Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Polymer Degradation and Stability 110 (2014) 184e194

Contents lists available at ScienceDirect

Polymer Degradation and Stability


journal homepage: www.elsevier.com/locate/polydegstab

Review article

Physicochemical characterization of ethanol organosolv lignin (EOL)


from Eucalyptus globulus: Effect of extraction conditions on the
molecular structure
n
Mauricio Ya ~ez-S a, *, Betty Matsuhiro a, Carolina Nun
~ ez a, Shaobo Pan b,
Christopher A. Hubbell , Poulomi Sannigrahi , Arthur J. Ragauskas b
b b

a
Laboratorio de Hidratos de Carbono, Departamento de Química, Universidad de Santiago de Chile (USACH), Departamento de Ciencias del Ambiente,
Facultad de Química y Biología, Universidad de Santiago de Chile, Casilla 40, Correo 33, Chile
b
Institute of Paper Science and Technology (IPST), School of Chemistry and Biochemistry, Georgia Institute of Technology, Atlanta, GA 30332, United States

a r t i c l e i n f o a b s t r a c t

Article history: The aim of the present study was to investigate the effect of extraction conditions, mainly severity factor
Received 7 July 2014 (H-factor ¼ 3980e14,500) on the molecular structure of ethanol organosolv lignins extracted from
Received in revised form Eucalyptus globulus. Isolated lignins were structurally characterized by 1H NMR, 31P NMR, UVeVis, FT-IR
24 August 2014
spectroscopy and gel permeation chromatography. The results showed that an increase in the severity of
Accepted 27 August 2014
Available online 11 September 2014
the pretreatment decreased the molecular weight of the lignins within a 36e56% range with respect to
the untreated lignin (MWL). Moreover, the increase in severity of the organosolv treatment was
accompanied by strong decrease in the content of aliphatic hydroxyl groups and by an increase of
Keywords:
Organosolv
syringyl phenolic units and condensed phenolic structures. The condensed phenolic structures quanti-
Lignin fied in the organosolv lignin correspond to resinol, phenylcoumaran, dibenzodioxocin, spirodienone and
Lignocellulosic b-10 linkages.
Eucalyptus © 2014 Elsevier Ltd. All rights reserved.

1. Introduction radicals that can randomly combine with each other [3]. This
combining of radicals generates different types of inter-unit link-
Lignin is the second most abundant terrestrial organic polymer ages including b-O-aryl ether (b-O-4), 4-O-5 (biphenyl ether) and
on earth after cellulose and plays an essential role in the architec- carboncarbon linkages like 1,2-diaryl propane (b-1), resinol (5-5),
ture of vascular plant cell walls [1]. Lignin contributes to the biphenyl (5-50 ), and phenylcoumaran (b-5/a-O-4) [4] (Fig. 1b).
compression strength of woody stems and to the water proofing of Currently, lignin is mainly obtained from the black liquor of
tracheids cells and vessel elements within the xylem for the con- pulping process kraft hardwood and softwood which represents
duction of water and minerals throughout the plant [2]. 94% of worldwide production of cellulose pulp [5]. However, this
Lignin is a branched and amorphous polyphenolic macromole- lignin is directly used as a boiler combustible for energy generation
cule (Fig. 1a) biosynthesized in the cell wall through copolymeri- which also allows to recover the process reagents and energy [6],
zation of one to three precursors consisting of p-hydroxy-cinnamyl only a small amount of this lignin has been used as raw material for
alcohols (monolignols) called coniferyl alcohol (R1 ¼ OMe, R2 ¼ H), the obtention of aromatic chemicals such as vanillin [7], phenols
sinapyl alcohol (R1 ¼ OMe, R2 ¼ OMe) and p-coumaryl alcohol and methoxyphenols [8] and polymeric materials such as epoxy
(R1 ¼ H, R2 ¼ H). The polymerization process begins with the resins [9], phenolic resins [10], polyolefins [11] and polyurethane
enzymatic dehydrogenation of the phenolic hydroxyl groups from foam [12].
the monolignols, which produced resonance stabilized phenoxy Despite the major advances in the use of lignin, especially kraft
and lignosulphonates [13], there is new interest on the use of
sulfur-free lignins, having always in mind that the renewable
source should be relatively pure and easy to obtain. Some of the
* Corresponding author. Avenida Libertador Bernardo O'Higgins No 3363,
n Central, Santiago, Chile. Tel.: þ56 (0)27181153.
lignins presenting these characteristics are obtained from ligno-
Estacio
E-mail addresses: mauricio.yanez@usach.cl, mauricioqc@gmail.com (M. Y
an~ez- cellulosic biomass pretreatment technologies which are mainly
S). focused on the generation of reactive substrates for the ethanol

http://dx.doi.org/10.1016/j.polymdegradstab.2014.08.026
0141-3910/© 2014 Elsevier Ltd. All rights reserved.
n
M. Ya ~ez-S et al. / Polymer Degradation and Stability 110 (2014) 184e194 185

Fig. 1. (a) Schematic representation of a hardwood lignin structure (b) Common linkages between phenylpropane units in hardwood Lignin.

