl radical adducts

(adapted from Ref. [6]).

Another reaction mode of the hydroxycyclohexadienyl radical (Scheme 7.12, 51

and 56) is the elimination of the hydroxyl radical as hydroxide anion (Scheme

7.12, paths A and B). This results in the formation of cation radicals (52 and 57)

followed by the generation of side-chain oxidation products and products of homolytic

Ca–Cb bond cleavage (58). The elimination of a proton leads to a re-aromatization

(59).

7.3 Oxygen Delignification 653

OR

OCH3

CH

OH

50

- OH -

OR

OCH3

HC

OH

55

CR1

OH

A

B

R = H or alkyl

OR

OCH3

CH

51

OH

OR

OCH3

CH

52

+ - H+

OR

OCH3

CH

53

OR

OCH3

C

54

R = alkyl; R1 = aryl or aroxyl

further oxidation

OR

OCH3

HC

56

CR1

OH

HO

H

OR

OCH3

HC

57

CR1

OH

- OH -

+

OR

OCH3

HC

58

CR1

OH

+

OR

OCH3

HC

59

O

- H+

Scheme 7.12 Reactions of the hydroxyl radical adducts of

aromatic and side chain structures (adapted from Ref. [6]).

Elimination of the hydroxyl radical as hydroxide anion results in the formation

of a cation radical (62 and 63), followed by a phenolic coupling (64) (Scheme 7.13)

and elimination of two protons to form a diphenyl (5–5) structure (58). The formation

of diphenyl structures is an undesirable reaction, because the 5–5 bond is

very stable and can hardly be cleaved.

O-

OH

60

H3CO

O-

61

H3CO

H

OH

O-

62

H3CO

- OH -

+

O

63

H3CO

2 x

O

64

H3CO

O

OCH3

H

H

O-

65

H3CO

O-

OCH3

- H+

Scheme 7.13 Phenolic coupling of the hydroxyl radical

adducts of aromatic structures (adapted from Ref. [6])

Singlet oxygen that can be generated during oxygen bleaching in different ways

[Eqs. (13–16)] has been of growing research interest for the past few years

[140,142,158–180]. Both, lignin model compounds and pulp have been investigated.

However, in most of the studies photosensitizers, such as rose bengal

654 7Pulp Bleaching

[159,162,164–167,174,179], methylene blue [140,141,160,163,169,175,177] or titandioxide

(TiO2) [139,169,175,177] have been used to generate singlet oxygen using

light from the visible range to the UV, the latter also used for direct irradiation of,

for example, an a-carbonyl group-containing lignin. Alternatively, singlet oxygen

was produced from sodium hypochlorite (NaOCl) and hydrogen peroxide [142]

according to Eq. (14). Some of these studies were performed in organic solvents

[162,167,174,176] and others in aqueous alkaline solution [142,168,177,179,181],

with the latter category being of main interest for this chapter. The photo- and radiation

chemical-induced degradation of lignin model compounds have been

summarized in a very good review [171], including other ROS, and the photochemical

oxidation of lignin models in the presence of singlet oxygen has been

studied by using ab initio calculations [178].

As mentioned, singlet oxygen has a pronounced electrophilic character, and hence

reacts well with electron-rich groups such as olefinic or aromatic derivatives. These

electron-rich groups tend to form an intermediate exciplex as a result of charge transfer

reactions between the electron-rich substrate and the singlet oxygen. This exciplex

is able to later form dioxetanes, hydroperoxides, or endoperoxides.

Photosensitized degradation studies of a-carbonyl group-containing lignin

model compounds (Scheme 7.14) show that a hydrogen atom transfer from the

phenolic OH group (66) to 1O2 might occur, leading to a phenoxyl radical (67) and

subsequently to quinonoid species (path A). However, formation of an endoperoxide

(68) leading ultimately to p-quinones (70) is also possible (path B).

H OH

HO H

H OR

H OH

R1

OH

HO H

H OR

H OH

R1

H OH

O2/OH-

O OH

HO H

H OR

H OH

R1

O

O HO

HO H

H OR

H OH

R1

OH

O

OH

H

+

OH-

I

II

I

1. -HOO-

2. -HOH

O

HO H

H OR

H OH

R1

O

OH

HO H

H OR

H OH

R1

BAR

H O HO O

HO O

HO H

H OR

H OH

R1

II

1 2 3 5

8

7

4

R1 = -H for xylan

BAR = Benzilic Acid Rearrangement R = Polysaccharide chain

H

HO

HO H

H OR

H OH

R1

HO O

6

+

R1 = -CH2OH for cellulose and glucomannan

- HCOOH

- ROH

+ OH-

HO O

HO

H

H OH

R1

9

+ OH-

Degradation

products

1 D-Glucose

2 1-Hydroperoxy-ketose

3 2-Hydroperoxy-aldose

4 D-arabino-hexosulose

5 Gluconic acid

6 Mannonic acid

8 Arabinonic acid

9 3-Deoxy-D-glycero-2-keto-pentonic acid

1 D-Xylose

2 1-Hydroperoxy-ketose

3 2-Hydroperoxy-aldose

4 D-threo-pentosulose

5 Xylonic acid

6 Lyxonic acid

8 Threonic acid

9 3-Deoxy-2-keto-tetronic acid

Scheme 7.14 Photodegradation of a-carbonyl group-containing

lignin model compounds (from Ref. [171]).

