Lignin

Last updated
Idealized structure of lignin from a softwood Lignin structure.svg
Idealized structure of lignin from a softwood

Lignin is a class of complex organic polymers that form key structural materials in the support tissues of most plants. [1] Lignins are particularly important in the formation of cell walls, especially in wood and bark, because they lend rigidity and do not rot easily. Chemically, lignins are polymers made by cross-linking phenolic precursors. [2]

Contents

History

Lignin was first mentioned in 1813 by the Swiss botanist A. P. de Candolle, who described it as a fibrous, tasteless material, insoluble in water and alcohol but soluble in weak alkaline solutions, and which can be precipitated from solution using acid. [3] He named the substance "lignine", which is derived from the Latin word lignum , [4] meaning wood. It is one of the most abundant organic polymers on Earth, exceeded only by cellulose and chitin. Lignin constitutes 30% of terrestrial non-fossil organic carbon [5] on Earth, and 20 to 35% of the dry mass of wood. [6]

Lignin is present in red algae, which suggest that the common ancestor of plants and red algae also synthesised lignin. This finding also suggests that the original function of lignin was structural as it plays this role in the red alga Calliarthron , where it supports joints between calcified segments. [7]

Composition and structure

The composition of lignin varies from species to species. An example of composition from an aspen [8] sample is 63.4% carbon, 5.9% hydrogen, 0.7% ash (mineral components), and 30% oxygen (by difference), [9] corresponding approximately to the formula (C31H34O11)n.

Lignin is a collection of highly heterogeneous polymers derived from a handful of precursor lignols. Heterogeneity arises from the diversity and degree of crosslinking between these lignols. The lignols that crosslink are of three main types, all derived from phenylpropane: coniferyl alcohol (3-methoxy-4-hydroxyphenylpropane; its radical, G, is sometimes called guaiacyl), sinapyl alcohol (3,5-dimethoxy-4-hydroxyphenylpropane; its radical, S, is sometimes called syringyl), and paracoumaryl alcohol (4-hydroxyphenylpropane; its radical, H, is sometimes called 4-hydroxyphenyl).[ citation needed ]

The relative amounts of the precursor "monomers" (lignols or monolignols) vary according to the plant source. [5] Lignins are typically classified according to their syringyl/guaiacyl (S/G) ratio. Lignin from gymnosperms (softwoods, grasses) is derived from the coniferyl alcohol, which gives rise to G upon pyrolysis. In angiosperms (hardwoods) some of the coniferyl alcohol is converted to S. Thus, lignin in angiosperms has both G and S components. [10] [11]

Lignin's molecular masses exceed 10,000 u. It is hydrophobic as it is rich in aromatic subunits. The degree of polymerisation is difficult to measure, since the material is heterogeneous. Different types of lignin have been described depending on the means of isolation. [12]

The three common monolignols:
.mw-parser-output .hlist dl,.mw-parser-output .hlist ol,.mw-parser-output .hlist ul{margin:0;padding:0}.mw-parser-output .hlist dd,.mw-parser-output .hlist dt,.mw-parser-output .hlist li{margin:0;display:inline}.mw-parser-output .hlist.inline,.mw-parser-output .hlist.inline dl,.mw-parser-output .hlist.inline ol,.mw-parser-output .hlist.inline ul,.mw-parser-output .hlist dl dl,.mw-parser-output .hlist dl ol,.mw-parser-output .hlist dl ul,.mw-parser-output .hlist ol dl,.mw-parser-output .hlist ol ol,.mw-parser-output .hlist ol ul,.mw-parser-output .hlist ul dl,.mw-parser-output .hlist ul ol,.mw-parser-output .hlist ul ul{display:inline}.mw-parser-output .hlist .mw-empty-li{display:none}.mw-parser-output .hlist dt::after{content:": "}.mw-parser-output .hlist dd::after,.mw-parser-output .hlist li::after{content:" * ";font-weight:bold}.mw-parser-output .hlist dd:last-child::after,.mw-parser-output .hlist dt:last-child::after,.mw-parser-output .hlist li:last-child::after{content:none}.mw-parser-output .hlist dd dd:first-child::before,.mw-parser-output .hlist dd dt:first-child::before,.mw-parser-output .hlist dd li:first-child::before,.mw-parser-output .hlist dt dd:first-child::before,.mw-parser-output .hlist dt dt:first-child::before,.mw-parser-output .hlist dt li:first-child::before,.mw-parser-output .hlist li dd:first-child::before,.mw-parser-output .hlist li dt:first-child::before,.mw-parser-output .hlist li li:first-child::before{content:" (";font-weight:normal}.mw-parser-output .hlist dd dd:last-child::after,.mw-parser-output .hlist dd dt:last-child::after,.mw-parser-output .hlist dd li:last-child::after,.mw-parser-output .hlist dt dd:last-child::after,.mw-parser-output .hlist dt dt:last-child::after,.mw-parser-output .hlist dt li:last-child::after,.mw-parser-output .hlist li dd:last-child::after,.mw-parser-output .hlist li dt:last-child::after,.mw-parser-output .hlist li li:last-child::after{content:")";font-weight:normal}.mw-parser-output .hlist ol{counter-reset:listitem}.mw-parser-output .hlist ol>li{counter-increment:listitem}.mw-parser-output .hlist ol>li::before{content:" "counter(listitem)"\a0 "}.mw-parser-output .hlist dd ol>li:first-child::before,.mw-parser-output .hlist dt ol>li:first-child::before,.mw-parser-output .hlist li ol>li:first-child::before{content:" ("counter(listitem)"\a0 "}
paracoumaryl alcohol, H
coniferyl alcohol, G
sinapyl alcohol, S Monolignols general.svg
The three common monolignols:

Many grasses have mostly G, while some palms have mainly S. [13] All lignins contain small amounts of incomplete or modified monolignols, and other monomers are prominent in non-woody plants. [14]

Biological function

Lignin fills the spaces in the cell wall between cellulose, hemicellulose, and pectin components, especially in vascular and support tissues: xylem tracheids, vessel elements and sclereid cells.[ citation needed ]

Lignin plays a crucial part in conducting water and aqueous nutrients in plant stems. The polysaccharide components of plant cell walls are highly hydrophilic and thus permeable to water, whereas lignin is more hydrophobic. The crosslinking of polysaccharides by lignin is an obstacle for water absorption to the cell wall. Thus, lignin makes it possible for the plant's vascular tissue to conduct water efficiently. [15] Lignin is present in all vascular plants, but not in bryophytes, supporting the idea that the original function of lignin was restricted to water transport.

It is covalently linked to hemicellulose and therefore cross-links different plant polysaccharides, conferring mechanical strength to the cell wall and by extension the plant as a whole. [16] Its most commonly noted function is the support through strengthening of wood (mainly composed of xylem cells and lignified sclerenchyma fibres) in vascular plants. [17] [18] [19]

Finally, lignin also confers disease resistance by accumulating at the site of pathogen infiltration, making the plant cell less accessible to cell wall degradation. [20]

Economic significance

Pulp mill at Blankenstein, Germany. In such mills, using the kraft or the sulfite process, lignin is removed from lignocellulose to yield pulp for papermaking. Zellstoffwerk Rosenthal 4.jpg
Pulp mill at Blankenstein, Germany. In such mills, using the kraft or the sulfite process, lignin is removed from lignocellulose to yield pulp for papermaking.

Global commercial production of lignin is a consequence of papermaking. In 1988, more than 220 million tons of paper were produced worldwide. [21] Much of this paper was delignified; lignin comprises about 1/3 of the mass of lignocellulose, the precursor to paper. Lignin is an impediment to papermaking as it is colored, it yellows in air, and its presence weakens the paper. Once separated from the cellulose, it is burned as fuel. Only a fraction is used in a wide range of low volume applications where the form but not the quality is important. [22]

Mechanical, or high-yield pulp, which is used to make newsprint, still contains most of the lignin originally present in the wood. This lignin is responsible for newsprint's yellowing with age. [4] High quality paper requires the removal of lignin from the pulp. These delignification processes are core technologies of the papermaking industry as well as the source of significant environmental concerns.[ citation needed ]

In sulfite pulping, lignin is removed from wood pulp as lignosulfonates, for which many applications have been proposed. [23] They are used as dispersants, humectants, emulsion stabilizers, and sequestrants (water treatment). [24] Lignosulfonate was also the first family of water reducers or superplasticizers to be added in the 1930s as admixture to fresh concrete in order to decrease the water-to-cement (w/c) ratio, the main parameter controlling the concrete porosity, and thus its mechanical strength, its diffusivity and its hydraulic conductivity, all parameters essential for its durability. It has application in environmentally sustainable dust suppression agent for roads. Also, lignin can be used in making biodegradable plastic along with cellulose as an alternative to hydrocarbon-made plastics if lignin extraction is achieved through a more environmentally viable process than generic plastic manufacturing. [25]

Lignin removed by the kraft process is usually burned for its fuel value, providing energy to power the paper mill. Two commercial processes exist to remove lignin from black liquor for higher value uses: LignoBoost (Sweden) and LignoForce (Canada). Higher quality lignin presents the potential to become a renewable source of aromatic compounds for the chemical industry, with an addressable market of more than $130bn. [26]

Given that it is the most prevalent biopolymer after cellulose, lignin has been investigated as a feedstock for biofuel production and can become a crucial plant extract in the development of a new class of biofuels. [27] [28]

Biosynthesis

Lignin biosynthesis begins in the cytosol with the synthesis of glycosylated monolignols from the amino acid phenylalanine. These first reactions are shared with the phenylpropanoid pathway. The attached glucose renders them water-soluble and less toxic. Once transported through the cell membrane to the apoplast, the glucose is removed, and the polymerisation commences. [29] Much about its anabolism is not understood even after more than a century of study. [5]

Polymerisation of coniferyl alcohol to lignin. The reaction has two alternative routes catalysed by two different oxidative enzymes, peroxidases or oxidases. LigninPolymerisation.png
Polymerisation of coniferyl alcohol to lignin. The reaction has two alternative routes catalysed by two different oxidative enzymes, peroxidases or oxidases.

