You are on page 1of 22

Course Code 6452

ASSIGNMENT No. 1

Q. 1 Write in detail diagnostic characteristics, economic importance and distribution pattern of family Brassicaceae
(Cruciferae).

Ans:- In this article we will discuss about:- 1. Characters of Brassicaceae 2. Distribution of Brassicaceae 3. Economic
Importance 4. Affinities 5. Important Types.

Characters of Brassicaceae:
Flowers actinomorphic rarely zygomorphic, hermaphrodite; sepals four in two whorls of two each, petals four, diagonally
arranged-cruciform; stamens six, tetradynamous; gynoecium bicarpellary, syncarpous, parietal placentation, bilocular due to
the formation of flase septum (replum); fruit siliqua or silicula.

A. Vegetative characters:

ADVERTISEMENTS:

Habit:

Generally herbs, annual (Brassica, Capsella) or biennial or shrubs. Common Indian herbs are Eruca, Alyssum, Nasturtium,
Lepidium, Coronopus etc. Vegetative reproduction is by bulbils (Dentaria bulbifera) or by coral roots.

Roots:

Tap root, swollen on account of stored food materials. It may be conical (Radish), fusiform or napiform (Turnip).

ADVERTISEMENTS:

Stem:

Herbaceous, erect, cylindrical (Iberis, Brassica) rarely woody or some times reduced (Raphanus & Brassica species), glabrous
or hairy, solid and branched.

Leaves:

Alternate or sub-opposite, simple, exstipulate (Brassica campestris). May be cauline or radical (Raphanus), generally sessile,
hairy, entire and with unicostate reticulate venation.

B. Floral characters:

Inflorescence:

1
Raceme (Brassica campestris) corymbose raceme (Iberis) or corymb.

Flower:

Pedicellate, ebracteate, hermaphrodite, actinomorphic rarely zygomorphic (Iberis and Teesdalia), hypogynous, complete or
incomplete (Lepidium) and tetramerous.

ADVERTISEMENTS:

Calyx:

Sepals 4 arranged in two whorls of two each, polysepalous (2 antero-posterior and 2 lateral), 2 lateral sepals may be saccate,
imbricate aestivation, inferior.

Corolla:

Petals 4, alternate with sepals, polypetalous, petals arranged in the form of across known as cruciform. This arrangement is
characteristic of the family Petals usually clawed, petals generally equal rarely unequal (Iberis, Teesdalia) or sometimes petals
may be replaced by stamens (Capsella bursa pastoris).

ADVERTISEMENTS:

Androecium:

Stamens 6, arranged in two whorls, outer two stamens short and inner four long (2+4), tetradynamous, polyandrous, anthers
dithecous basifixed, introrse. Disc like nectaries, variable in number, present at the base of stamens. In some cases the
number of stamens is variable – 16 (Megacarpaea), 4 (Cardamine hirsuta), 2 (Coronopus) etc.

Gynoecium:

Bicarpellary rarely tricarpellary (Lepidium sativum), syncarpous, ovary superior, unilocular, becomes bilocular due to the
development of false septum called replum: parietal placentation, ovules many, style short, stigma simple or bifid. The
crucifer carpel has been a puzzling subject for the morphologists and their attention attracted towards its for a long time.
According to some there are only two carpels while others hold that there are four carpels.

ADVERTISEMENTS:

Fruit:

Siliqua or silicula, sometimes lomentum (Raphanus); when the valves separate in a siliqua the seeds remain attached to the
replum.

2
Seed:

Ex-albuminous. The germination of seed is epigeal.

ADVERTISEMENTS:

Pollination:

Self or cross pollinated; flowers are visited by insects due to the presence of nectaries. Cleistogamy is found in Cardamine
chenopodifolia. Anemophilous pollination is found in Pringlea.

Floral formula:

Distribution of Brassicaceae:
This family is also called Brassica family. The family includes 375 genera and 3200 species according to Willis. It is distributed
all over the world but mainly confined to the Mediterranean region and north temperature regions.

Economic Importance of Brassicaceae:


This family is of considerable economic importance.

1. Food:

The plants of this family which are cultivated as vegetable crops are:

Brassica oleracea var. botrytis (H. Phul gobhi), B. oleracea var. capitata (H. Band gobhi), B. oleracea var. caulorapa (H. Gand-
gobhi), Brassica campestris var. sarson (white mustard), Brassica rapa (H. Shalgam), Raphanus satiuus (H. Muli), are edible and
cooked as vegetables.

2. Oil:

The seed of B. campestris (or white mustard) yield mustard oil or Karwa-tel which is widely used as a cooking medium. B. nigra
(H. Kalirai) and B. juncea (H. rai) also produce oil.

After extracting oil the cake is left behind which is highly nutritious as a cattle feed; the oil cake is also used as soil fertilizer.
Raphanus seeds also produce a pungent oil which is often used in adulteration of sarson oil; this oil has digestive properties.

3. Medicines:

The leaves and tender shoots of Lepidium sativum are used in liver complaints, asthma, cough and bleeding piles. Rorippa
montana is an appetizer and a stimulant. The seeds of Cheiranthus cheiri are used in bronchitis and fever. The flowers are
used in paralysis and impotency. Lobularia is used for gonorrhoea. Iberis amara is used in rheumatism and gout.

4. Ornamentals:

3
Some plants are grown in gardens for their beautiful flowers viz. Cheiran thus cheiri (wall flower), Iberis amara (candituft)
Lobularia, Matthiola (stock), Hesperis (rocket), Alyssum, Lunaria (honesty) etc.

Primitive characters:

1. Leaves simple and alternate.

2. Flowers hermaphrodite, hypogynous and actinomorphic.

3. Calyx and corolla free.

4. Stamens polyandrous.

5. Ovules anatropous.

Advanced characters:

1. Plants are generally herbs-annual or biennials.

2. Leaves exstipulate.

3. Flowers ebracteate and sometimes zygomorphic (Iberis).

4. Gynoecium bicarpellary and syncarpous.

5. Fruit simple.

Affinities of Brassicaceae:
Rendle placed this family under the order Rhoedales; Bentham-Hooker placed it under the cohort Parietales. The family is
related to the Papaveraceae on one hand and to the Capparidaceae on the other. Bentham & Hooker and Hutchinson (1948,
1964) hold the view that Brassicaceae is derived from the Papaveraceous ancestors whereas Eames, Arber, Hayek and Puri
believe it to have a Capparidaceous alliance.

The three families, Capparidaceae, Brassicaceae (Cruciferae) and Papaveraceae have in common the features of tetramerous
perianth, bicarpellary syncarpous gynoecium and parietal placentation. These characters gave problematic issues as to
whether the Brassicaceae (Cruciferae) originated from the Capparidaceae or descended from the Papaveraceae.

The anatomy and morphology of stamens and carpels of cruciferous flower bears testimony to a papaverous ancestory. But in
Brassicaceae the stamens are tetradynamous and not in Papaveraceae.