Moreover, a-carbonyl-containing b-O-4 lignin model compounds intensively

used in singlet oxygen degradation studies have been degraded to products deriving

from b-C–O bond cleavage. The main reactions were conversion of phenolic

aromatic units into carboxylic acids and cleavage of the b-O-4 ether bonds, leading

to a depolymerization of the lignin framework into smaller fragments [177]. Cleavage

of the b-O-4 aryl ether bond has been found for phenolic as well as nonpheno-

7.3 Oxygen Delignification 655

656 7Pulp Bleaching

lic derivatives [162]. Photochemical oxidation of the phenolic b-O-4 aryl ether gave

the same type of product, which confirmed that, in this case, the presence of the

carbonyl group is not indispensable for the cleavage reaction to occur [162]. When

the phenoxy portion of the molecule [1-(4-hydroxy-3-methoxyphenyl)-2-(2,6-

dimethoxyphenoxy)-3-hydroxy- 1-propanol] shows a lower reactivity towards singlet

oxygen, the oxidation of the phenol moiety to hydroquinone can occur. The

photochemical behavior of this model compound can be rationalized from a reaction

of singlet oxygen with the phenoxy part of the molecule [162].

Due to the unknown real contribution of singlet oxygen to lignin degradation

during oxygen bleaching, and the fact that in processes interconversions between

reactive species occur, this section of the text will be minimized.

One example of a rose bengal photosensitized degradation of loblolly pine

(Pinus taeda) kraft pulp, the final product of which contained 4% by mass of residual

lignin with the remainder being carbohydrates, is presented [179]. In this

study, the reactivity of singlet oxygen with kraft softwood substrates with respect

to the chemistry of lignin and cellulose has been investigated. The results revealed

that, despite the relatively high selectivity of singlet oxygen for lignin aromatic

units, degradation of the cellulose nevertheless occurred after approximately 50%

removal of the lignin. A decrease was observed in the number of aliphatic hydroxyls

(17%), condensed phenolics (4%), and guaiacyl phenolics (7%), and an

increase in carboxylic acids (54%). This result is typical of what is observed in the

reactions of ground-state oxygen with pulp or lignin, and suggests that despite the

initial electrophilic reactions of singlet oxygen with lignin, it is likely that ensuing

oxidations follow some of the typical reactions associated with ground-state oxygen

reactions, such as ring additions by hydroperoxide and oxygen followed by

ring openings to the muconic esters and acids. However, unlike ground-state oxygen

reactions, the levels of condensed phenolics (e.g., conjugated lignin monomers

at the C5 positions of the benzene moieties) were reduced during the singlet

oxygen reactions. Thismay be a consequence of the high electrophilic reactivity of singlet

oxygen, and was tested by subjecting substrates enriched in condensed phenolics

to singlet oxygen reactions [179]. The most salient difference between this systemand

a typical ground-state oxygen delignification system is the absence of condensed

phenolic units in the lignin. Subsequently, it was discovered that both the condensed

and noncondensed (guaiacyl) units react well with singlet oxygen [179].

This finding is important since 5-condensed phenolic subunits (5–5 and diphenylmethane;

DPM) in lignin are quite resistant. Their relative robustness does

not, however, appear to be the main rationale for the inactivity of lignin towards

oxygen delignification, but serves to suggest that the nature and reactivity of the

free phenolics deserve increasing scrutiny [182].

Residual lignins isolated from unbleached and oxygen-bleached eucalyptus

kraft pulps by acid hydrolysis and dissolved lignins in the kraft cooked and oxygen-

bleached liquors were studied, and the results compared with the corresponding

residual lignins. The data showed that etherified syringyl structures were

quite resistant towards degradation in the oxygen bleaching, causing little depolymerization

in residual lignin and a small increase in carboxylic acid content, but

producing appreciable amounts of saturated aliphatic methylene groups [105].

7.3 Oxygen Delignification 657

7.3.2.5 Carbohydrate Reactions in Dioxygen-Alkali Delignification Processes

The reactions of wood polysaccharides during dioxygen-alkali treatment can be

classified according to Malinen [183] into the following main categories:

_ Stabilization of the reducing end-groups.

_ Peeling reactions starting from the reducing end-groups.

_ Peeling reactions starting from stabilized end-groups.

_ Cleavage of the polysaccharide chain.

Reaction steps involving dioxygen are drawn with thicker lines (bold) and the

numbers given in italic.

7.3.2.5.1 Stabilization of the Reducing End-Groups

The rapid stabilization of the reducing end-groups of polysaccharides by transformation

to aldonic acid end-residues has been considered to be one great advantage

of the dioxygen-alkali delignification of wood or pulp [184–186]. Under the conditions

of dioxygen-alkali treatment, oxidation of the glucose unit (1) may proceed

via a 1-hydroperoxy-ketose (2 [187]) and a 2-hydroperoxy-aldose (3) (Scheme 7.15).