The polymerisation step, that is a radical-radical coupling, is catalysed by oxidative enzymes. Both peroxidase and laccase enzymes are present in the plant cell walls, and it is not known whether one or both of these groups participates in the polymerisation. Low molecular weight oxidants might also be involved. The oxidative enzyme catalyses the formation of monolignol radicals. These radicals are often said to undergo uncatalyzed coupling to form the lignin polymer. [30] An alternative theory invokes an unspecified biological control. [1]

Biodegradation

In contrast to other bio-polymers (e.g. proteins, DNA, and even cellulose), lignin resists degradation. It is immune to both acid- and base-catalyzed hydrolysis. The degradability varies with species and plant tissue type. For example, syringyl (S) lignin is more susceptible to degradation by fungal decay as it has fewer aryl-aryl bonds and a lower redox potential than guaiacyl units. [31] [32] Because it is cross-linked with the other cell wall components, lignin minimizes the accessibility of cellulose and hemicellulose to microbial enzymes, leading to a reduced digestibility of biomass. [15]

Some ligninolytic enzymes include heme peroxidases such as lignin peroxidases, manganese peroxidases, versatile peroxidases, and dye-decolourizing peroxidases as well as copper-based laccases. Lignin peroxidases oxidize non-phenolic lignin, whereas manganese peroxidases only oxidize the phenolic structures. Dye-decolorizing peroxidases, or DyPs, exhibit catalytic activity on a wide range of lignin model compounds, but their in vivo substrate is unknown. In general, laccases oxidize phenolic substrates but some fungal laccases have been shown to oxidize non-phenolic substrates in the presence of synthetic redox mediators. [33] [34]

Lignin degradation by fungi

Well-studied ligninolytic enzymes are found in Phanerochaete chrysosporium [35] and other white rot fungi. Some white rot fungi, such as Ceriporiopsis subvermispora, can degrade the lignin in lignocellulose, but others lack this ability. Most fungal lignin degradation involves secreted peroxidases. Many fungal laccases are also secreted, which facilitate degradation of phenolic lignin-derived compounds, although several intracellular fungal laccases have also been described. An important aspect of fungal lignin degradation is the activity of accessory enzymes to produce the H2O2 required for the function of lignin peroxidase and other heme peroxidases. [33]

Lignin degradation by bacteria

Bacteria lack most of the enzymes employed by fungi to degrade lignin, and lignin derivatives (aliphatic acids, furans, and solubilized phenolics) inhibit the growth of bacteria. [36] Yet, bacterial degradation can be quite extensive, [37] especially in aquatic systems such as lakes, rivers, and streams, where inputs of terrestrial material (e.g. leaf litter) can enter waterways. The ligninolytic activity of bacteria has not been studied extensively even though it was first described in 1930. Many bacterial DyPs have been characterized. Bacteria do not express any of the plant-type peroxidases (lignin peroxidase, Mn peroxidase, or versatile peroxidases), but three of the four classes of DyP are only found in bacteria. In contrast to fungi, most bacterial enzymes involved in lignin degradation are intracellular, including two classes of DyP and most bacterial laccases. [34]

In the environment, lignin can be degraded either biotically via bacteria or abiotically via photochemical alteration, and oftentimes the latter assists in the former. [38] In addition to the presence or absence of light, several of environmental factors affect the biodegradability of lignin, including bacterial community composition, mineral associations, and redox state. [39] [40]

Pyrolysis

Pyrolysis of lignin during the combustion of wood or charcoal production yields a range of products, of which the most characteristic ones are methoxy-substituted phenols. Of those, the most important are guaiacol and syringol and their derivatives. Their presence can be used to trace a smoke source to a wood fire. In cooking, lignin in the form of hardwood is an important source of these two compounds, which impart the characteristic aroma and taste to smoked foods such as barbecue. The main flavor compounds of smoked ham are guaiacol, and its 4-, 5-, and 6-methyl derivatives as well as 2,6-dimethylphenol. These compounds are produced by thermal breakdown of lignin in the wood used in the smokehouse. [41]

Chemical analysis

The conventional method for lignin quantitation in the pulp industry is the Klason lignin and acid-soluble lignin test, which is standardized procedures. The cellulose is digested thermally in the presence of acid. The residue is termed Klason lignin. Acid-soluble lignin (ASL) is quantified by the intensity of its Ultraviolet spectroscopy. The carbohydrate composition may be also analyzed from the Klason liquors, although there may be sugar breakdown products (furfural and 5-hydroxymethylfurfural). [42]

