11.12.2012 Views

The Components of Plant Tissue Culture Media II - Horticultural ...

The Components of Plant Tissue Culture Media II - Horticultural ...

The Components of Plant Tissue Culture Media II - Horticultural ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Chapter 4<br />

<strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> ll:<br />

Organic Additions, Osmotic and pH Effects,<br />

and Support Systems<br />

Growth and morphogenesis <strong>of</strong> plant tissue<br />

cultures can be improved by small amounts <strong>of</strong> some<br />

organic nutrients. <strong>The</strong>se are mainly vitamins<br />

(including some substances that are not strictly<br />

animal vitamins), amino acids and certain undefined<br />

supplements. <strong>The</strong> amount <strong>of</strong> these substances<br />

required for successful culture varies with the species<br />

and genotype, and is probably a reflection <strong>of</strong> the<br />

synthetic capacity <strong>of</strong> the explant.<br />

1.1. VITAMINS<br />

Vitamins are compounds required by animals in<br />

very small amounts as necessary ancillary food<br />

factors. Absence from the diet leads to abnormal<br />

growth and development and an unhealthy condition.<br />

Many <strong>of</strong> the same substances are also needed by plant<br />

cells as essential intermediates or metabolic catalysts,<br />

but intact plants, unlike animals, are able to produce<br />

their own requirements. <strong>Culture</strong>d plant cells and<br />

tissues can however become deficient in some<br />

factors; growth and survival is then improved by their<br />

addition to the culture medium.<br />

In early work, the requirements <strong>of</strong> tissue cultures<br />

for trace amounts <strong>of</strong> certain organic substances were<br />

satisfied by “undefined” supplements such as fruit<br />

juices, coconut milk, yeast or malt extracts and<br />

hydrolysed casein. <strong>The</strong>se supplements can contribute<br />

vitamins, amino acids and growth regulants to a<br />

culture medium. <strong>The</strong> use <strong>of</strong> undefined supplements<br />

has declined as the need for specific organic<br />

compounds has been defined, and these have become<br />

listed in catalogues as pure chemicals.<br />

1.2. THE DEVELOPMENT OF VITAMIN MIXTURES<br />

<strong>The</strong> vitamins most frequently used in plant tissue<br />

culture media are thiamine (Vit. B1), nicotinic acid<br />

(niacin) and pyridoxine (Vit. B6) and apart from these<br />

three compounds, and myo-inositol, there is little<br />

common agreement about which other vitamins are<br />

really essential.<br />

<strong>The</strong> advantage <strong>of</strong> adding thiamine was discovered<br />

almost simultaneously by Bonner (1937, 1938),<br />

Robbins and Bartley (1937) and White (1937).<br />

Nicotinic acid and pyridoxine appear, in addition to<br />

1. ORGANIC SUPPLEMENTS<br />

E. F. George et al. (eds.), <strong>Plant</strong> Propagation by <strong>Tissue</strong> <strong>Culture</strong> 3rd Edition, 115–173.<br />

© 2008 Springer.<br />

115<br />

thiamine, in media published by Bonner (1940),<br />

Gautheret (1942) and White (1943b); this was<br />

following the findings <strong>of</strong> Bonner and Devirian (1939)<br />

that nicotinic acid improved the growth <strong>of</strong> isolated<br />

roots <strong>of</strong> tomato, pea and radish; and the papers <strong>of</strong><br />

Robbins and Schmidt (1939a,b) which indicated that<br />

pyridoxine was also required for tomato root culture.<br />

<strong>The</strong>se four vitamins; myo-inositol, thiamine, nicotinic<br />

acid, and pyridoxine are ingredients <strong>of</strong> Murashige<br />

and Skoog (1962) medium and have been used in<br />

varying proportions for the culture <strong>of</strong> tissues <strong>of</strong> many<br />

plant species (Chapter 3). However, unless there has<br />

been research on the requirements <strong>of</strong> a particular<br />

plant tissue or organ, it is not possible to conclude<br />

that all the vitamins which have been used in a<br />

particular experiment were essential.<br />

<strong>The</strong> requirements <strong>of</strong> cells for added vitamins vary<br />

according to the nature <strong>of</strong> the plant and the type <strong>of</strong><br />

culture. Welander (1977) found that Nitsch and<br />

Nitsch (1965) vitamins were not necessary, or were<br />

even inhibitory to direct shoot formation on petiole<br />

explants <strong>of</strong> Begonia x hiemalis. Roest and<br />

Bokelmann (1975) on the other hand, obtained<br />

increased shoot formation on Chrysanthemum<br />

pedicels when MS vitamins were present. Callus <strong>of</strong><br />

Pinus strobus grew best when the level <strong>of</strong> inositol in<br />

MS medium was reduced to 50 mg/l whereas that <strong>of</strong><br />

P. echinata. proliferated most rapidly when no<br />

inositol was present (Kaul and Kochbar, 1985).<br />

Research workers <strong>of</strong>ten tend to adopt a ‘belt and<br />

braces’ attitude to minor media components, and add<br />

unusual supplements just to ensure that there is no<br />

missing factor which will limit the success <strong>of</strong> their<br />

experiment. Sometimes complex mixtures <strong>of</strong> as many<br />

as nine or ten vitamins have been employed.<br />

Experimentation <strong>of</strong>ten shows that some vitamins<br />

can be omitted from recommended media. Although<br />

four vitamins were used in MS medium, later work at<br />

Pr<strong>of</strong>essor Skoog’s laboratory showed that the<br />

optimum rate <strong>of</strong> growth <strong>of</strong> tobacco callus tissue on<br />

MS salts required the addition <strong>of</strong> only myo-inositol<br />

and thiamine. <strong>The</strong> level <strong>of</strong> thiamine was increased<br />

four-fold over that used by Murashige and Skoog<br />

(1962), but nicotinic acid, pyridoxine and glycine


116<br />

(amino acid) were unnecessary (Linsmaier and<br />

Skoog, 1965). A similar simplification <strong>of</strong> the MS<br />

vitamins was made by Earle and Torrey (1965) for<br />

the culture <strong>of</strong> Convolvulus callus.<br />

Soczck and Hempel (1988) found that in the<br />

medium <strong>of</strong> Murashige et al. (1974) devised for the<br />

shoot culture <strong>of</strong> Gerbera jamesonii, thiamine, pyridoxine<br />

and inositol could be omitted without any<br />

reduction in the rate <strong>of</strong> shoot multiplication <strong>of</strong> their<br />

local cultivars. Ishihara and Katano (1982) found that<br />

Malus shoot cultures could be grown on MS salts<br />

alone, and that inositol and thiamine were largely<br />

unnecessary.<br />

1.3. SPECIFIC COMPOUNDS<br />

Myo-inositol. Myo-inositol (also sometimes<br />

described as meso-inositol or i-inositol) is the only<br />

one <strong>of</strong> the nine theoretical stereoisomers <strong>of</strong> inositol<br />