production [14]. Among the lignocellulosic biomass pretreatments, approximately 2.0 cm  2.5 cm  0.5 cm. The wood chips were
the organosolv processes stands out as one of the most efficient dried in a mechanical convection oven at 50  C to 8% moisture and
methods for obtaining high purity lignin. This extraction systems stored in plastic bags at 4  C. All solvents used were analytical grade
use mixtures of organic solvent-water and under certain reaction from J.T. Baker. The chemicals and deuterated solvents were ob-
conditions has the ability to progressively break down and modify tained from SigmaeAldrich.
the lignin macromolecule until the resulting molecular fragments
become small enough to dissolve in the aqueous liquor [15].
Among the organosolv pretreatments, ethanol organosolv pre- 2.2. Organosolv pretreatments
treatment seems to be the most promising because the main
product of the bioprocess (ethanol) could be recycled and reused in E. globulus wood chips were ethanol-organosolv-pretreated
the first step of pretreatment [16]. By using ethanol organosolv under several conditions. The pretreatment processes were car-
pretreatment, lignins have been isolated and characterized from ried out in a 3.75 L 4551 pressure Parr reactor equipped with a 4843
several raw materials such as Miscanthus giganteus [17], Buddleja PID temperature controller, thermocouple, sampling and safety
davidii [18], wheat straw [19] and softwood as Loblolly Pine [20]; valves, and fitted with internal cooling coil (Parr Instrument
moreover under the conditions applied in such processes it has Company, Moline, IL. USA). The reactor was loaded with 200 g of
been possible to obtain fragments of low molecular weight relative wood chips (dry basis) and the solvent (30e70% v/v) ethanol at a
to native lignin. These depolymerized lignins could be interesting liquor:wood ratio of 6:1 (v/w). The reactor was heated at an average
starting materials for the preparation of high-value chemicals and rate of 2.5  C/min, starting at 20  C, to reach the desired set point
fuels. temperature (190  C or 200  C), which was kept within a range of
Despite a considerable number of works on the structural ±1  C; after a residence time (15e75 min) the reactor was rapidly
characterization of organosolv lignin there are very few studies cooled. The reaction mixture was vacuum filtered and the solid
related to the effect of extraction conditions on the molecular fraction (pretreated material) was washed with 1.2 L of the solvent
structure of lignin. In the present study, ethanol organosolv mixture at the same concentration as the liquor and then, with 1 L
pretreatments were performed on Eucalyptus globulus wood of hot water. The pretreated material was homogenized for 5 min at
under a wide range of severity of the process (H- 250 rpm in a laboratory disintegrator, and screened in a BH-6/12
factor ¼ 3980e14,500), in order to investigate the effect of REGMED fiber classifier with flat screen (0.2 mm  20 mm) to
extraction conditions on the molecular structure of isolated lig- remove the rejects (non-defiberized wood chips and knots). The
nins and also to better understand the lignin breakdown mech- rejects were dried in an oven at 50  C for 2 days and then weighed.
anism in the organosolv process. The screened pretreated material was centrifuged, weighed, and its
moisture content determined according to T 210 cm-03 TAPPI
method. The pretreated material (pulp) samples were analyzed to
2. Materials and methods determine Klason lignin and carbohydrate content, according to
T222 om-88 and T249 cm-85 TAPPI standard methods, respectively.
2.1. Raw material The severity of the process was calculated using the H-factor
parameter [21], which relates time and temperature in a single
Wood chips from 10 to 12 year-old E. globulus trees provided by variable and was calculated according to Sundquist using numerical
the Empresas CMPC S.A. (Santiago, Chile) were screened to integration methods (Eq (1)) Table 1.
186 n
M. Ya ~ez-S et al. / Polymer Degradation and Stability 110 (2014) 184e194

Table 1
Lignin Lignin
Experimental conditions of the ethanol organosolv pretreatments on Eucalyptus
globulus.
OH-
Condition Temperature ( C) Ethanol Time (min) H-factora + H2 O
concentration
(%, v/v) H3CO OCH3 H3CO OCH3
A 200 50 45 9624 OH O-
B 200 50 50 11,759
C 200 50 75 13,659
Fig. 2. Ionization of phenolic hydroxyl groups in lignin.
D 200 70 75 14,508
E 190 50 45 3948
F 190 50 75 6311 2.5.2. UVeVisible spectroscopy
G 190 30 45 3980 The analysis by UVeVis spectroscopy were performed on a Shi-
Z
a
t madzu Spectrophotometer UV-1650 PC spectrophotometer. The
H-factor, H ¼ eð43:2  16117=TÞdt; T: Reaction temperature, t: Reaction
0 determination of phenolic hydroxyl groups in the organosolv lignin
time. isolated samples was made by Ultraviolet difference spectroscopy
and with some modifications [23]. This method is based on the dif-
Zt   ference in absorption at 292 and 370 nm between phenolic units in
16117 neutral and alkaline solutions. The absorbance of EOL solution pH 12
H¼ e 43:2  dt (1)
T (boric acid/sodium borate buffer) was measured against a solution of
0
the same concentration at pH ¼ 6 (KH2PO4/NaOH buffer). The sub-
where T ¼ Reaction temperature, and t ¼ Reaction time. traction of two spectra provided the amount of phenolic hydroxyl
groups present in the sample. The measurements considered a molar
absorption coefficient (Dε) of the hydroxyl group of
2.3. Ethanol organosolv lignin (EOL) 4100 L cm1 mol1, taken as the average of the maximum absorbance
of a set of model compounds of lignin [24] Fig. 2.
The lignin was precipitated from the black liquor at pH 2 (adjusted
with 0.015 M sulfuric acid) [22]. The precipitated lignin was filtered 2.5.3. FT-IR spectroscopy
through a sintered glass Büchner funnel and washed with water at The FT-IR spectra were recorded on a BRUKER, VECTOR 22 IR
70  C to remove residual sugars. The purification was performed by spectrophotometer in KBr pellet.
dissolving 9.00 g of lignin in 45.00 mL of acetone (dried over mo-
lecular sieve) and filtered to remove impurities using a membrane 2.5.4. Determination of carbonyl groups by modified oximation
filter of 0.22 mm. Acetone was evaporated under reduced pressure The carbonyl group content of the organosolv lignin samples
and pure lignin was dried over P2O5 in a vacuum desiccator. was determined by a modified oximation, according to the

½ðOven  dry precipitated ligninðgÞ  100


Lignin recoveryð%Þ ¼
½ðOven  dry untreated woodðgÞ  ðKL=100Þ

where KL is the Klason lignin content of untreated wood (g Klason method described by Zakis et al. [25]. Lignin (80 mg), 2.00 mL of
lignin per 100 g of wood on dry basis). dimethylsulfoxide (DMSO) and the oximating mixture (5.00 mL)
were placed in a tube, and air was expelled using nitrogen. The
2.4. Acetylation of lignin sealed tube was heated at 80  C (±2  C) for 2 h. The reaction
mixture was cooled and then it was quantitatively transferred
Pure lignin (0.50 g) was dissolved in 6 mL of pyridine-acetic into a titration glass using a minimum volume of distilled water.
anhydride mixture (1:1, v/v) and kept in the dark at room tem- The excess of triethanolamine (TEA) was potentiometrically
perature for 72 h, then the solution was added dropwise to 120 mL titrated with 0.1 N HCl to pH 3.3. A blank experiment was per-
of cold water containing1 mL of hydrochloric acid, under constant formed without lignin, under the same conditions as before
stirring. The acetylated lignin precipitate was filtered using a Fig. 3.
membrane filter of 0.22 microns, this was washed with nanopure
water and dried over P2O5 in a vacuum desiccator.