Comparison of floral diagram indicates that Brassicaceae is closely allied to Capparidaceae. But in Brassicaceae gynophore and
variable number of stamens are absent where as these are the prominent characters of Capparidean flowers.

Within the Rhoedales reduction seems to have taken place in the number of stamens. In the Papaveraceae there are
numerous stamens but in its two subfamilies reduction has occurred. In the Hypecoideae there are only four stamens; in the
Fumarioideae the stamens are arranged in two bundles each with one dithecous and two monothecous anthers.

4
In the Capparidaceae the number of stamens range between several (as in Capparis) to six (as in Gynandropsis). Finally in
Cleome there are only four stamens. The floral diagram of Cleome spinosa with six stamens is remarkably similar to that of the
Brassicaceae (Cruciferae).

In this family the general condition is tetradynamous but may be reduced to only two (as in Coronopus). Celakovasky
considers the above view as most satisfactory.

Cronquist (1968) too considers that the Brassicaceae (Cruciferae) evolved from the Capparidaceae.

Common plants of the family:

1. Brassica campestris (Sarson) – a cultivated herb.

2. Iberis amara (Chandni) – annual, ornamental, herb cultivated in winter.

3. Cherianthus cheiri (Wall flower) – ornamental annual herb.

4. Rorippa monatna (Water cress) – semi wild.

5. Capsella bursa – pastoris (Shepherd’s purse) – common weed.

6. Farsetia jaquemontii – common weed.

7. Coronopus didymus (= Senebiera didyma) – wild in waste places.

8. Eruca sativa (Tara mira) – cultivated for seeds that yield an oil.

Division of the family and chief genera:

Linnaeus utilised the pod as the character for his classification.

He divided Tetradynamia into two orders:

Order 1. Siliculosae – fruit a silicula.

Order 2. Siliquosae – fruit a siliqua.

Prantl divided the family on the presence or absence of hairs into two series.

Series I. Hairs un-branched or hairs absent, never glandular:

This series includes two sub-families:

1. Thelypodieae:

With stigma developed equally all around, style undivided. Thelypodium, Pringlea.

2. Sinapeae:
5
With stigma better developed over placentae. Iberis, Brassica, Raphanus etc.

Series II. Hairs branched, sometimes glandular:

This series includes two sub-families:

1. Schizopetaleae:

With stigma equal all round. Physaria.

2. Hesparideae:

With stigma better developed above the placentae. Alyssum, Capsella, Cheiranthus etc.

O. E. Schulz (1936) divided the family into 19 tribes on the basis of a wide variety of characters.

Important Types of Brassicaceae:


1. Brassica campestris, Linn. (Fig. 31.1):

Brassica Campestris

Habit and habitat:

An annual herb, cultivated for seeds which yield oil.

Root:

Tap and branched.

Stem:

Herbaceous erect, cylindrical, solid, glabrous or hairy.

Leaf:

Simple, alternate, exstipulate, lower ones lyrate and upper oblong or lanceolate, unicostate reticulate venation, hairy, sessile.

Inflorescence:

A corymbose-raceme.

Flower:

Ebracteate, pedicellate, complete, actinomorphic, hermaphrodite, cruciform, tetramerous, hypogynous, and yellow.

Calyx:

6
Sepals 4 (2 + 2) in two whorls, outer whorl antero-posterior, the two lateral one saccate, green, polysepalous, inferior.

Corolla:

Androecium:

Stamens six, tetradynamous, in two whorls, the outer with two short lateral stamens and inner with four long stamens
arranged in two median pairs. Basifixed, polyandrous, introrse. Four green nectaries are present, on the inner side of each
short stamen and a similar one at the base but outside each pair of long median stamens, inferior.

Gynoecium:

Bicarpellary, syncarpous, superior, unilocular becoming bilocular by the development of false septum called – replum; parietal
placentation, style short, stigma bilobed.

Fruit:

Siliqua.

Seed:

Non-endospermic.

Floral formula:

2. Iberis amara, Linn. (Fig. 31.2):

Lberis Amara

Habit and habitat:

Herbaceous annual, cultivated in the gardens.

Root:

Tap, branched and annual.

Stem:

Herbaceous, erect, branched, cylindrical, solid, green and glabrous.

Leaf:

Cauline, ramal, alternate, simple, exstipulate, sessile, glabrous, or hairy, unicostate reticulate venation.
7
Inflorescence:

Corymbose raceme.

Flower:

Ebracteate, pedicellate, complete, zygomorphic, zygomorphy is due to two longer anterior petals, tetramerous, hypogynous,
white.

Calyx:

Sepals 4, polysepalous, in two whorls of two each, green, imbricate aestivation, inferior.

Corolla:

Petals 4, polypetalous, cruciform, petals unequal-2 posterior smaller and 2 anterior longer, valvate, white.

Androecium:

Stamens 6, polyandrous, arranged in two whorls; outer of 2 short stamens and inner of 4 longer stamens, tetradynamous,
anthers basifixed and introrse.

Gynoecium:

Bicarpellary, syncarpous, ovary superior, unilocular but becomes bilocular due to the formation of false septum – the replum,
parietal placentation; style short; stigma capitate.

Fruit:

A silicula.

Seed:

Non endospermic.

Floral formula:

3. Coronopus didymus, Linn. (Syn. Senebiera didyma) (Fig. 31.3):

Coronopus Didymus

Habit and habitat:

An annual, winter weed.

8
Root:

Tap, branched and annual.

Stem:

Prostrate, herbaceous, cylindrical, branched, solid, green and glabrous.

Leaf:

Alternate, exstipulate, sub-sessile or sessile, simple, pinnatifid, glabrous, unicostate reticulate venation.

Inflorescence:

A typical raceme.

Flower:

Pedicellate, bracteate, complete, hermaphrodite, actinomorphic, hypogynous, greenish white, small.

Calyx:

Sepals 4, polysepalous in two whorls of two each, antero-posterior sepals form the outer whorl, green, linear, inferior.

Corolla:

Petals 4 polypetalous, very small and scale-like, cruciform, whitish green, alternating with the sepals, valvate, inferior.

Androecium:

Stamens 2, anterio-posterior, polyandrous, small nectaries are present at the bases of stamens, anthers basifixed, dithecous,
introrse.

Gynoecium:

Bicarpellary, syncarpous superior, unilocular becoming bilocular due to the formation of false septum, parietal placentation, 2
ovules per placentum, style small, stigma bifid.

Fruit:

A silicula.

Seed:

Non-endospermic

Q. 2 Describe importance of binomial Nomenclature.

9
Ans:- The most well-known living things have common names. For example, you are probably familiar with the small,
red insects dotted with little black spots. You might call them 'ladybugs' or 'ladybird beetles.' But did you know there are
actually many different species of these insects? Just using common names may make it difficult for scientists to
differentiate between them, so every species is given a unique scientific name.

Binomial nomenclature is the formal naming system for living things that all scientists use. It gives every species a two-
part scientific name. For example, a ladybug found in the United States goes by the fancy name of Harmonia axyridis.