A solution of hydrochloric acid and phloroglucinol is used for the detection of lignin (Wiesner test). A brilliant red color develops, owing to the presence of coniferaldehyde groups in the lignin. [43]

Thioglycolysis is an analytical technique for lignin quantitation. [44] Lignin structure can also be studied by computational simulation. [45]

Thermochemolysis (chemical break down of a substance under vacuum and at high temperature) with tetramethylammonium hydroxide (TMAH) or cupric oxide [46] has also been used to characterize lignins. The ratio of syringyl lignol (S) to vanillyl lignol (V) and cinnamyl lignol (C) to vanillyl lignol (V) is variable based on plant type and can therefore be used to trace plant sources in aquatic systems (woody vs. non-woody and angiosperm vs. gymnosperm). [47] Ratios of carboxylic acid (Ad) to aldehyde (Al) forms of the lignols (Ad/Al) reveal diagenetic information, with higher ratios indicating a more highly degraded material. [31] [32] Increases in the (Ad/Al) value indicate an oxidative cleavage reaction has occurred on the alkyl lignin side chain which has been shown to be a step in the decay of wood by many white-rot and some soft rot fungi. [31] [32] [48] [49] [50]

Lignin and its models have been well examined by 1H and 13C NMR spectroscopy. Owing to the structural complexity of lignins, the spectra are poorly resolved and quantitation is challenging. [51]

Related Research Articles

<span class="mw-page-title-main">Cellulose</span> Polymer of glucose and structural component of cell wall of plants and green algae

Cellulose is an organic compound with the formula (C
6
H
10
O
5
)
n
, a polysaccharide consisting of a linear chain of several hundred to many thousands of β(1→4) linked D-glucose units. Cellulose is an important structural component of the primary cell wall of green plants, many forms of algae and the oomycetes. Some species of bacteria secrete it to form biofilms. Cellulose is the most abundant organic polymer on Earth. The cellulose content of cotton fiber is 90%, that of wood is 40–50%, and that of dried hemp is approximately 57%.

<span class="mw-page-title-main">Caffeic acid</span> Chemical compound

Caffeic acid is an organic compound with the formula (HO)2C6H3CH=CHCO2H. It is a yellow solid. Structurally, it is classified as a hydroxycinnamic acid. The molecule consists of both phenolic and acrylic functional groups. It is found in all plants as an intermediate in the biosynthesis of lignin, one of the principal components of biomass and its residues. It is unrelated to caffeine,

<span class="mw-page-title-main">ABTS</span> Chemical compound

In biochemistry, ABTS is a chemical compound used to observe the reaction kinetics of specific enzymes. A common use for it is in the enzyme-linked immunosorbent assay (ELISA) to detect the binding of molecules to each other.

Guaiacol is an organic compound with the formula C6H4(OH)(OCH3). It is a phenolic compound containing a methoxy functional group. Guaiacol appears as a viscous colorless oil, although aged or impure samples are often yellowish. It occurs widely in nature and is a common product of the pyrolysis of wood.

<span class="mw-page-title-main">Dirigent protein</span>

Dirigent proteins are members of a class of proteins which dictate the stereochemistry of a compound synthesized by other enzymes. The first dirigent protein was discovered in Forsythia intermedia. This protein has been found to direct the stereoselective biosynthesis of (+)-pinoresinol from coniferyl alcohol monomers:

<span class="mw-page-title-main">Phenylpropanoid</span>

The phenylpropanoids are a diverse family of organic compounds that are biosynthesized by plants from the amino acids phenylalanine and tyrosine in the shikimic acid pathway. Their name is derived from the six-carbon, aromatic phenyl group and the three-carbon propene tail of coumaric acid, which is the central intermediate in phenylpropanoid biosynthesis. From 4-coumaroyl-CoA emanates the biosynthesis of myriad natural products including lignols, flavonoids, isoflavonoids, coumarins, aurones, stilbenes, catechin, and phenylpropanoids. The coumaroyl component is produced from cinnamic acid.

<span class="mw-page-title-main">Lignocellulosic biomass</span> Plant dry matter

Lignocellulose refers to plant dry matter (biomass), so called lignocellulosic biomass. It is the most abundantly available raw material on the Earth for the production of biofuels. It is composed of two kinds of carbohydrate polymers, cellulose and hemicellulose, and an aromatic-rich polymer called lignin. Any biomass rich in cellulose, hemicelluloses, and lignin are commonly referred to as lignocellulosic biomass. Each component has a distinct chemical behavior. Being a composite of three very different components makes the processing of lignocellulose challenging. The evolved resistance to degradation or even separation is referred to as recalcitrance. Overcoming this recalcitrance to produce useful, high value products requires a combination of heat, chemicals, enzymes, and microorganisms. These carbohydrate-containing polymers contain different sugar monomers and they are covalently bound to lignin.