R1 R1
2.5. Chemical characterization of organosolv lignin
O + NH2OH*HCl NOH + H2O + HCl
R2 R2
2.5.1. Elemental analysis
Oxime
The elemental analysis of C, H, N of the purified samples was
R1 = alkyl, aryl.
performed on an Elemental Analyzer EA1108 FISONS Elemental
Analyzer Instruments, the column used was PORAPAKQ column R2 = alkyl or H
and used a thermal conductivity detector were used; the oxygen HCl + (CH2CH2OH)3N (CH2CH2OH)3N*HCl
content was determined by difference. The determination was
performed using a calibration curve using acetanilide as standard. Fig. 3. Determination of carbonyl groups in lignin by modified oximation.
n
M. Ya ~ez-S et al. / Polymer Degradation and Stability 110 (2014) 184e194 187

Table 3
Physicochemical characterization of ethanol organosolv lignin (EOL) extracted from
O
CH2OH Eucalyptus globulus.
P O
HO Condition Lignin MWD Elemental composition
O Lignin O H2 C O
P O recovery (%, w/w)
O O Lignin yielda (%)
O Mw Mn D %C %H %O %N
CDCl3
+ Cl P (g/mol) (g/mol) (Mw/Mn)
R2 R1 O Pyridine
OH R2 A 46.9 3990 1280 3.11 63.79 5.04 30.95 0.22
R1
B 51.2 3490 1240 2.81 63.96 8.10 27.54 0.46
R1=OMe, R2=H O O
P C 63.4 3820 1250 3.06 63.73 7.36 28.45 0.46
R1=R2=OMe
O D 53.4 4150 1350 3.07 63.18 7.46 28.99 0.37
R1=R2=H
E 40.5 4230 1340 3.16 62.33 7.33 29.81 0.52
F 40.5 4270 1360 3.14 64.09 8.16 27.06 0.48
Fig. 4. Derivatization of hydroxyl groups on the lignin polymer with TMDP. G 3.92 2960 1170 2.53 62.12 6.65 30.66 0.56
MWL 6700 2600 2.6 60.80 6.40 31.32
a
% on starting lignin in wood. MWD, molecular-weight distribution, Mn,
2.5.5. Lignin molecular-weight determination molecular-weight distribution number average molecular weight; MW, weight
Lignin sample (20 mg) was placed in a small vial to which average molecular weight, D, polydispersity. MWL: Milled wood lignin.
anhydrous pyridine (0.5 mL) and acetic anhydride (0.5 mL) were
added sequentially. The vial was capped, vortexed and allowed to 210 ppm spectra width in F1 (13C) dimension with 256 data points
stand at RT for 24 h, then ethanol (5 mL) was added to quench the (6.1 ms acquisition time), 3 s pulse delay, 1JCeH of 145 Hz, and 32
reaction. The vial was maintained uncapped overnight in a fume scans. The central solvent (DMSO-d6) peak (dC, 39.5 ppm; dC,
hood and then, then placed in a vacuum oven at 40  C for 3 h to 2.5 ppm) was used for chemical shift calibration. For quantitative
remove any residual volatiles. The acetylated lignin sample was 31
P NMR analysis, organosolv lignin (~25 mg) was dissolved in a
purified by dissolving in chloroform (1 mL) and then precipitated solution of pyridine/CDCl3 (500 mL) and derivatized with 100 mL of
into ethyl ether (75 mL). The resulting solid was collected by 2-chloro-4,4,5,5-tetramethyl-1,3,2-dioxaphospholane (TMDP)
filtration and dried under vacuum at 40  C for 1 h. Prior to GPC (Fig. 4). Endo-N-hydroxy-5-norbornene-2,3-dicarboximide (NHND)
analysis the acetylated lignin sample was dissolved in tetrahydro- was used as internal standard and the chemical shift of phosphi-
furan (1 mg/mL), filtered through a 0.45 mm filter and placed in a tylated NHND occurs at 151.9 ppm. The 31P NMR spectra were ac-
2 mL auto-sampler vial. The molecular weight distributions of the quired using an inverse gated decoupling pulse sequence with a
acetylated lignin samples were analyzed on an Agilent GPC SECu- 25 s pulse delay and 128 scans. NMR data were processed using the
rity 1200 system equipped with four Waters Styragel columns TopSpin 2.1 software (BrukerBioSpin).
(HR0.5, HR2, HR4, HR6), Agilent refractive index (RI) detector and
Agilent UV detector (270 nm) using THF as the mobile phase
3. Results and discussion
(1.0 mL/min) with injection volumes of 20 mL. A calibration curve
was created by fitting a 3rd order polynomial equation to the
3.1. Structural characterization
elution volumes of 16 narrow polystyrene standards ranging in
molecular weight from 580 to 3.6  106 g/mol. Data collection and
E. globulus wood was ethanol-organosolv-pretreated under
processing were performed using Polymer Standards Service
seven conditions as mentioned in the experimental section.
WinGPC Unity software. Number-averaged and weight-averaged
E. globulus has an initial lignin content of ~27% (w/w) oven-dried
molecular weights, Mn and Mw respectively, were calculated by
basis. The chemical composition of organosolv pulp was previ-
the PSS WinGPC software relative to the universal polystyrene
ously reported by Yan ~ ez-S et al. [14], some data are briefly shown
calibration curve.
(see Table 2). As can be seen in Table 3, at 50% ethanol concentra-
tion the increase in severity of the organosolv treatment from 3948
2.5.6. Structural analysis of EOL using 2D NMR and 31P NMR to 13,659 resulted in an increase in the recovery of EOL, reaching a
All NMR experiments were performed on a Bruker Avance-400 maximum of 63% of the Klason lignin in untreated wood. When
spectrometer operating at a frequency of 400.133 MHz for 1H, comparing the conditions C and D a decrease in the recovery of
100.59 MHz for 13C and 161.976 MHz for 31P. Conditions for the 2D lignin was observed, this is because that increasing the ethanol
HSQC NMR analysis were as follows: 13 ppm spectra width in F2 concentration (70%) decreases the concentration of hydronium ion
(1H) dimension with 1024 data points (95.9 ms acquisition time), (H3Oþ) in the solution and thus lowers the depolymerization of
lignin and its solubility in liquor.
Table 2
Chemical composition of pulp and physicochemical characterization of ethanol 3.2. Elemental analysis of EOL
organosolv lignin (EOL) extracted from Eucalyptus globulus.