The first part of a scientific name, like Harmonia, is called the genus. A genus is typically the name for a small group of
closely related organisms. The second part of a scientific name, axyridis in this example, is the specific epithet. It is used
to identify a particular species as separate from others belonging to the same genus. Together, the genus plus the
specific epithet is the full scientific name for an organism.

I bet that you actually already know the scientific name for at least one animal, although you may not have realized it.
Ever heard of the dinosaur T. rex? T. rex is actually a scientific name - the 'T' is just an abbreviation of the genus
Tyrannosaurus. So the scientific name is actually Tyrannosaurus rex.

Binomial Nomenclature Rules


Because scientific names are unique species identifiers, they ensure that there is never any confusion as to which
organism a scientist may be referring. Additionally, there are some important rules that must be followed to keep all
binomial names standardized:

Q. 3 Elaborate primary and secondary growth of stems.


Ans:- Growth in plants occurs as the stems and roots lengthen. Some plants, especially those that are woody, also increase in thickness during their life span.
The increase in length of the shoot and the root is referred to as primary growth. It is the result of cell division in the shoot apical meristem. Secondary growth is
characterized by an increase in thickness or girth of the plant. It is caused by cell division in the lateral meristem. Herbaceous plants mostly undergo primary
growth, with little secondary growth or increase in thickness. Secondary growth, or “wood”, is noticeable in woody plants; it occurs in some dicots, but occurs very
rarely in monocots.

Figure 30.2C.130.2C.1: Primary and secondary growth: In woody


plants, primary growth is followed by secondary growth, which allows the plant stem to increase in thickness or girth. Secondary vascular tissue is added as the
plant grows, as well as a cork layer. The bark of a tree extends from the vascular cambium to the epidermis.
Some plant parts, such as stems and roots, continue to grow throughout a plant’s life: a phenomenon called indeterminate growth. Other plant parts, such as
leaves and flowers, exhibit determinate growth, which ceases when a plant part reaches a particular size.
Primary Growth
10
Most primary growth occurs at the apices, or tips, of stems and roots. Primary growth is a result of rapidly-dividing cells in the apical meristems at the shoot tip
and root tip. Subsequent cell elongation also contributes to primary growth. The growth of shoots and roots during primary growth enables plants to continuously
seek water (roots) or sunlight (shoots).
The influence of the apical bud on overall plant growth is known as apical dominance, which diminishes the growth of axillary buds that form along the sides of
branches and stems. Most coniferous trees exhibit strong apical dominance, thus producing the typical conical Christmas tree shape. If the apical bud is removed,
then the axillary buds will start forming lateral branches. Gardeners make use of this fact when they prune plants by cutting off the tops of branches, thus
encouraging the axillary buds to grow out, giving the plant a bushy shape.
Secondary Growth
The increase in stem thickness that results from secondary growth is due to the activity of the lateral meristems, which are lacking in herbaceous plants. Lateral
meristems include the vascular cambium and, in woody plants, the cork cambium. The vascular cambium is located just outside the primary xylem and to the
interior of the primary phloem. The cells of the vascular cambium divide and form secondary xylem ( tracheids and vessel elements) to the inside and secondary
phloem (sieve elements and companion cells) to the outside. The thickening of the stem that occurs in secondary growth is due to the formation of secondary
phloem and secondary xylem by the vascular cambium, plus the action of cork cambium, which forms the tough outermost layer of the stem. The cells of the
secondary xylem contain lignin, which provides hardiness and strength.
In woody plants, cork cambium is the outermost lateral meristem. It produces cork cells (bark) containing a waxy substance known as suberin that can repel
water. The bark protects the plant against physical damage and helps reduce water loss. The cork cambium also produces a layer of cells known as phelloderm,
which grows inward from the cambium. The cork cambium, cork cells, and phelloderm are collectively termed the periderm. The periderm substitutes for the
epidermis in mature plants. In some plants, the periderm has many openings, known as lenticels, which allow the interior cells to exchange gases with the outside
atmosphere. This supplies oxygen to the living- and metabolically-active cells of the cortex, xylem, and phloem.
Figure 30.2C.130.2C.1: Example of lenticels: Lenticels on the bark of this cherry tree enable the woody stem to exchange gases with the surrounding atmosphere.
Annual Rings
The activity of the vascular cambium gives rise to annual growth rings. During the spring growing season, cells of the secondary xylem have a large internal
diameter; their primary cell walls are not extensively thickened. This is known as early wood, or spring wood. During the fall season, the secondary xylem
develops thickened cell walls, forming late wood, or autumn wood, which is denser than early wood. This alternation of early and late wood is due largely to a
seasonal decrease in the number of vessel elements and a seasonal increase in the number of tracheids. It results in the formation of an annual ring, which can
be seen as a circular ring in the cross section of the stem. An examination of the number of annual rings and their nature (such as their size and cell wall
thickness) can reveal the age of the tree and the prevailing climatic conditions during each season.
Figure 30.2C.130.2C.1: Annual growth rings: The rate of wood growth increases in summer and decreases in winter, producing a characteristic ring for each year
of growth. Seasonal changes in weather patterns can also affect the growth rate. Note how the rings vary in thickness.

11
Q. 4 What are biomolecules? Give introduction to structure and functions of bimolecular .
Ans:- A biomolecule or biological molecule is a loosely used term for molecules present in organisms that are essential
to one or more typically biological processes, such as cell division, morphogenesis, or development.[1] Biomolecules
include large macromolecules (or polyanions) such as proteins, carbohydrates, lipids, and nucleic acids, as well as small
molecules such as primary metabolites, secondary metabolites and natural products. A more general name for this class
of material is biological materials. Biomolecules are usually[citation needed] endogenous, produced within the
organism[2] but organisms usually need exogenous biomolecules, for example certain nutrients, to survive.

Biology and its subfields of biochemistry and molecular biology study biomolecules and their reactions. Most
biomolecules are organic compounds, and just four elements—oxygen, carbon, hydrogen, and nitrogen—make up 96%
of the human body's mass. But many other elements, such as the various biometals, are present in small amounts.