The hydroperoxy-group can easily be replaced by a hydroxide anion followed by

dehydration (path I) resulting in a a, b– dicarbonyl (glucosone = d-arabino-hexosulose,

4), which converts into gluconic acid (5) and mannonic acid (6) via benzilic

acid rearrangement (BAR) (see Section 4.2.4.2, Carbohydrate reactions). Glucosone

(d-arabino-hexosulose) end-groups have been suggested to be intermediates

in the formation of aldonic end-residues [188,189], and Theander [185] stated that

the fact that mannonic acid and gluconic acid end-residues are obtained on cellulose

treatment with dioxygen in basic solution is the best support for the view that

glucosone is really an intermediate. Alternatively, the hydroperoxy-intermediates

are split to formic acid (7) and arabinonic acid (8) (path II), the latter being converted

to 3-deoxy-d-glycero–2-keto-pentonic acid (9) and further degraded.

H OH

HO H

H OR

H OH

R1

OH

HO H

H OR

H OH

R1

H OH

O2/OH-

O OH

HO H

H OR

H OH

R1

O

O HO

HO H

H OR

H OH

R1

OH

O

OH

H

+

OH-

I

II

I

1. -HOO-

2. -HOH

O

HO H

H OR

H OH

R1

O

OH

HO H

H OR

H OH

R1

BAR

H O HO O

HO O

HO H

H OR

H OH

R1

II

1 2 3 5

8

7

4

R1 = -H for xylan

BAR = Benzilic Acid Rearrangement R = Polysaccharide chain

H

HO

HO H

H OR

H OH

R1

HO O

6

+

R1 = -CH2OH for cellulose and glucomannan

- HCOOH

- ROH

+ OH -

HO O

HO

H

H OH

R1

9

+ OH -

Degradation

products

1 D-Glucose

2 1-Hydroperoxy-ketose

3 2-Hydroperoxy-aldose

4 D-arabino-hexosulose

5 Gluconic acid

6 Mannonic acid

8 Arabinonic acid

9 3-Deoxy-D-glycero-2-keto-pentonic acid

1 D-Xylose

2 1-Hydroperoxy-ketose

3 2-Hydroperoxy-aldose

4 D-threo-pentosulose

5 Xylonic acid

6 Lyxonic acid

8 Threonic acid

9 3-Deoxy-2-keto-tetronic acid

Scheme 7.15 Stabilization of reducing end-residues through

formation of aldonic acids (5) and mannonic acid (6)

(adapted from Malinen [183] and Theander [185]).

In the absence of dioxygen, large amounts of 3-deoxy-pentonic acids are formed

and under oxidative conditions arabinonic, erythronic and mannonic acids are the

major reaction products [190]. A relative composition of aldonic acid residues

from various treatments is shown in Tab. 7.12.

Tab. 7.12 Relative composition (mol. %) of aldonic acid residues

from various treatment (from Ref. [185]).

From d-glucosone From cellulose

Acid NaOH/air

0.04 M, 100 °C

4 h [189]

NaOH/O2

0.04 M, 95 °C

1bar , 5 min [187]

NaOH/N2

0.04 M, 95 °C

1bar , 5 min [187]

NaOH/air

18%, 25 °C

200 h [191]

NaOH/O2

0.5%, 100 °C

5 bar, 2 h [184]

Mannonic 11 18 47 15 27

Gluconic 2 5 5 2 3

Arabinonic 58 37 26 58 50

Ribonic 4 6 0 2 2

Erythronic 25 35 22 23 18

Two different pathways can form erythronic acid (11) (Scheme 7.16). The first

entails rearrangement of the glucosone to d-erythro–2,3-hexodiulose (10), followed

by an oxidative cleavage and loss of glycolic acid (12) [183]. In the second pathway,

erythronic acid (11) results from alkaline and oxidative degradation of the glucosone

(4) through arabinose (13) and arabinosone (14) as intermediates. In the

absence of dioxygen arabinose (13, Scheme 7.16) and arabinonic acid (8, Scheme

7.15), it may be formed by hydroxide ion attack at C1 and C2 respectively [185].

Minor amounts of 3-deoxy-pentonic acids (17, 18) are formed from an arabinose

intermediate (13), and the main pathway starts with a direct b-hydroxy-elimination

in the glucosone (4) followed by loss of the elements of carbon monoxide

from the intermediate 4-deoxy-d-glycero–2,3-hexodiulose (15) [187].

The yield of 3-deoxy-pentonic acids is lower in the presence of dioxygen [185],

and the formation of arabinonic and erythronic acid is particularly important.

Theander [185] stated that an attack of dioxygen to the glucosone (4, Scheme 7.17)

should give a hydroperoxide (20), which should further yield arabinonic acid (8)

and carbon dioxide. A similar attack at C3 could, via formation of a hydroperoxide

(21), result in the formation of an erythronic acid end-group (11) plus glyoxylic

acid (22).

About the same proportions of aldonic acids were produced from glucosone

and glucose treated with dioxygen and alkali [183], and cellobiose [190] and cellotriose

[192] yielded glucosyl- and cellobiosyl-arabinonic acids as the main products.