<span class="mw-page-title-main">Syringol</span> Chemical compound

Syringol is the organic compound with the formula HO(CH3O)2C6H3. The molecule is a phenol, with methoxy groups in the flanking (2 and 6) positions. It is the symmetrically dimethylated derivative of pyrogallol. It is a colorless solid, although typical samples are brown owing to air-oxidized impurities. Together with guaiacol, syringol and its derivatives are produced by the pyrolysis of lignin. Specifically, syringol is derived from the thermal decomposition of the sinapyl alcohol component. As such, syringol is an important component of wood smoke.

<span class="mw-page-title-main">Monolignol</span>

Monolignols, also called lignols, are the source materials for biosynthesis of both lignans and lignin and consist mainly of paracoumaryl alcohol (H), coniferyl alcohol (G) and sinapyl alcohol (S). These monolignols differ in their degree of methoxilation of the aromatic ring.

Laccases are multicopper oxidases found in plants, fungi, and bacteria. Laccases oxidize a variety of phenolic substrates, performing one-electron oxidations, leading to crosslinking. For example, laccases play a role in the formation of lignin by promoting the oxidative coupling of monolignols, a family of naturally occurring phenols. Other laccases, such as those produced by the fungus Pleurotus ostreatus, play a role in the degradation of lignin, and can therefore be classed as lignin-modifying enzymes. Other laccases produced by fungi can facilitate the biosynthesis of melanin pigments. Laccases catalyze ring cleavage of aromatic compounds.

Lignin-modifying enzymes (LMEs) are various types of enzymes produced by fungi and bacteria that catalyze the breakdown of lignin, a biopolymer commonly found in the cell walls of plants. The terms ligninases and lignases are older names for the same class, but the name "lignin-modifying enzymes" is now preferred, given that these enzymes are not hydrolytic but rather oxidative by their enzymatic mechanisms. LMEs include peroxidases, such as lignin peroxidase, manganese peroxidase, versatile peroxidase, and many phenoloxidases of the laccase type.

<span class="mw-page-title-main">Wood-decay fungus</span> Any species of fungus that digests moist wood, causing it to rot

A wood-decay or xylophagous fungus is any species of fungus that digests moist wood, causing it to rot. Some species of wood-decay fungi attack dead wood, such as brown rot, and some, such as Armillaria, are parasitic and colonize living trees. Excessive moisture above the fibre saturation point in wood is required for fungal colonization and proliferation. In nature, this process causes the breakdown of complex molecules and leads to the return of nutrients to the soil. Wood-decay fungi consume wood in various ways; for example, some attack the carbohydrates in wood, and some others decay lignin. The rate of decay of wooden materials in various climates can be estimated by empirical models.

<i>Coriolopsis gallica</i> Species of fungus

Coriolopsis gallica is a fungus found growing on decaying wood. It is not associated with any plant disease, therefore it is not considered pathogenic. For various Coriolopsis gallica strains isolated, it has been found, as a common feature of the division Basidiomycota, that they are able to degrade wood components, mainly lignin and to lesser extent cellulose, which results in a degradation area covered by the accumulating -white- cellulose powder. Therefore, C. gallica might generically be called, as with many other basidiomycetes, a "white-rot" fungus.

In enzymology, a manganese peroxidase (EC 1.11.1.13) is an enzyme that catalyzes the chemical reaction

<span class="mw-page-title-main">Sinapaldehyde</span> Chemical compound

Sinapaldehyde is an organic compound with the formula HO(CH3O)2C6H2CH=CHCHO. It is a derivative of cinnamaldehyde, featuring one hydroxy group and two methoxy groups as substituents. It is an intermediate in the formation of sinapyl alcohol, a lignol that is a major precursor to lignin.

<span class="mw-page-title-main">Coniferyl aldehyde</span> Chemical compound

Coniferyl aldehyde is an organic compound with the formula HO(CH3O)C6H3CH=CHCHO. It is a derivative of cinnamaldehyde, featuring 4-hydroxy and 3-methoxy substituents. It is a major precursor to lignin.

Haem peroxidases (or heme peroxidases) are haem-containing enzymes that use hydrogen peroxide as the electron acceptor to catalyse a number of oxidative reactions. Most haem peroxidases follow the reaction scheme:

Versatile peroxidase (EC 1.11.1.16, VP, hybrid peroxidase, polyvalent peroxidase) is an enzyme with systematic name reactive-black-5:hydrogen-peroxide oxidoreductase. This enzyme catalyses the following chemical reaction

Myceliophthora thermophila is an ascomycete fungus that grows optimally at 45–50 °C (113–122 °F). It efficiently degrades cellulose and is of interest in the production of biofuels. The genome has recently been sequenced, revealing the full range of enzymes used by this organism for the degradation of plant cell wall material.