which has significant biological importance.<br />

Medically it has been classed as a member <strong>of</strong> the<br />

Vitamin B complex and is required for the growth <strong>of</strong><br />

yeast and many mammalian cells in tissue culture.<br />

Rats and mice require it for hair growth and can<br />

develop dermatitis when it is not in the diet. Myoinositol<br />

has been classed as a plant ‘vitamin’, but note<br />

that some authors think that it should be regarded as a<br />

supplementary carbohydrate, although it does not<br />

contribute to carbohydrate utilization as an energy<br />

source or as an osmoticum.<br />

Historical use in tissue cultures. Myo-inositol<br />

was first shown by Jacquiot (1951) to favour bud<br />

formation by elm cambial tissue when supplied at 20-<br />

1000 mg/l. Necrosis was retarded, though the<br />

proliferation <strong>of</strong> the callus was not promoted. Myoinositol<br />

at 100 mg/1 was also used by Morel and<br />

Wetmore (1951) in combination with six other<br />

vitamins for the culture <strong>of</strong> callus from the<br />

monocotyledon Amorphophallus rivieri (Araceae).<br />

Bud initials appeared on some cultures and both roots<br />

and buds on others according to the concentration <strong>of</strong><br />

auxin employed. <strong>The</strong> vitamin was adopted by both<br />

Wood and Braun (1961) and Murashige and Skoog<br />

(1962) in combination with thiamine, nicotinic acid<br />

and pyridoxine in their preferred media fur the<br />

culture <strong>of</strong> Catharanthus roseus and Nicotiana<br />

tabacum respectively. Many other workers have since<br />

included it in culture media with favourable results<br />

on the rate <strong>of</strong> callus growth or the induction <strong>of</strong><br />

morphogenesis. Letham (1966) found that myoinositol<br />

interacted with cytokinin to promote cell<br />

division in carrot phloem explants.<br />

<strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

Occurrence and biochemistry. Part <strong>of</strong> the<br />

growth promoting property <strong>of</strong> coconut milk is due to<br />

its myo-inositol content (Pollard et al., 1961).<br />

Coconut milk also contains scyllo-inositol (Table<br />

4.1). This can also promote growth but to a smaller<br />

extent than the myo-isomer (Pollard et al., 1961).<br />

Inositol is a constituent <strong>of</strong> yeast extract (Steiner et al.,<br />

1969; Steiner and Lester, 1972) and small quantities<br />

may also be contained in commercial agar (Wolter<br />

and Skoog, 1966). Myo-inositol is a natural<br />

constituent <strong>of</strong> plants and much <strong>of</strong> it is <strong>of</strong>ten<br />

incorporated into phosphatidyl-inositol which may be<br />

an important factor in the functioning <strong>of</strong> membranes<br />

(Jung et al., 1972; Harran and Dickinson, 1978). <strong>The</strong><br />

phosphatidylinositol cycle controls various cellular<br />

responses in animal cells and yeasts, but evidence <strong>of</strong><br />

it playing a similar role in plants is only just being<br />

accumulated. Enzymes which are thought to be<br />

involved in the cycle have been observed to have<br />

activities in plants and lithium chloride (which<br />

inhibits myo-inositol-1-phosphatase and decreases the<br />

cycle) inhibits callus formation in Brassica oleracea<br />

(Bagga et al., 1987), and callus growth in<br />

Amaranthus paniculatus (Das et al., 1987). In both<br />

plants the inhibition is reversed by myo-inositol.<br />

As the myo-inositol molecule has six hydroxyl<br />

units, it can react with up to six acid molecules<br />

forming various esters. It appears that inositol<br />

phosphates act as second messengers to the primary<br />

action <strong>of</strong> auxin in plants: phytic acid (inositol hexaphosphate)<br />

is one <strong>of</strong> these. Added to culture media it<br />

can promote tissue growth if it can serve as a source<br />

<strong>of</strong> inositol (Watanabe et al., 1971). In some species,<br />

auxin can be stored and may be transported as IAAmyo-inositol<br />

ester (Chapter 5). o-Methyl-inositol is<br />

present in quite large quantities in legumes; inositol<br />

methyl ethers are known to occur in plants <strong>of</strong> several<br />

other families, although their function is unknown<br />

(Phillips and Smith, 1974).<br />

<strong>The</strong> stimulatory effect <strong>of</strong> myo-inositol in plant<br />

cultures probably arises partly from the participation<br />

<strong>of</strong> the compound in biosynthetic pathways leading to<br />

the formation <strong>of</strong> the pectin and hemicelluloses needed<br />

in cell walls (Loewus et al., 1962; Loewus, 1974;<br />

Loewus and Loewus, 1980; Harran and Dickinson,<br />

1978; Verma and Dougall, 1979; Loewus and<br />

Loewus, 1980) and may have a role in the uptake and<br />

utilization <strong>of</strong> ions (Wood and Braun, 1961). In the<br />

experiments <strong>of</strong> Staudt (1984) mentioned below, when<br />

the P04 3– content <strong>of</strong> the medium was raised to 4.41<br />

mM, the rate <strong>of</strong> callus growth <strong>of</strong> cv. ‘Aris’ was


progressively enhanced as the myo- inositol in the<br />

medium was put up to 4000 mg/l. This result seems<br />

Chapter 4<br />

to stress the importance <strong>of</strong> inositol-containing<br />

phospholipids for growth.<br />

Table 4.1. Substances identified as components <strong>of</strong> coconut milk (water) from mature green fruits and market-purchased fruits.<br />

SUBSTANCE QUANTITY/REFERENCE SUBSTANCE QUANTITY/REFERENCE<br />

Mature green<br />

fruits<br />

Mature fresh<br />

fruits<br />

Mature<br />

green<br />

fruits<br />

Mature fresh<br />

fruits<br />

Amino acids (mg/l) Sugars (g/l)<br />

Alanine 127.3 (14) 312 (13), 177.1<br />

(14)<br />

Sucrose 9.2 (14) 8.9 (14)<br />

Arginine 25.6 (14) 133 (13), 16.8<br />

(14)<br />

Glucose 7.3 (14) 2.5 (14)<br />

Aspartic acid 35.9 (14) 65 (13), 5.4 (14) Fructose 5.3 (14) 2.5 (14)<br />

Asparagine 10.1 (14) ca.60 (13), 10.1<br />

(14)<br />

Sugar alcohols (g/l)<br />

γ-Aminobutyric 34.6 (14) 820 (13), 168.8 Mannitol (1)<br />

acid<br />

(14)<br />

Glutamine acid 70.8 (14) 240 (13), 78.7<br />

(14)<br />

Sorbitol 15.0 (12), (17)<br />

Glutamine 45.4 (14) ca.60 (13), 13.4<br />

(14)<br />

myo-Inositol 0.1 (12), (17)<br />

Glycine 9.7 (14) 13.9 (14) scyllo-Inositol 0.5 (12), (17)<br />

Histidine 6.3 (14) Trace (13,14) Vitamins (mg/l)<br />

Homoserine -- (14) 5.2 (14) Nicotinic acid 0.64 (4)<br />

Hydroxyproline Trace (13,14) Pantolhenic acid 0.52 (4)<br />

Lysine 21.4 65.8 (14) Biotin, Rib<strong>of</strong>lavin 0.02 (4)<br />

Methionine 16.9 (14) 8 (13), Trace (14) Rib<strong>of</strong>lavin 0.01 (4)<br />