However, the presence of the substituted erythronic and mannonic acids was

also significant, especially at higher alkali concentrations. Malinen and Sjostrom

658 7Pulp Bleaching

R1 = -H for xylan

R = Polysaccharide chain

R1 = -CH2OH for cellulose and glucomannan

O

HO H

H OR

H OH

R1

O

4

H

+ OH-

CH2OH

O

O

H OR

H OH

R1

10

+

O2/OH-

COOH

H OR

H OH

R1

11

CH2OH

COOH

12

7

- HCOOH

+ OHH

O

H

H OR

H OH

R1

13

HO

O2/OH- O2/OH-

H O

O

H OR

H OH

R1

14

COOH

H OR

H OH

R1

11

H O

H

H OH

R1

OH + OH-

- ROH-

H O

H

H OH

R1

16

O

H

O

OH

H

H OH

R1

H O

+ OH-

- ROH

O

O

H

H OH

R1

O

15

H

H

HO O

H

H OH

R1

17

HO

H

H

HO O

H

H OH

R1

18

H

H

OH

HO O

H

H OH

R1

19

+ + H

10 D-erythro-2,3-hexodiulose

11 Erythronic acid

12 Glycolic acid

13 Arabinose

14 Arabinosone

15 4-Deoxy-D-glycero-2,3-hexodiulose

16 3-Deoxy-D-glycero-pentosulose

17 3-Deoxy-D-threo-pentonic acid

18 3-Deoxy-D-erythro-pentonic acid

19 3,4-Dihydroxybutyric acid

10 D-glycero-2,3-pentodiulose

11 Glyceric acid

13 Threose

14 Threosone

15 4-Deoxy-2,3-pentodiulose

16 3-Deoxy-Tetrosulose

17 2,4-Dihydroxybutyric acid

18 2,4-Dihydroxybutyric acid

19 2-Deoxy-glyceric acid

Scheme 7.16 Degradation pathways of the glucosone and

xylosone end-groups (adapted from Malinen [183] and

Theander [185]).

R = Cellulose chain

R1 = -CH2OH

O

HO H

H OR

H OH

R1

O

4

H

+

O2/OHCOOH

H OR

H OH

R1

11

C

COOH

22

O

HO H

H OR

H OH

R1

O

20

H

O

HO O2H

H OR

H OH

R1

O

21

H

O2H

H

- CO2

HO O

HO H

H OR

H OH

R1

8

H O

O2/OHO2/

OH-

Scheme 7.17 Degradation pathways of the glucosone endgroups

to the formation of arabinonic acid (8) and erythronic

acid (11) (adapted from Theander [185]).

7.3 Oxygen Delignification 659

[192] reported that, when hydrocellulose was subjected to dioxygen-alkali treatment,

erythronic acid was the dominating end-group, and that the reaction conditions

actually have a marked effect on the composition of the aldonic acid endgroups.

Extensive studies on the formation of aldonic acid groups on cellulose [192],

mannan [193], xylan [194] and the corresponding oligosaccharides under various

conditions revealed that arabinonic acid was highly predominant after oxidation

of 4-b-linked mannobiose, mannotriose, and mannotetraose. The stabilization

(and also peeling) reactions of glucomannan and cellulose proceed in a similar

way (Schemes 7.15 and 7.16) [193]. In contrast, mannose end-groups – which

react more slowly than glucose end-groups – are converted to the same reactive

“fructose intermediates” as glucose, and the same aldonic acid end-groups in

about the same proportions have been found from manno-oligosaccharides and

mannan as from cello-oligosaccharides and hydrocellulose [193]. The monosaccharides

glucose, mannose, and xylose degrade much faster under dioxygen pressure

than the reducing end-groups of the corresponding oligosaccharides, the degradation

rates of which are almost the same in dioxygen and nitrogen atmospheres

[193].

The formation of aldonic acid end-groups after dioxygen-alkali treatment of

birch xylan studied by Kolmodin and Samuelson [195] showed that xylonic (5,

Scheme 7.15), lyxonic (6), threonic (8) and glyceric (11, Scheme 7.16) acids were

formed as the major terminal acid residues, and xylosone 2,4-dihydroxy-butyric

acid (17, 18) was also extensively formed in non-oxidative treatments [194]. Lyxonic

and xylonic groups are expected from a benzilic-type rearrangement (BAR) of

pentosulose end-unit (Scheme 7.15), whereas oxidative or hydrolytic cleavage

leads to threonic acid (8). Glyceric acid (11) is probably formed via cleavage of dglycero

2,3-pentodiulose (10) end-units formed by isomerization of pentosulose

units (4), and from alkaline and oxidative degradation of the xylosone (4) through

threose (13) and threosone (14) as intermediates.

7.3.2.5.2 Peeling Reactions Starting from the Reducing End-Groups

The peeling removes the terminal anhydro-sugar unit, generating a new reducing

end-group until a competitive stopping reaction sets in forming a stable saccharinic

acid end-group (see Section 4.2.4.2, Carbohydrate reactions).

In studying the oxidative alkaline peeling reaction of cellulose by using cello-oligosaccharides

and hydrocellulose, Malinen and Sjostrom ([190, 192]) found in

addition to the “normal” alkaline peeling products [isosaccharinic acid (27,

Scheme 7.18) and lactic acid (32)], large amounts of 3,4-dihydroxybutyric acid

(28), glycolic acid (33), 3-deoxy-pentonic acid (17, 18, Scheme 7.16), formic acid

(34) and glyceric acid (35). The formation of the two isomeric glucoisosaccharinic

acids (e.g., 27) by alkaline treatment of cellulose is much depressed in the presence

of dioxygen [185], and the 4-deoxy-d-glycero–2,3-hexodiulose (26) is instead

fragmented to 3,4-dihydroxybutyric acid (28) and glycolic acid (33). These are

formed via oxidative cleavage of 4-deoxy-d-glycero–2,3-hexodiulose (26), which can