<span class="mw-page-title-main">Fungal extracellular enzyme activity</span> Enzymes produced by fungi and secreted outside their cells

Extracellular enzymes or exoenzymes are synthesized inside the cell and then secreted outside the cell, where their function is to break down complex macromolecules into smaller units to be taken up by the cell for growth and assimilation. These enzymes degrade complex organic matter such as cellulose and hemicellulose into simple sugars that enzyme-producing organisms use as a source of carbon, energy, and nutrients. Grouped as hydrolases, lyases, oxidoreductases and transferases, these extracellular enzymes control soil enzyme activity through efficient degradation of biopolymers.

References

  1. 1 2 Saake, Bodo; Lehnen, Ralph (2007). Ullmann's Encyclopedia of Industrial Chemistry . Weinheim: Wiley-VCH. doi:10.1002/14356007.a15_305.pub3. ISBN   978-3527306732.
  2. Lebo, Stuart E. Jr.; Gargulak, Jerry D.; McNally, Timothy J. (2001). "Lignin". Kirk-Othmer Encyclopedia of Chemical Technology. Kirk‑Othmer Encyclopedia of Chemical Technology. John Wiley & Sons, Inc. doi:10.1002/0471238961.12090714120914.a01.pub2. ISBN   978-0-471-23896-6 . Retrieved 2007-10-14.
  3. de Candolle, M.A.P. (1813). Theorie Elementaire de la Botanique ou Exposition des Principes de la Classification Naturelle et de l'Art de Decrire et d'Etudier les Vegetaux. Paris: Deterville. See p. 417.
  4. 1 2 E. Sjöström (1993). Wood Chemistry: Fundamentals and Applications. Academic Press. ISBN   978-0-12-647480-0.
  5. 1 2 3 W. Boerjan; J. Ralph; M. Baucher (June 2003). "Lignin biosynthesis". Annu. Rev. Plant Biol. 54 (1): 519–549. doi:10.1146/annurev.arplant.54.031902.134938. PMID   14503002.
  6. "Lignin". Encyclopedia Brittanica. 2023-10-05. Retrieved 2023-10-26.
  7. Martone, Pt; Estevez, Jm; Lu, F; Ruel, K; Denny, Mw; Somerville, C; Ralph, J (Jan 2009). "Discovery of Lignin in Seaweed Reveals Convergent Evolution of Cell-Wall Architecture". Current Biology. 19 (2): 169–75. doi: 10.1016/j.cub.2008.12.031 . ISSN   0960-9822. PMID   19167225. S2CID   17409200.
  8. In the referenced article, the species of aspen is not specified, only that it was from Canada.
  9. Hsiang-Hui King; Peter R. Solomon; Eitan Avni; Robert W. Coughlin (Fall 1983). "Modeling Tar Composition in Lignin Pyrolysis" (PDF). Symposium on Mathematical Modeling of Biomass Pyrolysis Phenomena, Washington, D.C., 1983. p. 1. Archived from the original (PDF) on 2017-08-08. Retrieved 2024-01-29.
  10. Letourneau, Dane R.; Volmer, Dietrich A. (2021-07-22). "Mass spectrometry‐based methods for the advanced characterization and structural analysis of lignin: A review". Mass Spectrometry Reviews. 42 (1): 144–188. doi: 10.1002/mas.21716 . ISSN   0277-7037. PMID   34293221. S2CID   236200196.
  11. Li, Laigeng; Cheng, Xiao Fei; Leshkevich, Jacqueline; Umezawa, Toshiaki; Harding, Scott A.; Chiang, Vincent L. (2001). "The Last Step of Syringyl Monolignol Biosynthesis in Angiosperms is Regulated by a Novel Gene Encoding Sinapyl Alcohol Dehydrogenase". The Plant Cell. 13 (7): 1567–1586. doi:10.1105/tpc.010111. PMC   139549 . PMID   11449052.
  12. "Lignin and its Properties: Glossary of Lignin Nomenclature". Dialogue/Newsletters Volume 9, Number 1. Lignin Institute. July 2001. Archived from the original on 2007-10-09. Retrieved 2007-10-14.
  13. Kuroda K, Ozawa T, Ueno T (April 2001). "Characterization of sago palm (Metroxylon sagu) lignin by analytical pyrolysis". J Agric Food Chem. 49 (4): 1840–7. doi:10.1021/jf001126i. PMID   11308334. S2CID   27962271.
  14. J. Ralph; et al. (2001). "Elucidation of new structures in lignins of CAD- and COMT-deficient plants by NMR". Phytochemistry. 57 (6): 993–1003. doi:10.1016/S0031-9422(01)00109-1. PMID   11423146. Archived from the original on 2021-04-28. Retrieved 2018-12-29.