Phenylalanine -- (14) 12 (13), 10.2 (14) Folic acid 0.003 (4)<br />

Proline 31.9 97 (13), 21.6 (14) Thiamine, pyridoxine Trace (4)<br />

Serine 45.3 (14) Growth substances (mg/l)<br />

Typtophan 39 (13) Auxin 0.07 (7), (28)<br />

Threonine 16.2 (2) 44 (13), 26.3 (14) Gibberellin Yes (10,28)<br />

Tyrosine 6.4 (14) 16 (13), 3.1 (14) 1,3-Diphenylurea 5.8 (8), (6,17)<br />

Valine 20.6 (14) 27 (13), 15.1 (14) Zeatin (22,26)<br />

Other nitrogenous compounds Zeatin glucoside (26)<br />

Ammonium (19) Zeatin riboside (20), (24), (25)<br />

Ethanolamine (19) 6-Oxypurine growth<br />

promoter<br />

(27)<br />

Dihydroxyphenyl<br />

alanine<br />

(19) Unknown cytokinin/s 6, (18) (22)<br />

Inorganic elements (mg/100g dry wt.) Other (mg/l)<br />

Potassium 312.0 (3) RNA-polymerase (23)<br />

Sodium 105 (3) RNA-phosphorus 20.0 (14) 35.4 (14)<br />

Phosphorus 37.0 (3) DNA-phosphorus 0.1 (14) 3.5 (14)<br />

Magnesium 30.0 (3) Uracil, Adenine 21<br />

Organic acids (meq/ml) Leucoanthocyanins (11) (15,17)<br />

Malic acid 34.3 (14) 12.0 (14) Phyllococosine (16)<br />

Shikimic, Quinic<br />

and 2 unknowns<br />

0.6 (14) 0.41 (2) Acid Phosphatase (5,9)<br />

Pyrrolidone<br />

carboxylic acid<br />

0.4 (14) 0.2 (14) Diastase (2)<br />

Citric acid 0.4 (14) 0.3 (14) Dehydrogenase (5)<br />

Succinic acid -- (14) 0.3 (14) Peroxidase (5)<br />

Catalase (5)<br />

Numbered references (within brackets) in the above table are listed in Section 1.11 <strong>of</strong> this Chapter.<br />

117


118 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

Activity in tissue cultures. <strong>Culture</strong>d plant tissues<br />

vary in their capacity for myo-inositol biosynthesis.<br />

Intact shoots are usually able to produce their own<br />

requirements, but although many unorganised tissues<br />

are able to grow slowly without the vitamin being<br />

added to the medium (Murashige, 1974) the addition<br />

<strong>of</strong> a small quantity is frequently found to stimulate<br />

cell division. <strong>The</strong> compound has been discovered to<br />

be essential to some plants. In the opinion <strong>of</strong> Kaul<br />

and Sabharwal (1975) this includes all monocotyledons,<br />

the media for which, if they do not contain<br />

inositol, need to be complemented with coconut milk,<br />

or yeast extract.<br />

Fraxinus pennsylvanica callus had an absolute<br />

requirement for 10 mg/1 myo-inositol to achieve<br />

maximum growth; higher levels, up to 250 mg/l had<br />

no further effect on fresh or dry weight yields (Wolter<br />

and Skoog, 1966). <strong>The</strong> formation <strong>of</strong> shoot buds on<br />

callus <strong>of</strong> Haworthia spp was shown to be dependent<br />

on the availability <strong>of</strong> myo-inositol (Kaul and<br />

Sabharwal, 1972, 1975). In a revised Linsmaier and<br />

Skoog (1965) medium [Staudt (1984) containing 1.84<br />

mM PO4 3– ], callus tissue <strong>of</strong> Vitis vinifera cv ‘Müller-<br />

Thurgau’ did not require myo-inositol for growth, but<br />

that <strong>of</strong> Vitis vinifera x V. riparia cv. ‘Aris’ was<br />

dependent on it and the rate <strong>of</strong> growth increased as<br />

the level <strong>of</strong> myo-inositol was increased up to 250<br />

mg/l (Staudt, 1984).<br />

Gupta et al. (1988) found that it was essential to<br />

add 5 g/l myo-inositol to Gupta and Durzan (1985)<br />

DCR-1 medium to induce embryogenesis (embryonal<br />

suspensor masses) from female gametophyte tissue <strong>of</strong><br />

Pseudotsuga menziesii and Pinus taeda. <strong>The</strong><br />

concentration necessary seems insufficient to have<br />

acted as an osmotic stimulus (see section 3). myo-<br />

Inositol reduced the rate <strong>of</strong> proliferation in shoot<br />

cultures <strong>of</strong> Euphorbia fulgens (Zhang et al., 1986).<br />

Thiamine. Thiamine (Vit. B1, aneurine) in the<br />

form <strong>of</strong> thiamine pyrophosphate, is an essential c<strong>of</strong>actor<br />

in carbohydrate metabolism and is directly<br />

involved in the biosynthesis <strong>of</strong> some amino acids. It<br />

has been added to plant culture media more<br />

frequently than any other vitamin. <strong>Tissue</strong>s <strong>of</strong> most<br />

plants seem to require it for growth, the need<br />

becoming more apparent with consecutive passages,<br />

but some cultured cells are self sufficient. <strong>The</strong> maize<br />

suspension cultures <strong>of</strong> Polikarpochkina et al. (1979)<br />

showed much less growth in passage 2, and died in<br />

the third passage when thiamine was omitted from<br />

the medium.<br />

MS medium contains 0.3 μM thiamine. That this<br />

may not be sufficient to obtain optimum results from<br />

some cultures is illustrated by the results <strong>of</strong> Barwale<br />

et al. (1986): increasing the concentration <strong>of</strong><br />

thiamine-HCI in MS medium to 5 μM, increased the<br />

frequency with which zygotic embryos <strong>of</strong> Glycine<br />

max formed somatic embryos from 33% to 58%.<br />

Adding 30 μM nicotinic acid (normally 4 μM)<br />

improved the occurrence <strong>of</strong> embryogenesis even<br />

further to 76%. Thiamine was found to be essential<br />

for stimulating embryogenic callus induction in<br />

Zoysia japonica, a warm season turf grass from Japan<br />

(Asano et al., 1996). It has also been shown to<br />

stimulate adventitious rooting <strong>of</strong> Taxus spp. (Chée,<br />

1995).<br />

<strong>The</strong>re can be an interaction between thiamine and<br />

cytokinin growth regulators. Digby and Skoog (1966)<br />

discovered that normal callus cultures <strong>of</strong> tobacco<br />

produced an adequate level <strong>of</strong> thiamine to support<br />

growth providing a relatively high level <strong>of</strong> kinetin<br />

(ca. 1 mg/l) was added to the medium, but the tissue<br />

failed to grow when moved to a medium with less<br />

added kinetin unless thiamine was provided.<br />

Sometimes a change from a thiamine-requiring to<br />

a thiamine-sufficient state occurs during culture (see<br />

habituation – Chapter 7). In rice callus, thiamine<br />

influenced morphogenesis in a way that depended on<br />

which state the cells were in. Presence <strong>of</strong> the vitamin<br />

in a pre-culture (Stage I) medium caused thiaminesufficient<br />

callus to form root primordia on an<br />

induction (Stage <strong>II</strong>) medium, but suppressed the<br />

stimulating effect <strong>of</strong> kinetin on Stage <strong>II</strong> shoot<br />

formation in thiamine-requiring callus. It was<br />

essential to omit thiamine from the Stage I medium to<br />

induce thiamine-sufficient callus to produce shoots at<br />

Stage <strong>II</strong> (Inoue and Maeda, 1982).<br />

1.4. OTHER VITAMINS<br />

Pantothenic acid. Pantothenic acid plays an<br />

important role in the growth <strong>of</strong> certain tissues. It<br />

favoured callus production by hawthorn stem<br />

fragments (Morel, 1946) and stimulated tissue<br />

proliferation in willow and black henbane (Telle and<br />

Gautheret, 1947; Gautheret, 1948). However,<br />

pantothenic acid showed no effects with carrot, vine<br />

and Virginia creeper tissues which synthesize it in<br />

significant amounts (ca. 1 μg/ml).<br />

Vitamin C. <strong>The</strong> effect <strong>of</strong> Vitamin C (L-ascorbic<br />

acid) as a component <strong>of</strong> culture media will be<br />

discussed in Chapter 12. <strong>The</strong> compound is also used<br />

during explant isolation and to prevent blackening.