Condition Chemical composition (% on fibers) The elemental composition of the isolated EOLs of E. globulus
Glucana Xylanb Klason ligninc
from pretreatment 1e7 are summarized in Table 3. The elemental

A 85.3 ± 0.59 3.1 ± 0.04 7.7 ± 0.33


B 80.24 ± 0.25 0.0 ± 0.0 8.73 ± 0.3
C 80.3 ± 0.55 4.1 ± 0.11 6.8 ± 0.02
D 79.9 ± 0.51 10.4 ± 0.03 8.8 ± 0.30
E 80.3 ± 0.21 4.9 ± 0.20 10.4 ± 0.09
F 78.1 ± 0.36 1.0 ± 0.0 9.3 ± 0.16
G 78.1 ± 0.77 4.3 ± 0.38 17.3 ± 0.28
a
Glucan ¼ glucose * 0.9.
b
Xylan ¼ xylose * 0.88, Arabinan  0.03%. Fig. 5. Lignin condensation reaction during autocatalyzed organosolv pretreatment of
c
Klason lignin (includes both acid-soluble and acid-insoluble). lignocellulosic biomass.
188 n
M. Ya ~ez-S et al. / Polymer Degradation and Stability 110 (2014) 184e194

Fig. 6. Variation of the molecular weight (in gmol L1) with the severity of the process (H-factor).

analysis results show percentages of carbon between 62 and 64%


and hydrogen between 5 and 8%. The nitrogen content is between
0.2 and 0.6% which indicates a low level of contamination by pro-
teins and other nitrogenous compounds (Table 3). By comparing
the elemental composition of milled wood lignin (MWL) with EOLs,
we can see that the carbon content increased, while the oxygen
content decreased slightly, indicating that under the organosolv
conditions examined the EOLs undergo condensation reactions.
Hardwood such as E. globulus under severe temperature and
pressure conditions generates acetic acid by hydrolysis of acetyl
groups of xylans, such as O-acetyl-4-O-methyl-glucuronoxylan
[14]. The acetic acid released protonates the hydroxyl group in the
carbon Ca of the side chain of lignin, which is released as a
resonance-stabilized benzyl carbocation. This reactive carbocation
can readily form a bond with an electron-rich carbon atom in the
aromatic ring of another lignin unit i.e., through the free C5 or C6
position [15] (see Fig. 5).

Fig. 7. FT-IR spectrum of ethanol organosolv lignin isolated from E. globulus. Condition 3.3. GPC analysis of EOL
A (H-Factor: 9624).

The number-average molecular weight (Mn), weight-average


molecular weigth (Mw) and polydispersity (Mw/Mn) of milled
wood lignin and organosolv lignin from E. globulus were computed
from their chromatograms. The results are shown in Table 3. Under
the conditions of organosolv pretreatment applied (1e7), the mo-
lecular weight of the lignins decreased within the 36e56% range
with respect to the untreated lignin (MWL, Mw ¼ 6700 g/mol [26]),
indicating that the pretreatments degrade the structure of the
macromolecule to an appreciable degree. A similar effect in EOL
extracted of Miscanthus was reported by Brosse et al. [27], who
observed decreases in molecular weight of 43.5 and 48.5% g/mol for

Table 4
Results of CO group determination obtained by modified
oximation.

Condition H-factor CO (%, w/w)

A 9624 4.95
B 11,759 4.94
C 13,659 ND
D 14,508 3.42
E 3948 3.80
F 6311 3.04
G 3980 3.81
Fig. 8. Ionization difference spectra of organosolv lignins of E. globulus.
n
M. Ya ~ez-S et al. / Polymer Degradation and Stability 110 (2014) 184e194 189

Table 5
Determination of non-conjugated and conjugated phenolic hydroxyl groups (%, w/w) by Ultraviolet difference spectroscopy.

Condition Difference ArOH (1) (mmol/g) ArOH (1) (%, w/w) Difference ArOH) (2) (mmol/g) ArOH(2) (%, w/w)
absortivity (DD) l ¼ 293 nm l ¼ 293 nm absortivity (DD) l ¼ 372 nm l ¼ 372 nm
l ¼ 293 nm (Dε ¼ 4100) l ¼ 372 nm (Dε ¼ 37,250)

A 0.101 1.19 2.02 0.126 0.163 0.27


C 0.103 1.07 1.82 0.120 0.137 0.23
D 0.094 1.09 1.84 0.120 0.153 0.26
E 0.09 0.83 1.41 0.113 0.114 0.195
F 0.143 1.12 1.90 0.2047 0.176 0.30
G 0.09 0.75 1.28 0.113 0.103 0.176

ArOH (1) ¼ Non-conjugated phenolic structures.


ArOH (2) ¼ Conjugated phenolic structures.

Table 6
31
Organosolv lignin structural characteristics calculated from P NMR data.