The uniformity of both specific types of molecules (the biomolecules) and of certain metabolic pathways are invariant
features among the wide diversity of life forms; thus these biomolecules and metabolic pathways are referred to as
"biochemical universals"[3] or "theory of material unity of the living beings", a unifying concept in biology, along with
cell theory and evolution theory.[4]

Contents
1 Types of biomolecules
2 Nucleosides and nucleotides
2.1 DNA and RNA structure
3 Saccharides
4 Lignin
5 Lipid
6 Amino acids
6.1 Protein structure
6.1.1 Apoenzymes
6.1.2 Isoenzymes
7 See also
8 References
9 External links
Types of biomolecules
A diverse range of biomolecules exist, including:

Small molecules:
Lipids, fatty acids, glycolipids, sterols, monosaccharides
Vitamins
Hormones, neurotransmitters
Metabolites
Monomers, oligomers and polymers:
Biomonomers Bio-oligo Biopolymers Polymerization process Covalent bond name between monomers
Amino acids Oligopeptides Polypeptides, proteins (hemoglobin...) Polycondensation Peptide bond
Monosaccharides Oligosaccharides Polysaccharides (cellulose...) Polycondensation Glycosidic bond
Isoprene Terpenes Polyterpenes: cis-1,4-polyisoprene natural rubber and trans-1,4-polyisoprene gutta-
percha Polyaddition
12
Nucleotides Oligonucleotides Polynucleotides, nucleic acids (DNA, RNA) Phosphodiester bond
Nucleosides and nucleotides
Main articles: Nucleosides and Nucleotides
Nucleosides are molecules formed by attaching a nucleobase to a ribose or deoxyribose ring. Examples of these include
cytidine (C), uridine (U), adenosine (A), guanosine (G), and thymidine (T).

Nucleosides can be phosphorylated by specific kinases in the cell, producing nucleotides. Both DNA and RNA are
polymers, consisting of long, linear molecules assembled by polymerase enzymes from repeating structural units, or
monomers, of mononucleotides. DNA uses the deoxynucleotides C, G, A, and T, while RNA uses the ribonucleotides
(which have an extra hydroxyl(OH) group on the pentose ring) C, G, A, and U. Modified bases are fairly common (such as
with methyl groups on the base ring), as found in ribosomal RNA or transfer RNAs or for discriminating the new from old
strands of DNA after replication.[5]

Each nucleotide is made of an acyclic nitrogenous base, a pentose and one to three phosphate groups. They contain
carbon, nitrogen, oxygen, hydrogen and phosphorus. They serve as sources of chemical energy (adenosine triphosphate
and guanosine triphosphate), participate in cellular signaling (cyclic guanosine monophosphate and cyclic adenosine
monophosphate), and are incorporated into important cofactors of enzymatic reactions (coenzyme A, flavin adenine
dinucleotide, flavin mononucleotide, and nicotinamide adenine dinucleotide phosphate).[6]

DNA and RNA structure


Main articles: DNA and Nucleic acid structure
DNA structure is dominated by the well-known double helix formed by Watson-Crick base-pairing of C with G and A with
T. This is known as B-form DNA, and is overwhelmingly the most favorable and common state of DNA; its highly specific
and stable base-pairing is the basis of reliable genetic information storage. DNA can sometimes occur as single strands
(often needing to be stabilized by single-strand binding proteins) or as A-form or Z-form helices, and occasionally in
more complex 3D structures such as the crossover at Holliday junctions during DNA replication.[6]

Stereo 3D image of a group I intron ribozyme (PDB file 1Y0Q); gray lines show base pairs; ribbon arrows show double-
helix regions, blue to red from 5' to 3' end; white ribbon is an RNA product.
RNA, in contrast, forms large and complex 3D tertiary structures reminiscent of proteins, as well as the loose single
strands with locally folded regions that constitute messenger RNA molecules. Those RNA structures contain many
stretches of A-form double helix, connected into definite 3D arrangements by single-stranded loops, bulges, and
junctions.[7] Examples are tRNA, ribosomes, ribozymes, and riboswitches. These complex structures are facilitated by
the fact that RNA backbone has less local flexibility than DNA but a large set of distinct conformations, apparently
because of both positive and negative interactions of the extra OH on the ribose.[8] Structured RNA molecules can do
highly specific binding of other molecules and can themselves be recognized specifically; in addition, they can perform
enzymatic catalysis (when they are known as "ribozymes", as initially discovered by Tom Cech and colleagues).[9]

Saccharides
Monosaccharides are the simplest form of carbohydrates with only one simple sugar. They essentially contain an
aldehyde or ketone group in their structure.[10] The presence of an aldehyde group in a monosaccharide is indicated by
the prefix aldo-. Similarly, a ketone group is denoted by the prefix keto-.[5] Examples of monosaccharides are the
hexoses, glucose, fructose, Trioses, Tetroses, Heptoses, galactose, pentoses, ribose, and deoxyribose. Consumed
fructose and glucose have different rates of gastric emptying, are differentially absorbed and have different metabolic
fates, providing multiple opportunities for 2 different saccharides to differentially affect food intake.[10] Most
saccharides eventually provide fuel for cellular respiration.

13
Disaccharides are formed when two monosaccharides, or two single simple sugars, form a bond with removal of water.
They can be hydrolyzed to yield their saccharin building blocks by boiling with dilute acid or reacting them with
appropriate enzymes.[5] Examples of disaccharides include sucrose, maltose, and lactose.

Polysaccharides are polymerized monosaccharides, or complex carbohydrates. They have multiple simple sugars.
Examples are starch, cellulose, and glycogen. They are generally large and often have a complex branched connectivity.
Because of their size, polysaccharides are not water-soluble, but their many hydroxy groups become hydrated
individually when exposed to water, and some polysaccharides form thick colloidal dispersions when heated in water.[5]
Shorter polysaccharides, with 3 - 10 monomers, are called oligosaccharides.[11] A fluorescent indicator-displacement
molecular imprinting sensor was developed for discriminating saccharides. It successfully discriminated three brands of
orange juice beverage.[12] The change in fluorescence intensity of the sensing films resulting is directly related to the
saccharide concentration.[13]

Lignin
Lignin is a complex polyphenolic macromolecule composed mainly of beta-O4-aryl linkages. After cellulose, lignin is the
second most abundant biopolymer and is one of the primary structural components of most plants. It contains subunits
derived from p-coumaryl alcohol, coniferyl alcohol, and sinapyl alcohol[14] and is unusual among biomolecules in that it
is racemic. The lack of optical activity is due to the polymerization of lignin which occurs via free radical coupling
reactions in which there is no preference for either configuration at a chiral center.

Lipid
Lipids (oleaginous) are chiefly fatty acid esters, and are the basic building blocks of biological membranes. Another
biological role is energy storage (e.g., triglycerides). Most lipids consist of a polar or hydrophilic head (typically glycerol)
and one to three nonpolar or hydrophobic fatty acid tails, and therefore they are amphiphilic. Fatty acids consist of
unbranched chains of carbon atoms that are connected by single bonds alone (saturated fatty acids) or by both single
and double bonds (unsaturated fatty acids). The chains are usually 14-24 carbon groups long, but it is always an even
number.

For lipids present in biological membranes, the hydrophilic head is from one of three classes:

Glycolipids, whose heads contain an oligosaccharide with 1-15 saccharide residues.


Phospholipids, whose heads contain a positively charged group that is linked to the tail by a negatively charged
phosphate group.
Sterols, whose heads contain a planar steroid ring, for example, cholesterol.
Other lipids include prostaglandins and leukotrienes which are both 20-carbon fatty acyl units synthesized from
arachidonic acid. They are also known as fatty acids

Amino acids
Amino acids contain both amino and carboxylic acid functional groups. (In biochemistry, the term amino acid is used
when referring to those amino acids in which the amino and carboxylate functionalities are attached to the same
carbon, plus proline which is not actually an amino acid).