Besides, its role as an antioxidant, ascorbic acid is<br />

involved in cell division and elongation, e.g., in<br />

tobacco cells (de Pinto et al., 1999). Ascorbic acid<br />

(4-8 x 10 –4 M) also enhanced shoot formation in both<br />

young and old tobacco callus. (Joy et al., 1988). It<br />

speeded up the shoot-forming process, and<br />

completely reversed the inhibition <strong>of</strong> shoot formation<br />

by gibberellic acid in young callus, but was less<br />

effective in old callus. Clearly its action here was not<br />

as a vitamin.<br />

Vitamin D. Some vitamins in the D group,<br />

notably vitamin D2 and D3 can have a growth<br />

regulatory effect on plant tissue cultures. <strong>The</strong>ir effect<br />

is discussed in Chapter 7.<br />

Vitamin E. <strong>The</strong> antioxidant activity <strong>of</strong> vitamin E<br />

(α-tocopherol) will be discussed in Chapter 12.<br />

Other vitamins. Evidence has been obtained that<br />

folic acid slows tissue proliferation in the dark, while<br />

enhancing it in the light. This is probably because it<br />

is hydrolysed in the light to p-aminobenzoic acid<br />

(PAB). In the presence <strong>of</strong> auxin, PAB has been<br />

shown to have a weak growth-stimulatory effect on<br />

cultured plant tissues (de Capite, 1952a,b).<br />

Rib<strong>of</strong>lavin which is a component <strong>of</strong> some vitamin<br />

mixtures, has been found to inhibit callus formation<br />

but it may improve the growth and quality <strong>of</strong> shoots<br />

(Drew and Smith, 1986). Suppression <strong>of</strong> callus<br />

growth can mean that the vitamin may either inhibit<br />

or stimulate root formation on cuttings. Rib<strong>of</strong>lavin<br />

has been shown to stimulate adventitious rooting on<br />

shoots <strong>of</strong> Carica papaya (Drew et al., 1993), apple<br />

shoots (van der Krieken et al., 1992) and Eucalyptus<br />

globulus (Trindade and Pais, 1997). It also enhances<br />

embryogenic callus induction in Zoysia japonica in<br />

association with cytokinins and thiamine (Asano<br />

et al., 1996).<br />

Glycine is occasionally described as a vitamin in<br />

plant tissue cultures: its use has been described in the<br />

section on amino acids.<br />

Adenine. Adenine (or adenine sulphate) has been<br />

widely used in tissue culture media, but because it<br />

mainly gives rise to effects which are similar to those<br />

produced by cytokinins, it is considered in the chapter<br />

on cytokinins (Chapter 6).<br />

Stability. Some vitamins are heat-labile; see the<br />

section on medium preparation in Volume 2.<br />

1.5. UNDEFINED SUPPLEMENTS<br />

Many undefined supplements were employed in<br />

early tissue culture media. <strong>The</strong>ir use has slowly<br />

declined as the balance between inorganic salts has<br />

been improved, and as the effect <strong>of</strong> amino acids and<br />

Chapter 4<br />

119<br />

growth substances has become better understood.<br />

Nevertheless several supplements <strong>of</strong> uncertain and<br />

variable composition are still in common use.<br />

<strong>The</strong> first successful cultures <strong>of</strong> plant tissue<br />

involved the use <strong>of</strong> yeast extract (Robbins, 1922;<br />

White, 1934). Other undefined additions made to<br />

plant tissue culture media have been:<br />

• meat, malt and yeast extracts and fibrin digest;<br />

• juices, pulps and extracts from various fruits<br />

(Steward and Shantz, 1959; Ranga Swamy, 1963;<br />

Guha and Maheshwari, 1964, 1967), including those<br />

from bananas and tomatoes (La Rue, 1949);<br />

• the fluids which nourish immature zygotic<br />

embryos;<br />

• extracts <strong>of</strong> seedlings (Saalbach and Koblitz, 1978)<br />

or plant leaves (Borkird and Sink, 1983);<br />

• the extract <strong>of</strong> boiled potatoes and corn steep<br />

liquor (Fox and Miller, 1959);<br />

• plant sap or the extract <strong>of</strong> roots or rhizomes. <strong>Plant</strong><br />

roots are thought to be the main site <strong>of</strong> cytokinin<br />

synthesis in plants (Chapter 6);<br />

• protein (usually casein) hydrolysates (containing a<br />

mixture <strong>of</strong> all the amino acids present in the original<br />

protein). Casein hydrolysates are sometimes termed<br />

casamino acids: they are discussed in Chapter 3).<br />

Many <strong>of</strong> these amendments can be a source <strong>of</strong><br />

amino acids, peptides, fatty acids, carbohydrates,<br />

vitamins and plant growth substances in different<br />

concentrations. Those which have been most widely<br />

used are described below.<br />

1.6. YEAST EXTRACT.<br />

Yeast extract (YE) is used less as an ingredient <strong>of</strong><br />

plant media nowadays than in former times, when it<br />

was added as a source <strong>of</strong> amino acids and vitamins,<br />

especially inositol and thiamine (Vitamin B1) (Bonner<br />

and Addicott, l937; Robbins and Bartley, 1937). In a<br />

medium consisting only <strong>of</strong> macro- and micronutrients,<br />

the provision <strong>of</strong> yeast extract was <strong>of</strong>ten<br />

found to be essential for tissue growth (White, 1934;<br />

Robbins and Bartley, 1937). <strong>The</strong> vitamin content <strong>of</strong><br />

yeast extract distinguishes it from casein hydrolysate<br />

(CH) so that in such media CH or amino acids alone,<br />

could not be substituted for YE (Straus and La Rue,<br />

1954; Nickell and Maretzki, 1969). It was soon<br />

found that amino acids such as glycine, lysine and<br />

arginine, and vitamins such as thiamine and nicotinic<br />

acid, could serve as replacements for YE, for<br />

example in the growth <strong>of</strong> tomato roots (Skinner and<br />

Street, 1954), or sugar cane cell suspensions (Nickell<br />

and Maretzki, 1969).