Chemical shift Assignment Organosolv lignin from E. globulus (mmoles/g of lignin)


range d31P NMR
A B C D E F G

145.5e150.0 Aliphatic (OH) 2.355 1.985 1.915 2.275 2.800 2.577 2.536
151.9 NHND (IS)
140.3e144.7 Condensed phenolic (OH) þ 2.875 2.981 3.002 2.707 2.538 2.756 2.836
Syringyl OH
139.0e140.3 Guaiacyl phenolic (OH) 0.589 0.600 0.579 0.590 0.615 0.581 0.552
138.2e139.0 Catechol (OH) 0.127 0.135 0.122 0.116 0.106 0.107 0.090
137.3e138.2 P-hydroxyphenyl (OH) 0.070 0.072 0.064 0.063 0.048 0.051 0.043
133.6e136.6 Carboxylic acids (OH) 0.243 0.256 0.230 0.184 0.146 0.198 0.175

Mw and Mn respectively. In the same way, studies conducted by linkages. Such behavior is explained better as compared with lig-
Hallac et al. [18], determined that molecular weight (Mw) of milled nins extracted under conditions A to C, which used a lower ethanol
wood lignin of Buddleja davidii decreases by ~85% after ethanol- concentration (50% (v/v)) and thus generating a higher concen-
water pretreatment. tration of hydronium ion in liquor (lower pH value), which largely
Under the studied conditions D, E and F, lignin samples were facilitates the breakdown of a and b-ether bonds in lignin and the
obtained with a low degree of depolymerization in relation to formation of molecular fragments of lower molecular weight.
native lignin (MWL). In the case of condition D, this could be due to Under the reaction condition G were obtained lignin with the
a high concentration of ethanol in the liquor (70% (v/v)), which lowest molecular weight (2960 g/mol), this can be partly explained
generates lower hydrogen ion concentration (higher pH value), due to the low concentration of ethanol (30%, v/v), which generates
decreasing the likelihood that this catalyzes the cleavage of ether a high concentration of hydronium ion in the liquor. However, the

31
Fig. 9. Quantitative P NMR spectra and signal assignment of organosolv lignin A from Eucalyptus globulus.
190 n
M. Ya ~ez-S et al. / Polymer Degradation and Stability 110 (2014) 184e194

3,5

2,5

mmol.g -1of lignin 2 Aliphatic OH


Syringyl OH+ Condensed OH
Guaiacyl OH
1,5
Catechol OH
p-Hydroxyphenyl OH
1 Carboxylic acid

0,5

0
3000 5000 7000 9000 11000 13000

H-factor

Fig. 10. Variation of the hydroxyl group concentrations (in mmol g1 of lignin) with the severity of the process (H-factor).

low concentration of ethanol in the liquor decreases the degree of similar in all EOL isolated, which indicates that the “core” of the
delignification and hence the recovery of lignin in the liquor (3.9%). structure of lignin did not change significantly by changing the
Similar effect was reported by Sadler et al. during catalyzed ethanol reaction conditions. The absorption bands of vibration in the FT-IR
pulping of hybrid poplar wood [28]. Fig. 5 shows the effect of the spectrum of lignin were assigned based on previously reported data
severity of the process on the molecular weight (Mn and Mw) and [29]. The FT-IR spectra for sample of ethanol organosolv lignin,
polydispersity of the organosolv ethanol lignins. At a 50% ethanol condition A is shown in Fig. 7.
concentration, increasing the severity (H-factor) of the process The infrared spectrum of organosolv lignin shows at 3441 cm1
from 6300 to 11,760 decreases the molecular weight of lignin a broad intense OeH stretching absorption due to the phenolic
by~18%. However, further increase in severity up to 13,659 resulted hydroxyl groups and aliphatic hydroxyl groups. The two bands at
in an increase of molecular weight of 330 g/mol, which indicates 2938 and 2840 are assigned to CeH stretching vibrations. The band
that in this range of severity take place repolymerization reactions at 1705 cm1 is caused by stretching of unconjugated ketones,
of lignin. In this range of severity, the average polydispersity for the conjugated aldehydes and carboxylic acids aromatic ring, and the
lignins was 3.1, indicating a broad molecular weight distribution bands at 1605 and 1515 cm1 are caused by aromatic skeletal vi-
(MWL: Mw/Mn ¼ 2.6) Fig. 6. brations, while the band at 1461 cm1 corresponds to asymmetrical
CeH deformations in methyl (eCH3) and methylene (eCH2e)
3.4. FT-IR spectroscopy groups [29]. The band at 1424 cm1 also corresponds to aromatic
skeletal vibrations (C]C) (guaiacyl-syringyl) combined with CeH
The IR absorption spectra of the seven EOL studied were in-plane deformation. The band at 1328 cm1 corresponds to CeO
recorded in the 400e4000 cm1 region. The spectral profiles and stretching (syringyl). The bands at 1217 and 1031 cm1 corresponds
the relative intensities of the absorption bands of vibration are to phenolic hydroxyl groups PheO(H) and PheOeC ether in

Fig. 11. Solvolytic cleavage of b-aryl ether linkage by a mechanism that involves the elimination of formaldehyde.
n
M. Ya ~ez-S et al. / Polymer Degradation and Stability 110 (2014) 184e194 191

Table 7 3.5. Carbonyl groups


Molecular formulas C9 for the organosolv lignins studied.

Condition Formula C9 Molecular Carbonyl groups for all EOL lignins were determined quantita-
weight (g mol1) tively by the modified Oximating method [25]. In all the samples,
A C9H4.88O1.61(CH3O)1.38(OHAr)0.42 195.45 the content of CO is very close to the average of carbonyl in Alcell
(OHAlk)0.27(COOH)0.047 lignin of hardwoods (~5.0 w/w) [25]. However, there is not clear
C C9H8.92O1.13(CH3O)1.69(OHAr)0.44 201.07 relationship between increasing severity of the process and the
(OHAlk)0.23(COOH)0.046
increase of carbonyl groups in the molecular structure of the EOL
D C9H9.98O1.55(CH3O)1.25(OHAr)0.40 194.59
(OHAlk)0.26(COOH)0.036 (Table 4). These carbonyl groups are mainly present in the subunits
E C9H9.44O1.48(CH3O)1.52(OHAr)0.39 202.00 cinnamaldehyde ends and oxidized syringyl units, which were
(OHAlk)0.33(COOH)0.029 determined by 2D 13Ce1H HSQC NMR experiment (see section 3.8).
G C9H7.63O1.40(CH3O)1.78(OHAr)0.43 207.63
(OHAlk)0.31(COOH)0.036

In conditions B and F was not quantified the methoxyl groups. 3.6. UVeVis spectroscopy

syringyl and guaiacyl units, asymmetric and symmetric stretching The phenolic hydroxyl groups of all EOL samples were deter-
vibration respectively. Aliphatic ethers give one strong asymmetric mined by ultraviolet difference spectroscopy. The alkaline-neutral
stretch at 1117 cm1, and a very weak symmetric stretch around difference spectra of organosolv lignin present three absorption
834 cm1. Finally the band at 912 cm1 corresponds to CeH aro- bands at wavelengths of 254, 293 and 372 nm (see Fig. 7). For the
matic (G þ S) out of plane deformation. determination of phenolic hydroxyl groups was used the difference