Modified amino acids are sometimes observed in proteins; this is usually the result of enzymatic modification after
translation (protein synthesis). For example, phosphorylation of serine by kinases and dephosphorylation by
phosphatases is an important control mechanism in the cell cycle. Only two amino acids other than the standard twenty
are known to be incorporated into proteins during translation, in certain organisms:

Selenocysteine is incorporated into some proteins at a UGA codon, which is normally a stop codon.
14
Pyrrolysine is incorporated into some proteins at a UAG codon. For instance, in some methanogens in enzymes that are
used to produce methane.
Besides those used in protein synthesis, other biologically important amino acids include carnitine (used in lipid
transport within a cell), ornithine, GABA and taurine.

Protein structure
Main articles: Protein structure, Protein primary structure, Protein secondary structure, Protein tertiary structure, and
Protein quaternary structure
The particular series of amino acids that form a protein is known as that protein's primary structure. This sequence is
determined by the genetic makeup of the individual. It specifies the order of side-chain groups along the linear
polypeptide "backbone".

Proteins have two types of well-classified, frequently occurring elements of local structure defined by a particular
pattern of hydrogen bonds along the backbone: alpha helix and beta sheet. Their number and arrangement is called the
secondary structure of the protein. Alpha helices are regular spirals stabilized by hydrogen bonds between the backbone
CO group (carbonyl) of one amino acid residue and the backbone NH group (amide) of the i+4 residue. The spiral has
about 3.6 amino acids per turn, and the amino acid side chains stick out from the cylinder of the helix. Beta pleated
sheets are formed by backbone hydrogen bonds between individual beta strands each of which is in an "extended", or
fully stretched-out, conformation. The strands may lie parallel or antiparallel to each other, and the side-chain direction
alternates above and below the sheet. Hemoglobin contains only helices, natural silk is formed of beta pleated sheets,
and many enzymes have a pattern of alternating helices and beta-strands. The secondary-structure elements are
connected by "loop" or "coil" regions of non-repetitive conformation, which are sometimes quite mobile or disordered
but usually adopt a well-defined, stable arrangement.[15]

The overall, compact, 3D structure of a protein is termed its tertiary structure or its "fold". It is formed as result of
various attractive forces like hydrogen bonding, disulfide bridges, hydrophobic interactions, hydrophilic interactions, van
der Waals force etc.

When two or more polypeptide chains (either of identical or of different sequence) cluster to form a protein, quaternary
structure of protein is formed. Quaternary structure is an attribute of polymeric (same-sequence chains) or heteromeric
(different-sequence chains) proteins like hemoglobin, which consists of two "alpha" and two "beta" polypeptide chains.

Apoenzymes
An apoenzyme (or, generally, an apoprotein) is the protein without any small-molecule cofactors, substrates, or
inhibitors bound. It is often important as an inactive storage, transport, or secretory form of a protein. This is required,
for instance, to protect the secretory cell from the activity of that protein. Apoenzymes become active enzymes on
addition of a cofactor. Cofactors can be either inorganic (e.g., metal ions and iron-sulfur clusters) or organic compounds,
(e.g., [Flavin group|flavin] and heme). Organic cofactors can be either prosthetic groups, which are tightly bound to an
enzyme, or coenzymes, which are released from the enzyme's active site during the reaction.

Isoenzymes
Isoenzymes, or isozymes, are multiple forms of an enzyme, with slightly different protein sequence and closely similar
but usually not identical functions. They are either products of different genes, or else different products of alternative
splicing. They may either be produced in different organs or cell types to perform the same function, or several
isoenzymes may be produced in the same cell type under differential regulation to suit the needs of changing
development or environment. LDH (lactate dehydrogenase) has multiple isozymes, while fetal hemoglobin is an example
of a developmentally regulated isoform of a non-enzymatic protein. The relative levels of isoenzymes in blood can be
used to diagnose problems in the organ of secretion .
15
Q. 5 What is vascular cambium? Describe its importance in botany.
Ans:- The vascular cambium is the main meristem in the stem, producing undifferentiated wood cells inwards and
bark cells outwards. The thickness of the vascular cambium varies from around six cells during dormant periods to
around 14 during the most active periods of growth (Figure 5.4A–C). Being a meristem the cambium consists of
flattened, undifferentiated cells. These undifferentiated cells possess no defense capabilities, although the cambium
quickly can be reprogrammed to produce cells that are differentiated into PP cells or traumatic resin ducts. Since the
cambium itself is defenseless, but crucial for maintaining stem growth and tree integrity, it must be protected by the
different defense structures in the secondary phloem, cortex, and periderm.

From Cambium to Early Cell Differentiation Within the Secondary Vascular System
Peter Barlow, in Vascular Transport in Plants, 2005

Publisher Summary
Vascular cambium of both roots and shoots contains two types of cells: long, spindle-shaped fusiform cells and smaller,
cuboidal ray parenchyma cells. Ray initials are regularly interspersed with the fusiform initials on the cambial perimeter
and the radially elongated files to which they give rise intrude, like the spokes of a bicycle wheel, into both secondary
xylem and phloem. Irrespective of whether they are ray or fusiform cells, cambial initial cells are bidirectional in their cell
production. Each initial produces alternating sequences of new cells from either its inward- or outward-facing surfaces
that pass into the secondary xylem and phloem domains, respectively. Among the differentiated cells produced by the
cambial fusiform cells are those which have become adapted for long-distance vertical transport of solutes (tracheids,
xylem vessel elements, and phloem sieve cells) and for the assistance of these processes. Other cells (fibers, and also the
tracheids) are adapted for the mechanical support of the plant. Ray cells also synthesize and transport radially secondary
metabolites into the interior of the wood, as well as storing and transporting trophic materials to the cambium. From a
mechanical point of view, rays physically bolt together the annual rings of xylem, thus preventing shearing of these
groups of cells when the stem is bent. This chapter highlights the features of the cambial meristem, mainly in trees, that
bear on the development of the vertical and radial transport systems of stems and roots and discusses some of the
earliest stages of xylem vessel, phloem, and ray development.

Introduction
Donald E. Fosket, in Plant Growth and Development, 1994

The Vascular Cambium and Secondary Growth


The vascular cambium and cork cambium are secondary meristems that are formed in stems and roots after the tissues
of the primary plant body have differentiated. The vascular cambium is responsible for increasing the diameter of stems
and roots and for forming woody tissue. The cork cambium produces some of the bark. In dicot stems, the vascular
cambium initially differentiates from procambial cells within the vascular bundles (Fig. 1.8A). This fascicular cambium
may contribute additional cells to both the xylem and the phloem of the bundle. At some point the cambium expands
into the ground tissue between the vascular bundles, forming an interfascicular cambium, completing the ring of
vascular cambium (Fig. 1.8B). Cell division by the cambium produces cells that become secondary xylem and phloem. As
secondary phloem and xylem tissue accumulates, it both increases the girth of the stem and forms wood and bark.
Because cambial activity is seasonal in temperate zone plants, the wood and bark are laid down in distinct annual rings
(Fig. 1.8C). Monocots do not have a vascular cambium, even though some of them, such as palms and the Joshua tree,
exhibit secondary growth. Instead, they have a thickening meristem that produces secondary ground tissue. This
increases the girth of the stem and additional vascular bundles differentiate within the secondary ground tissue.