120 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

<strong>The</strong> percentage <strong>of</strong> amino acids in a typical yeast<br />

extract is high (e.g. 7% amino nitrogen - Nickell and<br />

Maretzki, 1969; Bridson, 1978; Thom et al., 1981),<br />

but there is less glutamic acid than in casein or other<br />

protein hydrolysate. Malt extract contains little<br />

nitrogen (ca. 0.5% in total).<br />

Yeast extract has been typically added to media in<br />

concentrations <strong>of</strong> 0.1-1 g/l; occasionally 5, 10 and<br />

even 20 g/l (Morel and Muller, 1964) have been<br />

included. It normally only enhances growth in media<br />

containing relatively low concentrations <strong>of</strong> nitrogen,<br />

or where vitamins are lacking. Addition <strong>of</strong> 125-5000<br />

mg/l YE to MS medium completely inhibited the<br />

growth <strong>of</strong> green callus <strong>of</strong> 5 different plants whereas<br />

small quantities added to Vasil and Hildebrandt<br />

(1966) THS medium (which contained 0.6 times the<br />

quantity <strong>of</strong> NO3 – and NH4 + ions and unlike MS did<br />

not contain nicotinic acid or pyridoxine) gave more<br />

vigorous growth <strong>of</strong> carrot, endive and lettuce callus<br />

than occurred on MS. <strong>The</strong>re was still no growth <strong>of</strong><br />

parsley and tomato callus on THS medium: these<br />

tissues only grew well on unmodified MS (Vasil and<br />

Hildebrandt, 1966a,b,c).<br />

Stage I media are sometimes fortified with yeast<br />

extract to reveal the presence <strong>of</strong> micro-organisms<br />

which may have escaped decontamination<br />

procedures: it is then omitted at later stages <strong>of</strong><br />

culture.<br />

Yeast extract has been shown to have some<br />

unusual properties which may relate to its amino acid<br />

content. It elicits phytoalexin accumulation in<br />

several plant species and in Glycyrrhiza echinata<br />

suspensions it stimulated chalcone synthase activity<br />

leading to the formation <strong>of</strong> narengin (Ayabe et al.,<br />

1988). It also stimulated furomocoumarin production<br />

in Glehnia littoralis cell suspensions (Kitamura et al.,<br />

1998). On Monnier (1976, 1978) medium 1 g/l yeast<br />

extract was found to inhibit the growth <strong>of</strong> immature<br />

zygotic embryos <strong>of</strong> Linum, an effect which, when<br />

0.05 mg/1 BAP and 400 mg/l glutamine were added,<br />

induced the direct formation <strong>of</strong> adventitious embryos<br />

(Pretova and Williams, 1986).<br />

Yeast extract is now purchased directly from<br />

chemical suppliers. In the 1930s and 1940s it was<br />

prepared in the laboratory. Brink et al. (1944)<br />

macerated yeast in water which was then boiled for<br />

30 minutes and, after cooling, the starchy material<br />

was removed by centrifugation. However, Robbins<br />

and Bartley (1937) found that the active components<br />

<strong>of</strong> yeast could be extracted with 80% ethanol.<br />

1.7. POTATO EXTRACT<br />

Workers in China found that there was a sharp<br />

increase in the number <strong>of</strong> pollen plants produced<br />

from wheat anthers when they were cultured on an<br />

agar solidified medium containing only an extract <strong>of</strong><br />

boiled potatoes, 0.1 mM FeEDTA, 9% sucrose and<br />

growth regulators. Potato extract alone or potato<br />

extract combined with components <strong>of</strong> conventional<br />

culture media (Chuang et al., 1978) has since been<br />

found to provide a useful medium for the anther<br />

culture <strong>of</strong> wheat and some other cereal plants. For<br />

example, the potato medium was found to be better<br />

for the anther culture <strong>of</strong> spring wheat than the<br />

synthetic (N6) medium (McGregor and McHughen,<br />

1990). Sopory et al. (1978) obtained the initiation <strong>of</strong><br />

embryogenesis from potato anthers on potato extract<br />

alone and Lichter (1981) found it beneficial to add<br />

2.5 g/l Difco potato extract to a medium for Brassica<br />

napus anther culture, but it was omitted by Chuong<br />

and Beversdorf (1985) when they repeated this work.<br />

We are not aware <strong>of</strong> potato extract being added to<br />

media for micropropagation, apart from occasional<br />

reports <strong>of</strong> its use for orchid propagation. Sagawa and<br />

Kunisaki (1982) supplemented 1 litre <strong>of</strong> Vacin and<br />

Went (1949) medium with the extract from 100g<br />

potatoes boiled for 5 minutes, and Harvais (1982)<br />

added 5% <strong>of</strong> an extract from 200g potatoes boiled in<br />

1 litre water to his orchid medium. Of interest was<br />

the finding that potato juice treatment enabled in vitro<br />

cultures <strong>of</strong> Doritaenopsis (Orchidaceae) to recover<br />

from hyperhydricity (Zou, 1995).<br />

1.8. MALT EXTRACT<br />

Although no longer commonly used, malt extract<br />

seems to play a specific role in cultures <strong>of</strong> Citrus.<br />

Malt extract, mainly a source <strong>of</strong> carbohydrates, was<br />

shown to initiate embryogenesis in nucellar explants<br />

(Rangan et al., 1968; Rangan, 1984). Several recent<br />

studies showed a role for the extract in the<br />

multiplication <strong>of</strong> Citrus sinensis somatic embryos<br />

(Das et al., 1995), and in other Citrus spp. (Jumin,<br />

1995), in the promotion <strong>of</strong> plantlet formation from<br />

somatic embryos derived from styles <strong>of</strong> different<br />

Citrus cultivars (De Pasquale et al., 1994), and in<br />

somatic embryogenesis and plantlet regeneration<br />

from pistil thin cell layers <strong>of</strong> Citrus (Carimi et al.,<br />

1999). Malt extract also promoted germination <strong>of</strong><br />

early cotyledonary stage embryos arising from the in<br />

vitro rescue <strong>of</strong> zygotic embryos <strong>of</strong> sour orange<br />

(Carimi et al., 1998). <strong>The</strong> extract is commercially<br />

available and used at a level <strong>of</strong> 0.5 – 1 g/l.


1.9. BANANA HOMOGENATE<br />

Homogenised banana fruit is sometimes added to<br />

media for the culture <strong>of</strong> orchids and is <strong>of</strong>ten reported<br />

to promote growth. <strong>The</strong> reason for its stimulatory<br />

effect has not been explained. One suggestion<br />

mentioned earlier is that it might help to stabilise the<br />

pH <strong>of</strong> the medium. Pierik et al. (1988) found that it<br />

was slightly inhibitory to the germination <strong>of</strong><br />

Paphiopedilum ciliolare seedlings but promoted the<br />

growth <strong>of</strong> seedlings once germination had taken<br />

place.<br />

1.10. FLUIDS WHICH NOURISH EMBRYOS<br />

<strong>The</strong> liquid which is present in the embryo sac <strong>of</strong><br />

immature fruits <strong>of</strong> Aesculus (e.g. A. woerlitzensis)<br />

(Shantz and Steward, 1956, 1964; Steward and<br />

Shantz, 1959; Steward and Rao, 1970) and Juglans<br />

regia (Steward and Caplin, 1952) has been found to<br />

have a strong growth-promoting effect on some plant<br />

tissues cultured on simple media, although growth<br />

inhibition has occasionally been reported<br />

(Fonnesbech, 1972). Fluid from the immature female<br />

gametophyte <strong>of</strong> Ginkgo biloba (Steward and Caplin,<br />

1952) and extracts from the female gametophyte <strong>of</strong><br />

Pseudotsuga menziesii (Mapes and Zaerr, 1981) and<br />

immature Zea mays grains (less than two weeks after<br />

pollination) can have a similar effect. <strong>The</strong> most<br />

readily obtained fluid with this kind <strong>of</strong> activity is<br />

coconut milk (water).<br />

1.11. COCONUT MILK/WATER<br />

When added to a medium containing auxin, the<br />

liquid endosperm <strong>of</strong> Cocos nucifera fruits can induce<br />

plant cells to divide and grow rapidly. <strong>The</strong> fluid is<br />

most commonly referred to as coconut milk, although<br />

Tulecke et al. (1961) maintained that the correct<br />

English term is ‘coconut water’, because the term<br />

coconut milk also describes the white liquid obtained<br />

by grating the solid white coconut endosperm (the<br />

‘meat’) in water and this is not generally used in<br />

tissue culture media. However, in this section, both<br />

terms are used.<br />

Coconut milk was first used in tissue cultures by<br />

Van Overbeek et al. (1941, 1942) who found that its<br />

addition to a culture medium was necessary for the<br />

development <strong>of</strong> very young embryos <strong>of</strong> Datura<br />