Fig. 12. A and B. HSQC spectrum of organosolv lignin from Eucalyptus globulus, condition A (a) Expanded side-chain region (dC/dH 50e95/2.5e6.0 ppm) and (b) Expanded aromatic
region (dC/dH 95e160/5.5e8.5 ppm).
192 n
M. Ya ~ez-S et al. / Polymer Degradation and Stability 110 (2014) 184e194

absorptivity (DD) at 293 and 370 nm. These bands indicate the inter-unit linkages (substructures) present in the molecular
presence of non-conjugated and conjugated phenolic hydroxyl structure of lignins obtained. The substructures present in orga-
groups, respectively. As can be seen from Table 4, increasing the nosolv lignin samples were determined based on the presence of
process severity between 3900 and 9600, the number of hydroxyl cross-signals of 13Ce1H correlation. This signal assignment was
groups unconjugated increases and then decreases slightly in the performed based on work previously reported in other hardwoods
range between 9600 and 13,600 Fig. 8(Table 5). as Paulownia fortunei [31] and Eucalyptus grandis [32]. The HSQC
spectrum of organosolv lignin condition A is shown in Fig. 12. The
31
3.7. Quantitative P NMR analysis spectrum shows two regions corresponding to side-chain (dC/dH
50e95/2.5e6.0 ppm) and aromatic 13Ce1H (dC/dH 95e160/
The seven organosolv lignin samples were phosphitylated with 5.5e8.5 ppm) correlations. The main substructures present
2-chloro-4,4,5,5-tetramethyl-1,2,3-dioxaphospholane in presence assigned in the HSQC spectrum are listed in Table 8 and are
of Endo-N-hydroxy-5-norbornene-2,3-dicarboximide (NHND) as depicted in Fig. 13.
an internal standard agent and analyzed with quantitative 31P NMR The presence of strong cross-signals of syringyl units (S) and of
according to a method described by Granata and Argyropoulos [30]. the substructures b-O-40 aryl ether linked to a G and S unit (Aa,b(G)/
The concentrations of the hydroxyl groups were calculated on the Aa,b(S)), confirm the results obtained by 31P NMR, i.e., the greater
basis of the internal standard (i.e., NHND) and the respective in- amount of substructures in the molecular structure of EOL of
tegrated peak area. The signals attributed to functional groups in E. globulus correspond to syringyl phenolic structures. Moreover,
the lignin samples and the analysis of alkyl/aryl substituted hy- condensed phenolic structures quantified by 31P NMR correspond
droxyl groups was determined and summarized in Table 6. to resinol (B), phenylcoumaran (C), dibenzodioxocin (F), spi-
0
The quantitative 31P NMR spectrum of organosolv lignin (con- rodienone (D) and b-1 -linkages (E). However, the cross-signals
dition A) is also shown in Fig. 9. In all samples of organosolv lignin CbeHb and CgeHg for the substructure dibenzodioxocin (F) were
(EOL) examined, it was observed that the phenolic structures that not observed in the spectrum, indicating that these structures are
are present in greater amount corresponding to condensed found in low amounts in hardwoods [31]. Similarly, for the spi-
phenolic moieties and syringyl phenolic units, accounting for about rodienone substructure (D) were not observed the correlations
77e80% of the total phenolic hydroxyl (eOH) content and 41e51% CbeHb and Cb0 eHb0 . The CaeHa and CbeHb correlations for the
of the total hydroxyl content present in the molecule. Moreover, the cinnamyl alcohol end-groups (I) were also not observed. The
aliphatic hydroxyl groups representing about 34e46% of the total
hydroxyl content in the structure of lignins studied.
Fig. 10 shows the variations of the phenolic hydroxyl concen- Table 8
trations (in mmol g1 of lignin) with the H-factor of the organosolv Assignment of 13Ce1H correlation signals in the HSQC spectrum of organosolv lignin
pretreatment. As can be seen, the increase in severity of the orga- of E. globulus condition A.
nosolv pretreatment was accompanied by an increase of syringyl Labels Substructure Assignment dC/dH(ppm)
phenolic units and condensed phenolic structures and a strong
Cb (Phenylcoumaran) (C) Cbe0Hb 53.0/3.49
decrease in the content of aliphatic hydroxyl groups. The decrease Bb (Resinol) (B) CbeHb 53.4/3.07
of aliphatic hydroxyl can be partly explained due to the solvolytic Eb b-10 (E) CbeHb 53.6/2.87
cleavage of b-aryl ether linkage by a mechanism that involves the eOMe Methoxyls CeH 55.7/3.77
loss of the g-methylol group with elimination of formaldehyde. The Ag b-O-40 (A) CgeHg 59.5/3.75 and
3.45
enol ether formed is then hydrolyzed, causing the ether bond Db Spirodienone (D) CbeHb Not observed
cleavage to give guaiacol and 2-(4-hydroxy-3-methoxyphenyl) Ig Cinnamyl alcohol CgeHg 59.5/4.06
acetaldehyde (homovanillin). The guaiacol could be demethylated end-groups (I)
to give catechol (see Fig. 11). The homolytic cleavage of b-O-4 via (sinapyl/coniferyl)
Cy Phenylcoumaran (C) CgeHg 62.76/3.75
quinone methide intermediate causes the formation of b-1 sub-
Aa(G) b-O-40 linked to a G unit (A) CaeHa 71.4/4.80
structures [18]. The b-10 substructures can be degraded in acid Bg Resinol (B) CgeHg 70.9/4.20
medium to give stilbenes. These structures were determined at low Aa(S) b-O-40 linked to a S unit (A) CaeHa 72.1/4.90
concentrations using UVeVis spectroscopy (see section 3.5). Db ′ Spirodienone (D) Cb0 eHb0 Not observed
Hydroxyl groups of units as catechol and p-hydroxyphenyl fol- Da Spirodienone (D) Ca-Ha 77.2/5.13
Fa 5-50 (dibenzodioxocin) (F) Ca-Ha 81.3/4.78
lowed the same trend and remain almost unchanged under all Ab(G) b-O-40 linked to a G unit (A) CbeHb 85.74/4.17
experimental conditions applied. The amounts of carboxylic acid Ba Resinol (B) CaeHa 85.12/4.65
increased from 0.14 to 0.25 (mmol g1 of lignin) by increasing the Fb 5-50 (dibenzodioxocin)(F) CbeHb Not observed
severity of the process in a range between 3900 and 12,000. This Da ’ Spirodienone (D) Ca0 eHa’ 84.9/4.80
Ab(S) b-O-40 linked to a S unit (A) CbeHb 86.61/4.08
increase might be due to the oxidation of aldehydes groups and
Ca Phenylcoumaran (C) CaeHa 87.0/5.47
hydrolysis of ester bonds [27]. S2,6 Etherified syringyl units (S) C2eH2/C6eH6 104.3/6.73
J2,6(S) Sinapaldehyde end-groups (J) C2eH2/C6eH6 106.5/7.06
3.8. Molecular formulas of organosolv lignin S′2,6 Oxidized (Ca]O) phenolic C2eH2/C6eH6 106.7/7.35 and
syringyl units (S′) 7.32
G2 Guaiacyl units (G) C2eH2 110.3/6.94
The molecular formula of the lignin samples was obtained from G5 Guaiacyl units (G) C5eH5 115.3/6.78 and
elemental analysis and the quantitative analysis of hydroxyl groups 6.97
by 31P NMR. Methoxyl content was determined by 1H NMR (Table 7). G6 Guaiacyl units (G) C6eH6 118.8/6.78
Jb Cinnamyl aldehyde CbeHb 126.1/7.00
end-groups (J)
3.9. Structural analysis of E. globulus EOL using qualitative 2D
1 Ib Cinnamyl alcohol CbeHb Not observed
He13C correlation NMR end-groups (I)
Ia Cinnamyl alcohol CaeHa Not observed
All organosolv lignin samples were structurally characterized end-groups (I)
by 2D 1He13C HSQC NMR (heteronuclear single quantum corre- Ja Cinnamyl aldehyde CaeHa 153.5/7.6
end-groups (J)
lation). This technique provided information of the structure of
n
M. Ya ~ez-S et al. / Polymer Degradation and Stability 110 (2014) 184e194 193