660 7Pulp Bleaching

H OH

HO H

H OR

H OH

R1

CH2OH

O

HO H

H OR

H OH

R1

H O

23

R1 = -H for xylan

R = Polysaccharide chain

R1 = -CH2OH for cellulose / glucomannan

23 Glucose end-group

24 Mannose end-group

25 Fructose end-group

26 4-Deoxy-D-glycero-2,3-hexodiulose

27 Isosaccharinic acid

28 3,4-Dihydroxybutyric acid

29 Glycolic acid

30 Dihydroxyacetone and glyceraldehyd

31 Methyl glyoxyl

32 Lactic acid

33 Glycolic acid

34 Formic acid

35 Glyceric acid

23 Xylose end-group

25 Xylulose end-group

26 4-Deoxy-2,3-pentodiulose

27 Xyloisosaccharinic acid

28 2-Deoxy-glyceric acid

29 Glycolic acid

30 Dihydroxyacetone and glycolaldehyd

31 Methyl glyoxyl

32 Lactic acid

33 Glycolic acid

34 Formic acid

35 Glyceric acid

25

- ROH

+ OH-

CH2OH

O

OH

H

H OH

R1

26

CH2OH

O

O

H

H OH

R1

COOH

HO

H

H OH

R1

27

+ OH-

+ O2/OHCH2OH

H

COOH

H

H OH

R1

28

H

+

COOH

CH2OH

29

HO H

HO H

H OH

H OH

CH2OH

H O

24

OH

30

H

H O

R1

CH2OH

O

CH2OH

OH

CH2

COOH

OH

CH2OH

COOH

CH2OH

33

HCOOH

34

+ OH-

O2/OH-

O2/OHH

O

31

+

O

CH3

H O

+ OH-

H

32

COOH

OH

CH2OH

H

35

Scheme 7.18 Peeling reactions of polysaccharides during

alkaline and oxidative alkaline conditions (redrawn from

Ref. [183]).

also rearrange to isosaccharinic acids (27) or cleave to yield glyceraldehyde (30)

[183]. Glyceraldehyde is further converted to lactic (32), glycolic (33) and glyceric

(35) acids.

Malinen and Sjostrom [192] reported that the extent of the peeling reaction for

cello-oligosaccharides was very low and that stabilization proceeded quickly. However,

the stabilization of hydrocellulose – that is, the formation of aldonic acid

end-groups – was less extensive, and peeling resulted in a loss of 10–50 sugar

units, depending on the reaction.

The peeling reactions of xylan and glucomannan that take place under alkaline

conditions have been described in detail (see Section 4.2.4.2, Carbohydrate reactions).

In the presence of dioxygen, the peeling of xylan is more extensive than in

alkali alone, and greater than that of cellulose and glucomannan. However, in the

absence of dioxygen the degradation rate is lower for xylan than for cellulose and

glucomannan [192,193,195]. 2,4-Dihydroxy-butyric acid (17, 18, Scheme 7.16), 2-

deoxy-glyceric acid (28, Scheme 7.18), glycolic acid (33), glyceric acid (35), xyloisosaccharinic

acid (27), lactic acid (32) and formic acid (34) are the main peeling

products of xylan, which are analogous to the peeling products of cellulose.

The xylan chains are partly substituted with 4-O-methyl-glucuronic acid units at

C2 [196], which prevent migration of the carbonyl group to the b-position relative

to the glycosidic bond constraining b-elimination (see Section 4.2.4.2, Carbohydrate

reactions; specific reactions of xylan). Model studies with aldobiuronic acid

7.3 Oxygen Delignification 661

[194,197] revealed that, under alkaline conditions at 80 °C, the degradation rate

was rapid but much slower than that of xylobiose. Under dioxygen alkali conditions,

aldobiuronic acid degraded almost as fast as xylobiose, suggesting that the

substituent at C2 has a low retarding effect on the peeling reaction. The arabinose

substituent at C3 position of softwood xylan is easily cleaved by b-elimination

through the peeling process, and the chain is partly stabilized to xylometasaccharinic

acid end-groups [198].

7.3.2.5.3 Peeling Reactions Starting from Stabilized End-Groups

The formation of aldonic acid end-groups serves as a possible means of stabilizing

the reducing end of the polysaccharide chain. In the presence of dioxygen, arabinonic

acid end-groups (8, Scheme 7.15) are formed that are relatively stable under

typical oxygen bleaching conditions, but degrade rather rapidly above 120 °C

under both oxygen and nitrogen atmospheres (Scheme 7.15). The formed erythronic

acid (11) and gluconic acid (5) end-groups are essentially stable up to 150 °C

[192,199]. Glucitol end-groups, which are more stable against dioxygen-alkali

treatment than the reducing end-groups, are relatively rapidly oxidized at higher

temperatures to arabinose, and are cleaved further by b-elimination [183,199].

Mannitol end-groups are oxidized through the same arabinose intermediates as

the glucitol end-groups. The model-compound methyl-a-d-mannopyranoside was

oxidized more rapidly than methyl-a-d-glucopyranoside giving similar oxidation

products, whereas the yield of furanosidic carboxylic acid was greater for methyla-

d-mannopyranoside. This suggests that the oxidative attack is favored by the cisposition

of the C2 and C3 hydroxyl groups [183]. Furthermore, the threonic acid

end-groups that have been formed during oxidative stabilization of the reducing

end-groups of xylan, show a similar degradation rate to that of arabinonic acid

end-residues.