16
Sign in to download full-size image
Figure 1.8. Secondary growth: the origin and structure of vascular cambium in the stem

The vascular cambium is formed in mature dicot stems after stem elongation stops. (A) Primary xylem and phloem
differentiate from procambial tissue in the vascular bundles, and a fascicular cambium is formed from procambial tissue
separating these tissues. (B) Later, an interfascicular cambium appears between the vascular bundles that is continuous
with the fascicular cambium. (C) The further development of the cambium results in the formation of a cylinder of
vascular tissue. (D) The vascular cambium is a layer of pluripotent dividing cells whose derivatives differentiate as either
xylem elements (vessel members, tracheids, fibers, or xylem parenchyma) or phloem elements (sieve tube members,
companion cells, fibers, or parenchyma). (E) The dividing cells of the vascular cambium consist of long, narrow fusiform
initials, from which the tracheary elements are derived, and ray initials, from which ray parenchyma is formed.

Based on Wilson, C. L., and Loomis, W. E. (1967). Botany. Holt, New York.Copyright © 1967
The vascular cambium is composed of two kinds of cells, ray initials and fusiform initials. In cross section these look very
similar. Both are small, flattened cells with thin walls. When viewed in tangential section, however, ray initials can be
seen to be relatively short, small cells, whereas fusiform initials are very long and narrow (Fig. 1.8D). In gymnosperms
the fusiform initials often are several millimeters in length. Dicot fusiform initials are much shorter, but some still are up
to 0.5 mm in length. Cell division in the fusiform initials usually is tangential and the cell is partitioned down its long axis,
forming two equally long, narrow cells. Some of the cells produced by the cambial initials continue to divide, whereas
others differentiate. Tracheary elements or sieve elements differentiate from derivatives of the fusiform initials, and
derivatives of the ray initials differentiate as ray parenchyma. The ray parenchyma permits transport of water from the
xylem into the cambium and the tissues of phloem, as well as transport of photosynthate from the phloem into the
cambium and the living cells of the xylem.

The cork cambium also is a secondary meristem, containing meristematic cells. The cork cambium forms a major portion
of the bark of woody plants. The secondary phloem also is part of the bark, but of course phloem is produced by the
vascular cambium. The cork cambium first arises within the cortex as a concentric layer forming a cylinder of dividing
cells (Fig. 1.9). The derivatives of this meristematic cell layer differentiate as cork, or phellem, toward the outside of the
stem, whereas derivatives produced toward the inner part of the stem differentiate as phelloderm. Suberin is deposited
in the cell walls of the phellem and they are dead at maturity. They protect the stem from water loss and from
mechanical damage. As the tree increases in girth, the outer layers of bark are sloughed off. Additional cork cambia arise
within the secondary phloem as the plant develops.

Sign in to download full-size image


Figure 1.9. Cross section through the stem of a woody dicot showing the development of a cork cambium

(A) Based on Raven, P. H., and Curtis, H. (1970). Biology of Plants. Worth Publishing Company, New York. (B) Redrawn
with permission from Wilson, C. L. and Loomis, W. E. (1967). Botany. Holt, New York.Copyright © 1970
Introduction to Vascular Plant Morphology and Anatomy
Thomas N. Taylor, ... Michael Krings, in Paleobotany (Second Edition), 2009

Vascular Cambium
The vascular cambium arises between the primary xylem and phloem of a young stem or root. Parenchymatous cells
become meristematic and begin to produce secondary xylem or wood toward the inside of the cambium and secondary
phloem toward the outside of the cambium. The cambium itself remains meristematic, except in some unusual cases,
for example, in the Carboniferous arborescent lycopsids (Chapter 9) and may range from a single layer to several layers
of meristematic cells (FIG. 7.26). If the primary xylem is a solid core, as in some fossils, the cambium begins development
17
as a complete cylinder (a ring, as seen in cross section) between the primary xylem and phloem. If the primary vascular
tissue occurs in bundles, as is the case in woody dicots and gymnosperms, the cambium begins development within the
bundle—the fascicular cambium. Then, parenchyma cells between the bundles become meristematic—the
interfascicular cambium—and connect the fascicular cambia together so that the cambium eventually forms a complete
ring around the axis, between the primary xylem and phloem.

Sign in to download full-size image


FIGURE 7.26. Cross section of Pinus sp. stem showing radial files of vascular cambium initials (C) (Extant). Bar=100 μm.

Cambial cells or initials divide primarily by periclinal divisions (parallel to the surface of the axis) on their inner and outer
faces, producing files of cells along the radii of the axis. The presence of these orderly files is one way to distinguish
secondary growth in fossil axes. Cambial initials must also divide anticlinally (perpendicular to the surface) to produce
more cambial cells as the circumference of the axis continues to increase due to the production of secondary tissue.
There are two types of initial cells in the vascular cambium. Fusiform initials are elongate cells that produce the
conducting cells in both the secondary xylem and secondary phloem and the other cells in the axial system. Ray initials
are shorter, generally rectangular cells, which give rise to cells in the ray system (see section “Secondary Xylem”).
Generally, many more secondary xylem cells are produced than secondary phloem; indeed, in most living trees the bulk
of the trunk represents secondary xylem or wood.

The vascular cambium in roots arises in the same place as in stems, that is, between the primary xylem and phloem, but
since the primary xylem in many roots is lobed or furrowed, the cambium initially also has this shape. As the root
continues to develop, however, more secondary xylem is produced in the furrows so that the cambium eventually has a
cylindrical shape, just as it does in stems. See section “Secondary Xylem” and “Phloem” (later) for the cell types
produced by the vascular cambium.

The Vascular Cambium of Trees and its Involvement in Defining Xylem Anatomy
Uwe Schmitt, ... Risto Jalkanen, in Secondary Xylem Biology, 2016

Abstract
The vascular cambium of trees is a secondary meristem and is responsible for the formation of the xylem and phloem.
The main focus of this chapter is on the xylem, specifically on the following three topics, demonstrating that the
cambium is not only responsible for the quantitative side of xylem formation, but also for the expression of stable
anatomical features essential for wood identification. In this complex process, we first describe the seasonal cambial
activity and its environmental control. Second, we discuss the cambium’s involvement in the restoration of tissues after
injuries. Third, we examine the cambium-dependent shaping of taxa-specific wood anatomical characteristics. The
results are mainly based on light microscopy; however, electron microscopy was also occasionally used to reveal
structural features on the cellular level.