Fig. 13. Main substructures identified in ethanol organosolv lignin from Eucalyptus globulus.

aliphatic hydroxyl groups, quantified by 31P NMR belongs to the Acknowledgment


substructures (A), (C), (E) and (G). Finally the HSQC spectra for all
EOLs analyzed present the same signals, so no changes were The authors gratefully thank FONDECYT for financial support
observed in the molecular structure depending on the extraction n de Capital Humano Avanzado en la academia
through Insercio
conditions applied. Project N 79100010 and DICYT Iniciacio
n (USACH) Project codigo
021241YS.

4. Conclusions References

[1] (a) Chen Y, Sarkanen S. Macromolecular replication during lignin biosynthesis.


In this work, we evaluated the effect of process severity on the Phytochemistry 2010;71:453e62.
molecular structure of E. globulus ethanol organosolv lignin by [2] Sarkanen KV, Ludwig CH. Lignin, occurrence, formation, structure and re-
spectroscopic and cromatographic techniques. It can be concluded actions. New York: Wiley/Interscience; 1971. p. 95e240.
[3] Freudenberg K, Neish AC. In: Kleinzeller A, Springer GF, Whittman HG, editors.
from the structural analysis by 31P NMR that increasing the severity Constitution and biosynthesis of lignin. New York: Springer-Verlag; 1968.
factor in the range of 3980e13,600 the phenolic hydroxyl of [4] (a) Ralph J. Lignin structure: recent developments. In: Proceedings of the 6th
syringyl units increases and also increases the content of brazilian symposium chemistry of lignins and other wood components,
Guaratingueta, Brazil; October 1999. p. 97e112;
condensed phenolic units. On the contrary, the content of aliphatic
(b) Lewis NG. A 20th century roller coaster ride: a short account of lignifica-
hydroxyl content decreases. This can be explained by solvolytic tions. Curr Opin Plant Biol 1999;2(2):153e62;
cleavage b-aryl ether linkage by a mechanism that involves the loss (c) Ralph J, Peng J, Lu F. Isochroman structures in lignin: a new b-1 pathway.
of the g-methylol group with elimination of formaldehyde. HSQC Tetrahedron Lett 1998;39(28):4963e4.
[5] World pulp annual historical data e excerpt. 2013.
and FT-IR showed that the severity of process did not significantly [6] (a) Lora JH, Glasser WG. Recent application of lignin: a sustainable alternative
change the core of the organosolv lignin. to nonrenewable materials. J Polym Environ 2002;10:39e48;
194 n
M. Ya ~ez-S et al. / Polymer Degradation and Stability 110 (2014) 184e194