stramonium. Gautheret (1942) found that coconut<br />

milk could be used to initiate and maintain growth in<br />

tissue cultures <strong>of</strong> several plants, and Caplin and<br />

Steward (1948) showed that callus derived from<br />

phloem tissue explants <strong>of</strong> Daucus carota roots grew<br />

much more rapidly when 15% coconut milk was<br />

Chapter 4<br />

121<br />

added to a medium containing IAA. Unlike other<br />

undefined supplements to culture media (such as<br />

yeast extract, malt extract and casein hydrolysate)<br />

coconut milk has proved harder to replace by fully<br />

defined media. <strong>The</strong> liquid has been found to be<br />

beneficial for inducing growth <strong>of</strong> both callus and<br />

suspension cultures and for the induction <strong>of</strong><br />

morphogenesis. Although commercial plant tissue<br />

culture laboratories (particularly those in temperate<br />

countries) would endeavour not to use this ingredient<br />

on account <strong>of</strong> its cost, it is still frequently employed<br />

for special purposes in research.<br />

It is possible to get callus growth on coconut milk<br />

alone (Steward et al., 1952), but normally it is added<br />

to a recognised medium. Effective stimulation only<br />

occurs when relatively large quantities are added to a<br />

medium; the incorporation <strong>of</strong> 10-15 percent by<br />

volume is quite usual. For instance, Burnet and<br />

Ibrahim (1973) found that 20% coconut milk (i.e.<br />

one-fifth <strong>of</strong> the final volume <strong>of</strong> the medium) was<br />

required for the initiation and continued growth <strong>of</strong><br />

callus tissue <strong>of</strong> various Citrus species in MS medium;<br />

Rangan (1974) has obtained improved growth <strong>of</strong><br />

Panicum miliaceum in MS medium using 2,4-D in<br />

the presence <strong>of</strong> 15% coconut milk. By contrast, Vasil<br />

and co-workers (e.g. Vasil and Vasil, 1981a,b)<br />

needed to add only 5% coconut milk to MS medium<br />

to obtain somatic embryogenesis from cereal callus<br />

and suspension cultures.<br />

Many workers try to avoid having to use coconut<br />

milk in their protocols. It is an undefined supplement<br />

whose composition can vary considerably (Swedlund<br />

and Locy, 1988). However, adding coconut milk to<br />

media <strong>of</strong>ten provides a simple way to obtain<br />

satisfactory growth or morphogenesis without the<br />

need to work out a suitably defined formulation.<br />

Suggestions that coconut milk is essential for a<br />

particular purpose need to be treated with some<br />

caution. For instance, in the culture <strong>of</strong> embryogenic<br />

callus from root and petiole explants <strong>of</strong> Daucus<br />

carota, coconut milk could be replaced satisfactorily<br />

either by adenine or kinetin, showing that it did not<br />

contribute any unique substances required for<br />

embryogenesis (Halperin and Wetherell, 1964).<br />

Preparation. Ready prepared coconut water<br />

(milk) can be purchased from some chemical<br />

suppliers, but the liquid from fresh nuts (obtained<br />

from the greengrocer) is usually perfectly adequate.<br />

One nut will usually yield at least 100 ml. <strong>The</strong> water<br />

is most simply drained from dehusked coconuts by<br />

drilling holes through two <strong>of</strong> the micropyles. Only<br />

normal uncontaminated water should be used and so


122 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

nuts should be extracted one by one, and the liquid<br />

endosperm from each examined to ascertain that it is<br />

unfermented before addition to a bulk supply. Water<br />

from green but mature coconuts may contain slightly<br />

different quantities <strong>of</strong> substances to that in the nuts<br />

purchased in the local market (Table 4.1) and has<br />

been said to be a more effective stimulant in plant<br />

media than that from ripe fruits, but Morel and<br />

Wetmore (1951) found to the contrary. Tulecke et al.<br />

(1961) discovered that the water from highly<br />

immature coconuts contained smaller quantities <strong>of</strong><br />

the substances normally present in mature nuts.<br />

References to composition <strong>of</strong> coconut water<br />

(numbers refer to citations in Table 4.1). (1)<br />

Dunstan (1906), (2) DeKruijff (1906), (3) McCance<br />

and Widdowson (1940), (4) Vandenbelt (1945), (5)<br />

Sadasivan (1951), (6) Shantz and Steward (1952), (7)<br />

Paris and Duhamet (1953), (8) Shantz and Steward<br />

(1955), (9) Wilson and Cutter (1955), (10) Radley<br />

and Dear (1958), (11) Steward and Shantz (1959),<br />

(12) Pollard et al. (1961), (13) Figures <strong>of</strong> Steward<br />

et al. (1961), given by Raghavan (1977), (14)<br />

Tulecke et al. (1961), (15) Steward and Mohan Ram<br />

(1961), (16) Kuraishi and Okumura (1961), (17)<br />

Steward (1963), (18) Zwar et al. (1963), (19) Steward<br />

et al. (1964), (20) Letham (1968), (21) Steward et al.<br />

(1969), (22) Zwar and Bruce (1970), (23) Mondal<br />

et al. (1972), (24) Letham (1974), (25) Van Staden<br />

and Drewes (1975), (26) Van Staden (1976), (27)<br />

Letham (1982), (28) Dix and Van Staden (1982).<br />

Use <strong>of</strong> Coconut water. Coconut water is usually<br />

strained through cloth and deproteinized by being<br />

heated to 80-100°C for about 10 minutes while being<br />

stirred. It is then allowed to settle and the supernatant<br />

is separated from the coagulated proteins by filtration<br />

through paper. <strong>The</strong> liquid is stored frozen at -20°C.<br />

Borkird and Sink (1983) did not boil the water from<br />

fresh ripe coconuts, but having filtered it through<br />

several layers <strong>of</strong> cheesecloth, adjusted the pH to 10<br />

with 2 N NaOH and then kept it overnight at 4°C.<br />

<strong>The</strong> following day the pH was re-adjusted to 7.0 with<br />

5 N HCl, and the preparation was refiltered before<br />

being stored frozen at –20°C.<br />

Some workers autoclave media containing<br />

coconut milk; others filter-sterilise coconut milk and<br />

add it to a medium after autoclaving has been carried<br />

out. Morel and Wetmore (1951) used filter<br />

sterilisation, but found that the milk lost its potency if<br />

stored sterile (but presumably unfrozen) for 3<br />

months. Street (1977) advocated autoclaving coconut<br />

milk after it had been boiled and filtered; it was then<br />

stored at -20°C until required.<br />

Active ingredients. <strong>The</strong> remarkable growth<br />

stimulating property <strong>of</strong> coconut milk has led to<br />

attempts to isolate and identify the active principles.<br />

This has proved to be difficult because the fractions<br />

into which coconut milk has been separated each<br />

possess only a small proportion <strong>of</strong> the total activity<br />

and the different components appear to act<br />

synergistically. Substances so far identified include<br />

amino acids, organic acids, nucleic acids, purines,<br />

sugars, sugar alcohols, vitamins, growth substances<br />

and minerals (Table 4.1). <strong>The</strong> variable nature <strong>of</strong> the<br />