  15. 1 2 K.V. Sarkanen & C.H. Ludwig, eds. (1971). Lignins: Occurrence, Formation, Structure, and Reactions. New York: Wiley Intersci.
  16. Chabannes, M.; et al. (2001). "In situ analysis of lignins in transgenic tobacco reveals a differential impact of individual transformations on the spatial patterns of lignin deposition at the cellular and subcellular levels". Plant J. 28 (3): 271–282. doi: 10.1046/j.1365-313X.2001.01159.x . PMID   11722770.
  17. Arms, Karen; Camp, Pamela S. (1995). Biology. Saunders College Pub. ISBN   978-0030500039.
  18. Esau, Katharine (1977). Anatomy of Seed Plants. Wiley. ISBN   978-0471245209.
  19. Wardrop; The (1969). "Eryngium sp.;". Aust. J. Bot. 17 (2): 229–240. doi:10.1071/bt9690229.
  20. Bhuiyan, Nazmul H; Selvaraj, Gopalan; Wei, Yangdou; King, John (February 2009). "Role of lignification in plant defense". Plant Signaling & Behavior. 4 (2): 158–159. doi:10.4161/psb.4.2.7688. ISSN   1559-2316. PMC   2637510 . PMID   19649200.
  21. Rudolf Patt etal. (2005). "Pulp". Paper and Pulp. Ullmann's Encyclopedia of Industrial Chemistry. Weinheim: Wiley-VCH. pp. 1–92. doi:10.1002/14356007.a18_545.pub4. ISBN   978-3527306732.
  22. Higson, A; Smith, C (25 May 2011). "NNFCC Renewable Chemicals Factsheet: Lignin". Archived from the original on 20 July 2011.
  23. "Uses of lignin from sulfite pulping". Archived from the original on 2007-10-09. Retrieved 2007-09-10.
  24. Barbara A. Tokay (2000). "Biomass Chemicals". Ullmann's Encyclopedia Of Industrial Chemistry. doi:10.1002/14356007.a04_099. ISBN   978-3527306732.
  25. Patt, Rudolf; Kordsachia, Othar; Süttinger, Richard; Ohtani, Yoshito; Hoesch, Jochen F.; Ehrler, Peter; Eichinger, Rudolf; Holik, Herbert; Hamm (2000). Ullmann's Encyclopedia of Industrial Chemistry . Weinheim: Wiley-VCH. doi:10.1002/14356007.a18_545. ISBN   978-3527306732.
  26. "Frost & Sullivan: Full Speed Ahead for the Lignin Market with High-Value Opportunities as early as 2017".
  27. Folkedahl, Bruce (2016), "Cellulosic ethanol: what to do with the lignin", Biomass, retrieved 2016-08-10.
  28. Abengoa (2016-04-21), The importance of lignin for ethanol production , retrieved 2016-08-10.
  29. Samuels AL, Rensing KH, Douglas CJ, Mansfield SD, Dharmawardhana DP, Ellis BE (November 2002). "Cellular machinery of wood production: differentiation of secondary xylem in Pinus contorta var. latifolia". Planta. 216 (1): 72–82. doi:10.1007/s00425-002-0884-4. PMID   12430016. S2CID   20529001.
  30. Davin, L.B.; Lewis, N.G. (2005). "Lignin primary structures and dirigent sites". Current Opinion in Biotechnology. 16 (4): 407–415. doi:10.1016/j.copbio.2005.06.011. PMID   16023847.
  31. 1 2 3 Vane, C. H.; et al. (2003). "Biodegradation of Oak (Quercus alba) Wood during Growth of the Shiitake Mushroom (Lentinula edodes): A Molecular Approach". Journal of Agricultural and Food Chemistry. 51 (4): 947–956. doi:10.1021/jf020932h. PMID   12568554.
  32. 1 2 3 Vane, C. H.; et al. (2006). "Bark decay by the white-rot fungus Lentinula edodes: Polysaccharide loss, lignin resistance and the unmasking of suberin". International Biodeterioration & Biodegradation. 57 (1): 14–23. doi:10.1016/j.ibiod.2005.10.004.
  33. 1 2 Gadd, Geoffrey M; Sariaslani, Sima (2013). Advances in applied microbiology. Vol. 82. Oxford: Academic. pp. 1–28. ISBN   978-0124076792. OCLC   841913543.
  34. 1 2 de Gonzalo, Gonzalo; Colpa, Dana I.; Habib, Mohamed H.M.; Fraaije, Marco W. (2016). "Bacterial enzymes involved in lignin degradation". Journal of Biotechnology. 236: 110–119. doi: 10.1016/j.jbiotec.2016.08.011 . PMID   27544286.
  35. Tien, M (1983). "Lignin-Degrading Enzyme from the Hymenomycete Phanerochaete chrysosporium Burds". Science. 221 (4611): 661–3. Bibcode:1983Sci...221..661T. doi:10.1126/science.221.4611.661. PMID   17787736. S2CID   8767248.