7.3.2.5.4 Cleavage of the Polysaccharide Chain

Cleavage of the cellulose chain under dioxygen-alkaline conditions has been studied

with simple model compounds such as methyl-4-O-methyl-b-d-glucopyranoside

[200], methyl-b-d-glucopyranoside [201–203] and methyl-b-d-cellobioside

[204]. These compounds represent the inner cellulose units, and result in the formation

of glycolic acid, lactic acid, formic acid, acetic acid and carbon dioxide

[183] and methyl-b-d-glucoside, d-glucose, d-arabinose, d-arabinonic acid, d-erythronic

acid, and d-glyceric acid [204]. Additionally, carboxy-furanosides, methyl-2-

C-carboxy-b-d-pentafuranosides, have been identified as oxidation products of

both glycosides [200] and the corresponding methyl-3-C-carboxy-b-d-pentafuranoside

has also been formed from methyl-b-d-glucopyranoside. The formation of

these furanosidic acids is suggested via benzilic acid rearrangement of a diketo

intermediate [201].

It has been generally suggested that the oxidative peeling of a cellulose chain

proceeds via oxidation of the C2 or C3 hydroxyl group, followed by b-alkoxy-elim-

662 7Pulp Bleaching

ination at C4 [188]. In contrast, the b-elimination is more pronounced when the 4-

hydroxyl-group is substituted (as in cellulose), as is known from model-compound

studies [185,200]. As a result of b-elimination at C1, preceded by oxidation at C2

or C3, the formation of methyl-b-d-glucopyranoside from oxidation of methyl-b-dcellobioside

can be regarded [205]. The acids which clearly result from the oxidative

cleavage of the C1–C2, C2–C3, and C3–C4 linkages have been identified

among the oxidation products [183]. Furthermore, an attack of the C6 hydroxyl

group by a ROS seems very probable [205,206], because methyl-b-d-glucopyranoside

was more rapidly oxidized than methyl-b-d-xylopyranoside [183,206,207] and

methyl-6-deoxy-b-d-glucopyranoside [206]. Because the products formed from

methyl-4-O-methyl-b-d-glucopyranoside under alkaline hydrogen peroxide treatment

corresponded to those from alkaline dioxygen experiments with glycosides,

a common reactive species was inferred [206,208,209].

Cleavage of the xylan chain studied with methyl-b-d-xyloside as a model compound

[207] showed that the oxidation reaction products were similar to those of

methyl-b-d-glucopyranoside, methyl-4-O-methyl-b-d-glucopyranoside and methyla-

d-mannopyranoside suggesting the same mechanism. Although the oxidation

of methyl-b-d-xyloside was slower, the oxidative depolymerization of xylan was

more drastic compared with cellulose, but this may have been due to physical factors

[183,195] such as crystallinity [80,210].

The common reactive oxygen species [206,208,209] noted previously is thought

to be the hydroxyl radical [3,202–204,211]. A possible degradation mechanism for

carbohydrates proposed by Gierer [3] starts with an attack of a hydroxyl radical

(_OH) at the C2 position in the polysaccharide chain (Scheme 7.19), followed by

oxygenation of the resultant carbon-centered radical and elimination of superoxide

anion radical. This leads to the formation of a ketone in the polysaccharide

chain that allows cleavage of the glycosidic linkage by b-elimination (see Section

4.2.4.2, Carbohydrate reactions).

O

O

O

OH

OH

CH2OH

O

O

O

OH

OH

CH2OH

.OH

-H2O -H+

O

O

O

O-

OH

CH2OH

pH > 10 O

O

O

O-

OH

CH2OH

O2

O

O

O

O

OH

CH2OH

-O2

-

H

H

H

H

H H

H H

H

H

H

O2

Fragmentation by β-elimination

("peeling")

Scheme 7.19 Mechanism for oxidative cleavage of carbohydrates

by hydroxyl radicals proposed by Gierer [3].

Guay et al. [204] have examined the proposed mechanism by using computational

methods, which revealed that the step involving elimination of superoxide

7.3 Oxygen Delignification 663

is energetically unfavorable. The highly reactive hydroxyl radical, which has been

generated by using hydrogen peroxide and UV light [Eq. (20)] [204], is capable of

reacting with most organic compounds, typically by hydrogen abstraction [139].

Hydroxyl radicals can react with both hydrogen peroxide and hydroperoxy anions

through Eq. (21) and Eq. (22), producing hydroperoxy radicals and superoxide

anions, respectively [212]. The reaction producing superoxide [Eq. (22)] is significantly

faster than the hydroperoxy radical formation [Eq. (21)] [213]. As shown in

Scheme 7.4, approximately half of the hydrogen peroxide is present as the conjugate

base at a pH of 11.8, and formation of superoxide anions should be more

important. At a lower pH, more hydroxyl radicals will be present to react with the

carbohydrates.