product is illustrated in the table by the analytical<br />

results obtained by different authors.<br />

Auxin activity. <strong>The</strong> liquid has been found to have<br />

some auxin activity which is increased by<br />

autoclaving, probably because any such growth<br />

substances exist in a bound form and are released by<br />

hydrolysis. But although coconut milk can stimulate<br />

the growth <strong>of</strong> some in vitro cultures in the absence <strong>of</strong><br />

exogenous auxin, it normally contains little <strong>of</strong> this<br />

kind <strong>of</strong> growth regulator and an additional exogenous<br />

supply is generally required. In modern media,<br />

where organic compounds are <strong>of</strong>ten added in defined<br />

amounts, the main benefit from using coconut milk is<br />

almost certainly due to its providing highly active<br />

natural cytokinin growth substances.<br />

Cytokinin activity. Coconut milk was shown to<br />

have cytokinin activity by Kuraishi and Okumura<br />

(1961) and recognised natural cytokinin substances<br />

have since been isolated [9-β-D-ribo-furanosyl zeatin<br />

(Letham, 1968); zeatin and several unidentified ones<br />

(Zwar and Bruce, 1970); N, N′-diphenyl urea (Shantz<br />

and Steward, 1955)] but the levels <strong>of</strong> these<br />

compounds in various samples <strong>of</strong> coconut milk have<br />

not been published. An unusual cytokinin-like<br />

growth promoter, 2-(3-methylbut-2-enylamino)-purin-<br />

6-one was isolated by Letham (1982).<br />

Because coconut milk contains natural cytokinins,<br />

adding it to media <strong>of</strong>ten has the same effect as adding<br />

a recognised cytokinin. This means that a beneficial<br />

effect on growth or morphogenesis is <strong>of</strong>ten dependent<br />

on the presence <strong>of</strong> an auxin. Steward and Caplin<br />

(1951) showed that there was a synergistic action<br />

between 2,4-D and coconut milk in stimulating the<br />

growth <strong>of</strong> potato tuber tissue. Lin and Staba (1961)<br />

similarly found that coconut milk gave significantly<br />

improved callus growth on seedling explants <strong>of</strong><br />

peppermint and spearmint initiated by 2,4-D, but only<br />

slightly improved the growth initiated by the auxin 2-<br />

BTOA (2-benziothiazoleoxyacetic acid). <strong>The</strong><br />

occurrence <strong>of</strong> gibberellin-like substances in coconut<br />

milk has also been reported (Radley and Dear, 1958).


Suboptimum stimulation and inhibition. In<br />

cases where optimal concentrations <strong>of</strong> growth<br />

adjuvants have been determined, it has been found<br />

that the level <strong>of</strong> the same or analogous substances in<br />

coconut milk may be suboptimal. La Motte (1960)<br />

noted that 150 mg/l <strong>of</strong> tyrosine most effectively<br />

induced morphogenesis in tobacco callus cultures, but<br />

coconut milk added at 15% would provide only 0.96<br />

mg/l <strong>of</strong> this substance (Tulecke et al., 1961). Fresh<br />

and autoclaved coconut milk from mature nuts has<br />

proved inhibitory to growth or morphogenesis (Noh<br />

et al., 1988) in some instances. It is not known which<br />

ingredients cause the inhibition but the growth <strong>of</strong><br />

cultured embryos seems particularly liable to be<br />

prevented, suggesting that the compound responsible<br />

might be a natural dormancy-inducing factor such as<br />

abscisic acid. Van Overbeck et al. (1942, 1944)<br />

found that a factor was present in coconut milk which<br />

Organic acids can have three roles in plant culture<br />

media:<br />

• they may act as chelating agents, improving the<br />

availability <strong>of</strong> some micronutrients,<br />

• they can buffer the medium against pH change,<br />

• they may act as nutrients.<br />

A beneficial effect is largely restricted to the acids<br />

<strong>of</strong> the Krebs’ cycle. Dougall et al. (1979) found that<br />

20 mM succinate, malate or fumarate supported<br />

maximum growth <strong>of</strong> wild carrot cells when the<br />

medium was initially adjusted to pH 4.5. Although 1<br />

mM glutarate, adipate, pimelate, suberate, azelate or<br />

phthalate controlled the pH <strong>of</strong> the medium, little or<br />

no cell growth took place.<br />

2.1. USE AS BUFFERS<br />

<strong>The</strong> addition <strong>of</strong> organic acids to plant media is not<br />

a recent development. Various authors have found<br />

that some organic acids and their sodium or<br />

potassium salts stabilise the pH <strong>of</strong> hydroponic<br />

solutions (Trelease and Trelease, 1933) or in vitro<br />

media (Van Overbeek et al., 1941, 1942; Arnow<br />

et al., 1953), although it must be admitted that they<br />

are not as effective as synthetic biological buffers in<br />

this respect (see Section 5). Norstog and Smith<br />

(1963) discovered that 100 mg/l malic acid acted as<br />

an effective buffering agent in their medium for<br />

barley embryo culture and also appeared to enhance<br />

growth in the presence <strong>of</strong> glutamine and alanine.<br />

Malic acid, now at 1000 mg/l was retained in the<br />

improved Norstog (1973) Barley <strong>II</strong> medium. In the<br />

Chapter 4<br />

2. ORGANIC ACIDS<br />

123<br />

was essential for the growth <strong>of</strong> Datura stramonium<br />

embryos, but that heating the milk or allowing it to<br />

stand could lead to the release <strong>of</strong> toxic substances.<br />

<strong>The</strong>se could be removed by shaking with alcohols or<br />

ether or lead acetate precipitation. Duhamet and<br />

Mentzer (1955) isolated nine fractions <strong>of</strong> coconut<br />

milk by chromatography, and found one <strong>of</strong> these to<br />

be inhibitory to cultured crown gall tissues <strong>of</strong> black<br />

salsify when more than 10-20% coconut milk was<br />

incorporated into the medium. Norstog (1965)<br />

showed that autoclaved coconut milk could inhibit<br />

the growth <strong>of</strong> barley embryos but that filter-sterilised<br />

milk was stimulatory. Coconut water inhibited<br />

somatic embryo induction in Pinus taeda (Li and<br />

Huang, 1996) and both autoclaved or filter-sterilized<br />

coconut milk inhibited the growth <strong>of</strong> wheat embryoshoot<br />

apices (Smith, 1967).<br />

experiments <strong>of</strong> Schenk and Hildebrandt (1972) low<br />

levels <strong>of</strong> citrate and succinate ions did not impede<br />

callus growth <strong>of</strong> a wide variety <strong>of</strong> plants and<br />

appeared to be stimulatory in some species. <strong>The</strong><br />

acids were also effective buffers between pH 5 and<br />

pH 6, but autoclaving a medium containing sodium<br />

citrate or citric acid caused a substantial pH increase..<br />

2.1.1. Complexing with metals<br />

Divalent organic acids such is citric, maleic, malic<br />

and malonic (depending on species) are found in the<br />

xylem sap <strong>of</strong> plants, where together with amino acids<br />

they can complex with metal ions and assist their<br />

transport (White et al., 1981). <strong>The</strong>se acids can also<br />

be secreted from cultured cells and tissues into the<br />

growth medium and will contribute to the<br />

conditioning effect. Ojima and Ohira (1980)<br />

discovered that malic and citric acids, released into<br />

the medium by rice cells during the latter half <strong>of</strong> a<br />

passage, were able to make unchelated ferric iron<br />

available, so correcting an iron deficiency.<br />

2.1.2. Nutritional role<br />

As explained in Chapter 3, adding Krebs’ cycle<br />

organic acids to the medium can enhance the<br />

metabolism <strong>of</strong> NH4 + . Gamborg and Shyluk (1970)<br />

found that some organic acids could promote<br />

ammonium utilisation and the incorporation <strong>of</strong> small<br />

quantities <strong>of</strong> sodium pyruvate, citric, malic and<br />

fumaric acids into the medium, was one factor which<br />

enabled Kao and Michayluk (1975) to culture Vicia<br />

hajastana cells at low density. <strong>The</strong>ir mixture <strong>of</strong>