  36. Cerisy, Tristan (May 2017). "Evolution of a Biomass-Fermenting Bacterium To Resist Lignin Phenolics". Applied and Environmental Microbiology. 83 (11). Bibcode:2017ApEnM..83E.289C. doi: 10.1128/AEM.00289-17 . PMC   5440714 . PMID   28363966.
  37. Pellerin, Brian A.; Hernes, Peter J.; Saraceno, JohnFranco; Spencer, Robert G. M.; Bergamaschi, Brian A. (May 2010). "Microbial degradation of plant leachate alters lignin phenols and trihalomethane precursors". Journal of Environmental Quality. 39 (3): 946–954. doi:10.2134/jeq2009.0487. ISSN   0047-2425. PMID   20400590.
  38. Hernes, Peter J. (2003). "Photochemical and microbial degradation of dissolved lignin phenols: Implications for the fate of terrigenous dissolved organic matter in marine environments". Journal of Geophysical Research. 108 (C9): 3291. Bibcode:2003JGRC..108.3291H. doi: 10.1029/2002JC001421 . Retrieved 2018-11-27.
  39. "Persistence of Soil Organic Matter as an Ecosystem Property". ResearchGate. Retrieved 2018-11-27.
  40. Dittmar, Thorsten (2015-01-01). "Reasons Behind the Long-Term Stability of Dissolved Organic Matter". Biogeochemistry of Marine Dissolved Organic Matter. pp. 369–388. doi:10.1016/B978-0-12-405940-5.00007-8. ISBN   978-0124059405.
  41. Wittkowski, Reiner; Ruther, Joachim; Drinda, Heike; Rafiei-Taghanaki, Foroozan (1992). Formation of smoke flavor compounds by thermal lignin degradation. ACS Symposium Series (Flavor Precursors). Vol. 490. pp. 232–243. ISBN   978-0-8412-1346-3.
  42. "TAPPI. T 222 om-02 – Acid-insoluble lignin in wood and pulp" (PDF).
  43. Harkin, John M. (November 1966). "Lignin production and detection in wood" (PDF). U.S. Forest Service Research. Note FPL-0148. Archived from the original (PDF) on 2020-03-05. Retrieved 2012-12-30.
  44. Lange, B. M.; Lapierre, C.; Sandermann, Jr (1995). "Elicitor-Induced Spruce Stress Lignin (Structural Similarity to Early Developmental Lignins)". Plant Physiology. 108 (3): 1277–1287. doi:10.1104/pp.108.3.1277. PMC   157483 . PMID   12228544.
  45. Glasser, Wolfgang G.; Glasser, Heidemarie R. (1974). "Simulation of Reactions with Lignin by Computer (Simrel). II. A Model for Softwood Lignin". Holzforschung. 28 (1): 5–11, 1974. doi:10.1515/hfsg.1974.28.1.5. S2CID   95157574.
  46. Hedges, John I.; Ertel, John R. (February 1982). "Characterization of lignin by gas capillary chromatography of cupric oxide oxidation products". Analytical Chemistry. 54 (2): 174–178. doi:10.1021/ac00239a007. ISSN   0003-2700.
  47. Hedges, John I.; Mann, Dale C. (1979-11-01). "The characterization of plant tissues by their lignin oxidation products". Geochimica et Cosmochimica Acta. 43 (11): 1803–1807. Bibcode:1979GeCoA..43.1803H. doi:10.1016/0016-7037(79)90028-0. ISSN   0016-7037.
  48. Vane, C. H.; et al. (2001). "The effect of fungal decay (Agaricus bisporus) on wheat straw lignin using pyrolysis–GC–MS in the presence of tetramethylammonium hydroxide (TMAH)". Journal of Analytical and Applied Pyrolysis. 60 (1): 69–78. doi:10.1016/s0165-2370(00)00156-x.
  49. Vane, C. H.; et al. (2001). "Degradation of Lignin in Wheat Straw during Growth of the Oyster Mushroom (Pleurotus ostreatus) Using Off-line Thermochemolysis with Tetramethylammonium Hydroxide and Solid-State 13C NMR". Journal of Agricultural and Food Chemistry. 49 (6): 2709–2716. doi:10.1021/jf001409a. PMID   11409955.
  50. Vane, C. H.; et al. (2005). "Decay of cultivated apricot wood (Prunus armeniaca) by the ascomycete Hypocrea sulphurea, using solid state 13C NMR and off-line TMAH thermochemolysis with GC–MS". International Biodeterioration & Biodegradation. 55 (3): 175–185. doi:10.1016/j.ibiod.2004.11.004.
  51. Ralph, John; Landucci, Larry L. (2010). "NMR of Lignin and Lignans". Lignin and Lignans: Advances in Chemistry. Boca Raton, FL: Taylor & Francis. pp. 137–244. ISBN   978-1574444865.

Further reading