(b) Chum HL, Parker SK, Feinberg DA, Wright JD, Rice PA, Sinclair SA, et al. [19] Xu F, Sun J, Sun R, Fowler P, Baird MS. Comparative study of organosolv lignins
SERI/TR 231e488. 1985. p. 1e86; from wheat straw. Industrial Crops Prod 2006;23:180e93.
(c) Gargulak JD, Lebo SE. ACS symp. Ser. no. 742. 2000. p. 304e20. [20] Sannigrahi P, Ragauskas A, Miller SJ. Lignin structural modifications resulting
[7] Borges da Silva EA, Zabkova M, Araújo JD, Cateto CA, Barreiro MF, from ethanol organosolv treatment of loblolly pine. Energy Fuels 2010;24:
Belgacem MN, et al. An integrated process to produce vanillin and lignin based 683e9.
polyurethanes from kraft lignin. Chem Eng Res Des 2009;87:1276e92. [21] (a) Vroom KE. The ‘‘H’’ factor: a means of expressing cooking times and
[8] (a) Yoshikawa T, Yagi T, Shinohara S, Fukunaga T, Nakasaka Y, Tago T, et al. temperatures as a single variable. Pulp Pap Mag Can 1957;58:228e31;
Production of phenols from lignin via depolymerization and catalytic cracking. (b) Sundquist J. Organosolv pulping. In: Chemical pulping. Helsink: Fapet Oy;
Fuel Process Technol 2013;108:69e75; 1999. p. 411e26.
(b) Yoshikawa T, Shinohara S, Yagi T, Ryumon N, Nakasaka Y. Production of [22] Thrin DT. The measurement of lignin in kraft pulping liquors using an auto-
phenols from lignin-derived slurry liquid using iron oxide catalyst. Appl Catal matic colorimetric method. J Pulp Pap Sci 1988;14:19e22.
B Environ 2014;146:289e97. [23] Lai YZ. Determination of phenolic hydroxyl groups. In: Lin SY, Dence CW,
[9] (a) Cazacu G, Popa VI. Lignin-based blends. In: Vasile C, Kulshreshtha AK, edi- editors. Methods in lignin chemistry. Berlin, Germany: Springer-Verlag; 1992.
tors. Handbook of polymer blends and composites, vol. 4B. Shawbury, UK: Rapra p. 423e34.
Technology Ltd.; 2003. p. 565e614; [24] (a) El Mansouri N, Salvado  J. Analytical methods for determining functional
(b) Kosbar LL, Gelorme J, Japp RM, Fotorny WT. Introducing biobased materials groups in various technical lignins. Industrial Crops Prod 2007;26:116e24;
into the electronics industry. J Ind Ecol 2001;4:93e8; (b) Tejado A, Pen ~ a C, Labidi J, Echeverria JM, Mondragon I. Physico-chemical
(c) Simionescu CI, Rusan V, Macoveanu MM, Cazacu G, Lipsa R, Vasile C, et al. characterization of lignins from different sources for use in phe-
Lignin/epoxy composites. Compos Sci Technol 1993;48:317e23; noleformaldehyde resin synthesis. Bioresour Technol 2007;98:1655e63.
(d) Simionescu CI, Macoveanu MM, Vasile C, Ciobanu F, Esanu M, Ioanid A, et al. [25] Faix O, Andersons B, Zakis G. Determination of carbonyl groups of six round
Polyolefins/lignosulfonates blends. Cell Chem Technol 1996;30(5e6):411e29. robin lignins by modified oximation and FTIR spectroscopy. Holzforschung
[10] (a) Cetin N, Ozmen N. Use of organosolv lignin in phenol-formaldehyde resins 1998;52:268e74.
for particleboard production I. Particleboard production and properties. Int J [26] Guerra A, Filpponen I, Lucia LA, Argyropoulos D. Comparative evaluation of
Adhesion Adhesives 2002;22:477e80; three lignin isolation protocols for various wood species. J Agric Food Chem
(b) Vazquez G, Rodríguez-Bona C, Freire S, Gonzalez-Alvarez J, Antorrena G. 2006;54:9696e705.
Acetosolv pine lignin as copolymer in resins for manufacture of exterior grade [27] El Hage R, Brosse N, Chrusciel L, Sanchez C, Sannigrahi P, Ragauskas A.
plywoods. Bioresour Technol 1999;70:209e14. Characterization of milled wood lignin and ethanol organosolv lignin from
[11] Cazacu G, Mihaies M, Pascu C, Profire L, Kowarskik AL, Vasile C. Polyolefin/ miscanthus. Polym Degrad Stab 2009;94:1632e8.
lignosulfonate blends. Macromol Mater Eng 2004;289:880e9. [28] Pan XJ, Kadla JF, Ehara K, Gilkes N, Saddler JN. Organosolv ethanol lignin from
[12] Kubo S, Kadla JF. Poly(ethylene oxide)/organosolv lignin blends: relationship hybrid poplar as a radical scavenger: relationship between lignin structure,
between thermal properties, chemical structure, and blend behavior. Mac- extraction conditions, and antioxidant activity. J Agric Food Chem 2006;54:
romolecules 2004;37:6904e11. 5806e13.
[13] Zoumpoulakis L, Simitzis J. Ion-exchange resins from PF resin modified lignin. [29] Hussin MH, Rahim AA, Mohamad Ibrahim MN, Brosse N. Physicochemical
Polym Int 2001;50:277e83. characterization of alkaline and ethanol organosolv lignins from oil palm
[14] Y ~ ez-S M, Rojas J, Castro J, Ragauskas AJ, Baeza J, Freer J. Fuel ethanol pro-
an (Elaeis guineensis) fronds as phenol substitutes for green material applica-
duction from Eucalyptus globulus wood by autocatalized organosolv pre- tions. Industrial Crops Prod 2013;49:23e32.
treatment ethanol-water and SSF. J Chem Technol Biotechnol 2013;88:39e48. [30] Granata A, Argyropoulos DS. 2-chloro-4,4,5,5-tetramethyl-1,3,2-dioxaphos-
[15] McDonough TJ. The chemistry of organosolv delignification. Tappi J 1993;76: pholane reagent for the accurate determination of the uncondensed and
186e93. condensed phenolic moieties in lignins. J Agric Food Chem 1995;43:1538e44.
[16] Pan X, Gilkes N, Kadla J, Pye K, Saka S, Gregg D, et al. Bioconversion of hybrid [31] Rencoret J, Gisela Marques, Gutie rrez A, Nieto L, Jime nez-Barbero J,
poplar to ethanol and coproducts using an organosolv fractionation process: Martínez AT, et al. Isolation and structural characterization of the milled-
optimization of process yields. Biotechnol Bioeng 2006;94:851e61. wood lignin from Paulownia fortunei wood. Indus Crop Prod 2009;30:
[17] El Hage R, Brosse N, Sannigrahi P, Ragauskas A. Effects of process severity on 137e43.
the chemical structure of miscanthus ethanol organosolv lignin. Polym Degrad [32] Prinsen P, Gutierrez A, Rencoret J, Nieto L, Jimenez-Barbero J, Burnet A, et al.
Stab 2010;95:997e1003. Morphological characteristics and composition of lipophilic extractives and
[18] Hallac BB, Pu Y, Ragauskas A. Chemical transformations of buddleja davidii lignin in Brazilian woods from different eucalypt hybrids. Indus Crop Prod
lignin during ethanol organosolv pretreatment. Energy Fuels 2010;24: 2012;36:572e83.
2723e32.

You might also like