H2O2 _

hm 2_OH _20_

_OH _ H2O2→H2O _ HO_

2 _21_

_OH _ HO_2 →H2O__O_2 _22_

The experiments of Guay et al. with methyl-b-cellobioside have been conducted

with and without hydrogen peroxide at pH 10 and 12, and under oxygen pressure

(about 4 bar) at 90 °C [204]. Beside the predominant degradation products of

methyl-b-glucoside and d-glucose, d-arabinose, d-cellobionic acid, d-arabinonic

acid d-erythronic acid, d-glyceric acid, and glycolic acid, products that have also

been found by other groups [183,192,202,214–219], were identified. Moreover, no

degradation products were found in the control reactions, suggesting that dioxygen,

hydroxide ions, hydrogen peroxide, and hydroperoxy anions are not capable

of degrading carbohydrates without a radical initiator, such as lignin or metal ions

[204]. Due to the lower reactivity of methyl-b-cellobioside at higher pH (12) [202],

and the pH-dependence of the oxygen-species distribution [see Eqs. (21) and (22)],

the extent of the degradation decreased but the overall chemistry was unchanged

[204].

The mechanism of the formation of d-cellobioside is proposed to occur through

a two-step process (Scheme 7.20), starting with a hydroxyl ion attack at the anomeric

carbon displacing the methoxy radical. This radical can then abstract a hydrogen

from hydrogen peroxide or another hydrogen donor, forming a hydroperoxyl

radical and methanol (found experimentally).

The second degradative pathway (Scheme 7.21) is very similar to the first

(Scheme 7.20), except that the cleavage is between two pyranose rings, starting

with a hydroxyl attack at the anomeric carbon displacing d-glucose and methyl bglucoside

oxy radical at C4. The methyl b-glucoside radical then abstracts a hydrogen

from hydrogen peroxide, forming methyl b-d-glucoside [204].

664 7Pulp Bleaching

O

H

O

H

HO

H

H

H OH

OCH3

OH

O

H

HO

H

HO

H

H

H OH

OH

+H2O2

.OH

O

H

O

H

HO

H

OH

OH

H

H

OH

O

H

HO

H

HO

H

H

OH

H

OH

+ CH3O.

HOO+ .

O

H

O

H

HO

H

H

OH

H

OH

OH

O

H

HO

H

HO

H

H

OH

H

OH

CH3OH

Scheme 7.20 Proposed mechanism for cellobiose formation

(redrawn from Guay et al. [204]).

O

H

O

H

HO

H

H

H OH

OCH3

OH

O

H

HO

H

HO

H

H

H OH

OH

.OH

OH

O

H

HO

H

HO

H

H

OH

H

OH

O

H

O

H

HO

H

H

OH

H

OCH3

OH

O

H

HO

H

HO

H

H

OH

H

OCH3

OH

+H2O2

OH

O

H

HO

H

HO

H

H

OH

H

OH

+

Scheme 7.21 Proposed mechanism for formation of methylb-

glucoside and D-glucose (redrawn from Guay et al. [204]).

Guay et al. [204] concluded that their experiments supported the view that hydroxyl

radicals are responsible for the degradation of carbohydrates during oxygen

delignification. Molecular oxygen, hydrogen peroxide, and hydroperoxy anions do

not appear to degrade carbohydrates directly. Previous studies also suggested that

superoxide anions do not degrade carbohydrates [3]. Guay et al. [204] reported that

no experimental evidence has been found to support the reaction mechanism

depicted in Scheme 7.19, though this may be due to different experimental conditions

being used in these studies and in previous research, which employed pulse

radiolysis to generate hydroxyl radicals. Evidence has been published suggesting

that cellulose degradation during pulse radiolysis arises from direct ionization of

the fibers rather than from hydroxyl radicals [216]. Moreover, the mechanism

(cleavage of the glycosidic linkage) shown in Scheme 7.21is supported by the

model-compound study with 1,4-anhydrocellobiotol and cellulose [211].

Details of the mechanisms regarding the involvement of superoxide elimination

[3,130] or no superoxide [204] are discussed – albeit controversially – in the literature,

there appears to be no doubt that the hydroxyl ion attacks the carbohydrates,

thereby starting the degradation reaction [4,220–226].

7.3 Oxygen Delignification 665

The hydroxyl radical (_OH) is one of the most reactive and short-lived of the

ROS, with a lifetime of about 1ns in biological systems [227]. Because of this,

methods used to detect _OH include electron spin resonance (ESR) [228] (using a

spin trap such as dimethylsulfoxide, DMSO), HPLC [229,230], rapid-flow ESR

[231], and fluorescence [232–237]. Two different methods can be used for the

detection of _OH. One is the direct reaction of a probe molecule with .OH. The

other method is to use a scavenger that creates a radical species with a longer lifetime.

The probe molecule then reacts with this radical species [229,234]. Superoxide

detection system have also been developed using ESR spin trapping [238], cytochrome

C [239,240], amperometric detection [241], or a chemiluminescence

assay [242,243], which may help to clarify whether the superoxide anion radical is

formed as a consequence of oxygen treatment. Moreover, a new chromatographic

method to determine hydroperoxides in cellulose [244], and a new colorimetric

method to determine hydroxyl radicals during the aging of cellulose [245] have

been published.

A compilation of important carbohydrate degradation products in dioxygenalkali

delignification processes of kraft and sulfite pulps (glycolic acid, 2,4-dihydroxibutyric

acid, 3,4-dihydroxibutyric acid, isosaccharinic acid, 2-deoxy-glyceric

acid, lactic acid, glyceric acid, formic acid, and acetic acid) according to Sjostrom

and Valttila [246] sums up this section.