Functional Significance of Cambial Development in Vertebraria Roots


Anne-Laure Decombeix, Nicholas P. Rowe, in Transformative Paleobotany, 2018

4.1 Implications of a Derived Cambial Development


The derived vascular cambium present in Vertebraria resulted in a complex geometrical organization that likely had a
significant effect on the functional biology and life history of the whole Glossopteris plant. The ensemble of
developmental motifs behind this structural organization in Vertebraria is a remarkable example of how simple changes
in developmental timing can lead to (1) a strong departure from a typical anatomical structure, (2) a wide diversity of
geometries and shapes between developmental stages, and (3) potentially major changes in mechanical and hydraulic
18
functioning between young and old stages and from the distal to proximal parts of the root system. So just what are the
functional implications of these changes? How can they be interpreted at the level of the whole plant? And to what
extent can they represent adaptations for life in high-latitude wetlands in the Palaeozoic?

Apical Dominance and Some Other Phenomena Illustrating Correlative Effects of Hormones
Lalit M. Srivastava, in Plant Growth and Development: Hormones and Environment, 2002

3.1. IAA Is an Important Factor in Reactivation of Cambium in Spring


In temperate climates, vascular cambium becomes dormant in the fall and resumes meristematic activity in the spring. It
is commonly assumed that IAA is involved in cambial reactivation, i.e., induction of cell division activity. It has also been
assumed that cambial activity proceeds from the top of the trunk to the base, a view that may be derived from the fact
that IAA is produced in flushing apical and lateral buds and young shoots and flows basipetally. There is some evidence
for a basipetal progression of cambial activation in diffuse porous woods based on bioassays. However, studies on
cambia of conifers as well as diffuse- and ring-porous dicot woods, while demonstrating that IAA is required for cell
divisions in the cambial zone, do not support the assumption that cambial activation proceeds basipetally in the main
trunk. Such basipetal progression is seen only in young parts of a tree, usually the first year's growth; the rest of the
trunk is reactivated more or less simultaneously.

Measurements of endogenous IAA in tree trunks at different heights using modern methods of analysis and quantitation
are very few. They are also difficult because sampling pieces of bark, cambium, and wood from tree trunks takes time
and quick freezing of relatively large samples in liquid nitrogen or isopentane still does not stop the mobility of small
molecules and ions instantaneously. Nonetheless, studies have been made and indicate that the situation is more
complex than previously realized. A vertical gradient in IAA concentration is seen mostly in young stems and branches
and in trees that are growing vigorously. The gradient is not so clear and may even be nonexistent in older stems or in
slow-growing trees. Moreover, not all IAA moving down basipetally comes from the shoot apex. Feeding 13C-labeled IAA
to a decapitated pine shoot showed isotopic dilution down the trunk, which suggested that at least some IAA in the
trunk is synthesized locally at lower levels. Finally, dormant cambium also has significant amounts of IAA, which could be
mobilized in spring.

The site of polar transport of IAA in tree trunks is thought to be the cambial zone. It has been mentioned before that it is
possible to measure very small quantities of hormones in tissue sections or small samples (see Chapter 5). In several
papers, IAA concentrations were monitored in individual tangential sections of a pine stem and data were integrated to
give a profile of IAA concentrations in the cambial zone and differentiating and mature secondary xylem and phloem
cells on either side (Fig. 14-39). Data show that the highest concentrations of IAA occur in the cambial zone and fall off in
a gradient on either side in the differentiating secondary xylem and secondary phloem, with fully mature tissues showing
very little IAA.

Sign in to download full-size image

Sign in to download full-size image


FIGURE 14-39. Schematic drawing of the specimen block and radial distribution of IAA in the cambial zone and secondary
tissues of pine (Pinus sylvestris). (A) Tangential longitudinal sections (30 μm in thickness, using a cryomicrotome at
−20°C) were obtained starting from the outer phloem and into the xylem tissue. IAA content was measured in each
section (sample) using a modified GC-MS procedure. Transverse sections at ends were used for the determination of
sample position. (B) Radial distribution of IAA in two representative trees; one sampled in late June at the height of
cambial activity and the other sampled during dormancy in mid-January. Each column represents the 30-μm tangential
section. Endogenous IAA content per cm2 section is indicated with black squares. NFP, nonfunctional phloem; FP,
19
functional phloem; CZ, cambial zone; ET and DT, expanding and differentiating tracheids; MT, mature tracheids. The
average number of radial file cells in each developmental zone is given on the right.

With permission from Uggla et al. (1996), ©1996 National Academy of Sciences, USA.
It would be expected that the IAA concentration in the cambial zone at any one location in the trunk would be higher in
spring/summer when cambium is actively producing xylem and phloem than in winter when it is dormant. However, the
summer and winter samples did not show much seasonal fluctuation, although there was a broadening of the IAA
gradient in spring/summer and a narrowing of the gradient in winter (Fig. 14-39B). The presence of IAA in the dormant
cambium suggests, by inference, that the cessation of cambial activity in late summer-early fall is not controlled by IAA,
a suggestion that is supported by feeding experiments where IAA supplied to shoots does not prevent the cambium
from becoming dormant. Environmental factors, such as temperature and shortening daylength, seem to be involved in
the induction of cambial dormancy. Although the concentration of IAA did not show much seasonal variation, the active
cambium contained a greater amount of IAA than the dormant cambium, which indicates that higher amounts of IAA are
produced and utilized, i.e., there is a higher flux of IAA in the cambial zone in the summer months. The observation that
the IAA content in differentiating xylem and phloem tissues was about the same is difficult to explain because higher IAA
concentrations are known to promote xylem differentiation (see below). It could be that other factors besides IAA, such
as sugars and gibberellins, may also control the developmental fate of cambial derivatives.

Special Features of Plant Development


Lalit M. Srivastava, in Plant Growth and Development: Hormones and Environment, 2002

3.2.3. Secondary Growth and Vascular Cambium


In gymnosperms and woody dicots, a vascular cambium makes its appearance in that region of root or stem that has
ceased elongating and produces secondary xylem and phloem. The addition of secondary vascular tissues, especially
xylem, adds to the girth of these organs and provides the needed structural support to trees. Small amounts of
secondary growth may also occur in some species in petioles and midveins of leaves and in axes that bear flowers, but
because these organs have only a limited life span, it is never extensive. Many herbaceous dicots also develop a
cambium, but it may not form a complete ring and its activity may be restricted to the vascular bundles.

The vascular cambium is a layer of meristematic cells (or initials) that arises between primary xylem and phloem.
Although it is a single layer of cells, in actual practice it is difficult to distinguish that layer from its immediate derivatives
on either side. Hence, the term cambial zone is used (Fig. 1-14A). With few exceptions, the cambium consists of two
types of initials; the fusiform and ray initials (Fig. 1-14B-D). Fusiform initials are elongated cells that divide periclinally
and give rise to axially elongated cells in the xylem and phloem, i.e., is, tracheary cells, sieve elements, fibres, and
parenchyma cells or vertical files of parenchyma cells, called parenchyma strands. Ray initials are more or less
isodiametric and occur in clusters that appear spindle shaped in tangential sections. Ray initials give rise to xylem and
phloem rays, which extend radially into the xylem and phloem and provide for the radial transport of water, minerals,
and photoassimlate.