7.3.2.6 Residual Lignin–Carbohydrate Complexes (RLCC)

It is well known that lignin and carbohydrates are linked in wood, and that new

linkages are formed during a kraft cook.

During oxygen delignification of pine pulp, the polysaccharides dissolve together

with lignin in the form of lignin–carbohydrate complexes (LCC) [247]. The

structures of these dissolved polysaccharides from pine and birch kraft pulps

treated under oxygen delignification conditions [247], when determined by using

methylation analysis [248], included 1,4-linked xylan, 1,3(,6)-linked and 1,4-linked

galactan, 1,5-linked arabinan, and notable amounts of a 1,3-linked glucan,

whereas the glucose-containing polysaccharide in the pine pulp effluent was 1,3-

linked glucan and not cellulose [247]. From the birch pulp mainly xylan, but also

traces of arabinan, 1,3-linked galactan and 1,4-linked glucan have been removed

[247].

Softwood kraft pulps with a kappa number between 50 and 20 and oxygendelignified

to a similar lignin content (kappa ~6) led to the isolation of LCCs using

a method based on selective enzymatic hydrolysis of the cellulose, and quantitative

fractionation of the LCC [63]. The large majority (85–90%) of the residual lignin

in the unbleached kraft pulp, and all of that in the oxygen-delignified pulps,

when isolated as LCC, was found as one of three types of complex, namely xylan–

lignin, glucomannan–lignin–xylan and glucan–lignin. Most of the lignin was

linked to xylan in high-kappa number pulps, but to glucomannan when the pulping

was extended to a low kappa number. Lawoko et al. [63] reported that, with

increasing degree of oxygen delignification, a similar trend in the delignification

666 7Pulp Bleaching

rates of LCC was observed; thus, the residual lignin was increasingly linked to glucomannan.

From this it was concluded that complex LCC network structures

appear to be degraded into simpler structures during delignification. Two excellent

schemes for the degradation of hemicellulose networks during pulping, and

possible differences in the accessibility of lignin under alkaline conditions between

a xylan–lignin complex and a glucomannan–lignin complex, were

described by Lawoko et al. [63]. Moreover, the chemical structure of the residual

lignin bound to xylan was different from that bound to glucomannan.

Enzymatically isolated residual lignin–carbohydrate complexes (RLCC) from

spruce and pine pulp (kappa number ca. 30) contained 4.9–9.4% carbohydrates,

with an enrichment of galactose and arabinose compared to the original pulp

samples. The main carbohydrate units present in the RLCC were 4-substituted

xylose, 4-, 3- and 3,6-substituted galactose, 4-substituted glucose, while 4- and 4,6-

substituted mannose were assigned to carbohydrate residues of xylan, 1,4- and

1,3/6-linked galactan, cellulose and glucomannan [65]. The comparison of RLCC

of surface material and the inner part of spruce kraft pulp fiber revealed that the

1,4-linked galactan was the major galactan in RLCC of fiber surface material of

spruce kraft pulp, and towards the inner part the proportion of 1,3/6-linked galactan

increased relative to 1,4-linked galactan [65]. It has been suggested that 1,3/6-

linked galactan structures may have a role in restricting lignin removal from the

secondary fiber wall. The RLCC of three different alkaline pine pulps studied by

Lawoko et al. [65] before and after oxygen delignification revealed small differences

in the carbohydrate structures of the unbleached pulps resulting from the

cooking method [conventional kraft pine pulp, a polysulfide/anthraquinone (AQ)

pine pulp and a soda/AQ pine pulp]. These authors found that all RLCC of oxygen-

delignified pulps had more nonreducing ends and less 1,3/ 6-linked galactan

than the corresponding RLCC of the unbleached pulps. Moreover, the oxygendelignified

soda/AQ pulp had a higher ratio of 1,4-galactan to 1,3/6- linked galactan

and shorter xylan residues than the RLCCs of oxygen-delignified conventional

kraft pine pulp and polysulfide/AQ pulps [65]. From the above results and the calculated

degree of polymerization, conclusions were drawn on the possible positions

of lignin–carbohydrate bonds (Fig. 7.27).

These authors concluded that xylan residues were partly bound to lignin via the

reducing end-groups, and that the RLCC contained either long galactan chains or

bonds linking galactans to lignin via the reducing ends [65]. Oxygen delignification

shortened the oligosaccharide chains present in RLCC and removed preferably

the 1,3/6- linked galactan compared to 1,4-linked galactan structures connected

to residual lignin. The RLCC of oxygen-delignified soda/AQ pulp differed

from those of the other two pulps after oxygen- delignification in that it had a

higher ratio of 1,4- to 1,3/6-linked galactan, and shorter xylan residues. However,

even this detailed analysis did not reveal any major differences in the soda/AQ

pulp that could explain its poor bleaching response. It is possible that factors other

than the chemical composition and interactions between lignin and carbohydrates

affect the bleachability of the pulps. These factors may be physical rather than

chemical [65].

7.3 Oxygen Delignification 667

gal-(1 4)-gal-(1 4)-gal-(1 4)-gal-(1 4)-gal-1

n

Cellulose and 1.3-glucan residues

glc-(1 3)-glc-(1 3)-glc-(1 3)-glc-(1 3)-glc-1

1-6

xyl-(1 4)-xyl-(1 4)-xyl-(1 4)-xyl-1

1-3