Sign in to download full-size image


FIGURE 1-14. (A) Cross section of a pine (Pinus sp.) stem showing the location of the vascular cambium, secondary
xylem, and secondary phloem. Tangential longitudinal sections through cambia of three woody trees, pine (B), birch
(Betula sp.) (C), and black locust (Robinia pseudo-acacia) (D), showing the arrangement and orientation of the fusiform
and ray initials. Note that in pine and birch the fusiform initials have ends that overlap with each other, whereas in black
locust they are in tiers one upon another. Cambia with the former type of arrangement of fusiform initials are referred
to as nonstoried cambia, whereas those with latter type of arrangement are referred to as storied cambia. Also note the
differences in the width and the height of rays in the three species.
20
Reproduced with permission from Arnoldia (1973).
The vascular cambium originates in roots and stems in slightly different locations (for origin in stems, see Fig. 1-1), but
eventually in woody plants it forms a complete ring—it extends up and down the stem or root like a cylindrical sheath.
How this sheath of cells with two distinct types of initials and a specific spatial arrangement comes to originate in
procambial strands has not been studied closely and the details of transition are unknown.

Procambial strands are composed of narrow elongated cells. In dicots and gymnosperms, some of these cells escape
differentiation as primary xylem or phloem cells and are left in a potentially meristematic state. Most likely, some of
these cells become committed as fusiform initials, which, likewise, are elongated cells, whereas others give rise to ray
initials after divisions. The actual process is probably more complicated and occurs over some time, but eventually
results in the conferment of a new polarity, which is unique to cambium. Cambial cells divide in a strict periclinal plane
and give rise to derivatives whose destinies are predetermined as xylem or phloem cells.

Cambium is not, however, a static cell layer placidly cutting out derivatives on each side, which differentiate as xylem
and phloem cells; rather it is a seat of constant and dynamic change in interrelationships among fusiform and ray initials.
In addition to dividing periclinally, cambial initials also divide periodically in an anticlinal plane (at right angles to the
periphery of the stem or root) to add to their numbers and thus cope with the increasing diameter of the wood cylinder,
a result of their own activity. In cambia that have been studied in detail, fusiform initials divide anticlinally with much
greater frequency than required—far more cells are produced than needed. Excess cells are converted to ray initials by
further divisions or they cease dividing and are lost from the cambial ring by differentiating as xylem or phloem cells. As
a result, interrelationships among cambial initials are constantly changing and confer upon the cambium an added
measure of plasticity. Such plasticity is useful in accommodating pathogens, such as mistletoe, which draw nutrients
from host xylem and/or phloem, or in producing more wood on one side to cope with gravity or other environmental
stresses, such as snow drifts and leaning boulders.

Plant Morphology
Michael G. Simpson, in Plant Systematics (Second Edition), 2010

Twigs, Trunks, and Buds


Twigs are the woody, recent-growth branches of trees or shrubs, and buds are immature shoot systems that develop
from meristematic regions (Figure 9.6). In many woody plants, especially those with seasonal dormancy, the outermost
leaves of the buds may develop into protective bracts (modified leaves) known as bud scales. The bud of a twig that
contains the original apical meristem of the shoot (which by later growth may result in further extension of the shoot) is
called the terminal or apical bud. Buds formed in the axils of leaves are called axillary [axial] or lateral buds.

Sign in to download full-size image


FIGURE 9.6. Twigs parts and bud types. (l.s. = longitudinal-section)

A given bud may be vegetative, if it develops into a vegetative shoot bearing leaves; floral or inflorescence, if it develops
into a flower or inflorescence; or mixed, if it develops into both flower(s) and leaves. In some species more than one
axillary bud forms per node. Two or more axillary buds that are oriented sideways are called collateral buds; two or
more axillary buds oriented vertically are called superposed buds. If the original terminal apical meristem of a shoot
aborts (e.g., by ceasing growth or maturing into a flower), then an axillary bud near the shoot apex may continue
extension growth; because this axillary bud assumes the function of a terminal bud, it is called a pseudoterminal bud.

21
Several scars may be identified on a woody, deciduous twig. These include the leaf scar, leaf vascular bundle scars,
stipule scars (if present), and bud scale scars. Bud scale scars represent the point of attachment of the bud scales of the
original terminal bud after resumption of growth during the new season. Thus, bud scale scars represent the point
where the branch ceased elongation the previous growing season; the region between adjacent bud scale scars
represents a single year's growth.

Bark technically comprises all the tissue outside the vascular cambium of a plant with true wood (see Chapter 10). The
outer bark, or periderm, are the tissues derived from the cork cambium itself. Morphologically, bark may refer to the
outermost protective tissues of the stems or roots of a plant with some sort of secondary growth, whether derived from
a true cork cambium or not. Bark types are often good identifying characteristics of plant taxa, particularly of deciduous
trees during the time that the leaves have fallen. Various bark types include:

1.
Exfoliating, a bark that cracks or splits into large sheets

2.
Fissured, a bark split or cracked into vertical or horizontal grooves

3.
Plated, a bark split or cracked, with flat plates between the fissures

4.
Shreddy, bark coarsely fibrous

5.
Smooth, a nonfibrous bark without fissures, fibers, plates, or exfoliating sheets

Genetic Engineering for Secondary Xylem Modification: Unraveling the Genetic Regulation of Wood Formation
Jae-Heung Ko, ... Kyung-Hwan Han, in Secondary Xylem Biology, 2016

Secondary growth and wood formation


During secondary growth, cell division in the vascular cambium and subsequent cell differentiation result in the
production of secondary xylem and phloem elements. The vascular cambium normally consists of 5 to 15 cambium initial
cells occurring as a continuous ring of cells between the xylem and the phloem throughout the length of fully expanded
shoots and roots (the so-called cambial zone) (Larson, 1994; Mauseth, 1998) (Fig. 10.1). Two types of initials are present
in the cambium: (1) the fusiform initials leading to the axial system and (2) the ray initials, which produce the cells that
differentiate into the system of rays throughout the wood of the stem (Lev-Yadun and Aloni, 1995). These initials serve
as a conduit for radial (across the cambium) and longitudinal (along the cambium) transfer of developmental signals and
nutrients. Adjusting to the demands of water transport required by the leaf biomass and of the mechanical strength
necessary to support the crown and to withstand wind forces (Zimmermann and Brown, 1971), cambial growth
promotes an increase in stem enlargement by the production of functional vascular elements through radial (or
anticlinal) and tangential (or periclinal) divisions (Catesson et al., 1994). Diameter growth is also coordinated with
changes in crown architecture and plant height (Larson, 1963), indicating a signaling system that integrates these growth
responses. The exact molecular mechanisms underlying the regulation of cambial growth have not been elucidated.

22

You might also like