124 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

organic acid ions has been copied into many other<br />

media designed for protoplast culture. <strong>Culture</strong>s may<br />

not tolerate the addition <strong>of</strong> a large quantity <strong>of</strong> a free<br />

acid which will acidify the medium. For example,<br />

Triticale anther callus grew well on Chu et al. (1975)<br />

N6 medium supplemented with 35 mg/l <strong>of</strong> a mixture<br />

<strong>of</strong> sodium pyruvate, malic acid, fumaric acid, citric<br />

acid, but not when 100 mg/l was added (Chien and<br />

Kao, 1983). When organic anions are added to the<br />

medium from the sodium or potassium salts <strong>of</strong> an<br />

acid there are metallic cations to counterbalance the<br />

organic anions, and it seems to be possible to add<br />

larger quantities without toxicity. Five mM (1240<br />

mg/l .3H2O) potassium succinate enhanced the<br />

growth <strong>of</strong> cultured peach embryos (Ramming, 1990),<br />

and adding 15 mM (4052 mg/1 .4H2O) sodium<br />

succinate to MS medium (while also increasing the<br />

sucrose content from 3% to 6%) increased the cell<br />

volume and dry weight <strong>of</strong> Brassica nigra suspensions<br />

by 2.7 times (Molnar, 1988).<br />

Some plants seem to derive nutritional benefit<br />

from the presence <strong>of</strong> one particular organic acid.<br />

Murashige and Tucker (1969) showed that orange<br />

juice added to a medium containing MS salts<br />

promoted the growth <strong>of</strong> Citrus albedo callus. Malic<br />

and other Krebs’ cycle acids also have a similar<br />

effect; <strong>of</strong> these, citric acid produces the most<br />

pronounced growth stimulation. A concentration <strong>of</strong><br />

up to 10.4 mM can be effective (Goldschmidt, 1976;<br />

Einset, l978; Erner and Reuveni, 1981). Succulent<br />

plants, in particular those in the family Crassulaccae,<br />

such as Bryophyllum and Kalanchoe fix relatively<br />

large amounts <strong>of</strong> carbon dioxide during darkness,<br />

converting it into organic acids, <strong>of</strong> which malic acid<br />

is particularly important. <strong>The</strong> organic acids are<br />

metabolised during daylight hours. In such plants,<br />

malic acid might be expected to prove especially<br />

efficient in enhancing growth if added to a culture<br />

medium. Lassocinski (1985) has shown this to be the<br />

case in chlorophyll-deficient cacti <strong>of</strong> three genera.<br />

<strong>The</strong> addition <strong>of</strong> L-malic acid to the medium <strong>of</strong><br />

Savage et al. (1979) markedly improved the rate <strong>of</strong><br />

survival and vigour <strong>of</strong> small cacti or areoles.<br />

Organic acid (citrate, lactate, succinate, tartrate,<br />

and oxalate) pretreatment <strong>of</strong> alfalfa callus<br />

dramatically decreased the growth <strong>of</strong> callus, but<br />

increased the subsequent yield <strong>of</strong> somatic embryos<br />

and embryo development, as well as conversion to<br />

plantlets (Nichol et al., 1991). <strong>The</strong>y suggested that<br />

the acids may act in the physiological selection for<br />

embryogenic callus, by inducing preferential growth<br />

<strong>of</strong> slower-growing-compact cell aggregates compared<br />

to the faster growing friable callus.<br />

3. SUGARS -NUTRITIONAL AND REGULATORY EFFECTS<br />

Carbohydrates play an important role in in vitro<br />

cultures as an energy and carbon source, as well as an<br />

osmotic agent. In addition, carbohydrate-modulated<br />

gene expression in plants is known (Koch, 1996).<br />

<strong>Plant</strong> gene responses to changing carbohydrate status<br />

can vary markedly. Some genes are induced, some<br />

are repressed, and others minimally affected. As in<br />

microorganisms, sugar-sensitive plant genes are part<br />

<strong>of</strong> an ancient system <strong>of</strong> cellular adjustment to critical<br />

nutrient availability. However, there is no evidence<br />

that this role <strong>of</strong> carbohydrate is important in normal<br />

growth and organized development in cell<br />

cultures.3.1. Sugars as energy sources<br />

3.1.1. Carbohydrate autotrophy.<br />

Only a limited number <strong>of</strong> plant cell lines have<br />

been isolated which are autotrophic when cultured in<br />

vitro. Autotrophic cells are capable <strong>of</strong> fully<br />

supplying their own carbohydrate needs by carbon<br />

dioxide assimilation during photosynthesis<br />

(Bergmann, 1967; Tandeau de Marsac and Peaud-<br />

Lenoel, 1972a,b; Chandler et al., 1972; Anon, 1980;<br />

Larosa et al., 1981). Many autotrophic cultures have<br />

only been capable <strong>of</strong> relatively slow growth (e.g.<br />

Fukami and Hildebrandt, 1967), especially in the<br />

ambient atmosphere where the concentration <strong>of</strong><br />

carbon dioxide is low (see Chapter 12). However,<br />

since these early trials, very good progess is being<br />

made with photoautrophic shoot cultures and photoautotrophic<br />

micropropagation is now possible (Kozai,<br />

1991). Success is dependent on enriching the CO2<br />

concentrations in the vessels during the photoperiod,<br />

reducing or eliminating sugar from the medium, and<br />

optimising the in vitro environment.<br />

Nevertheless, for the normal culture <strong>of</strong> either<br />

cells, tissues or organs, it is necessary to incorporate a<br />

carbon source into the medium. Sucrose is almost<br />

universally used for micropropagation purposes as it<br />

is so generally utilisable by tissue cultures. Refined<br />

white domestic sugar is sufficiently pure for most<br />

practical purposes. <strong>The</strong> presence <strong>of</strong> sucrose in tissue<br />

culture media specifically inhibits chlorophyll<br />

formation and photosynthesis (see below) making<br />

autotrophic growth less feasible.


3.2 ALTERNATIVES TO SUCROSE<br />

3.2.1. Other Sugars.<br />

<strong>The</strong> selection <strong>of</strong> sucrose as the most suitable<br />

energy source for cultures follows many comparisons<br />

between possible alternatives. Some <strong>of</strong> the first work<br />

<strong>of</strong> this kind on the carbohydrate nutrition <strong>of</strong> plant<br />

tissue was done by Gautheret (1945) using normal<br />

carrot tissue. Sucrose was found to be the best source<br />

<strong>of</strong> carbon followed by glucose, maltose and raffinose;<br />

fructose was less effective and mannose and lactose<br />

were the least suitable. <strong>The</strong> findings <strong>of</strong> this and other<br />

work is summarized in Table 4.2. Sucrose has almost<br />

invariably been found to be the best carbohydrate;<br />

glucose is generally found to support growth equally<br />

well, and in a few plants it may result in better in<br />

Chapter 4<br />

vitro growth than sucrose, or promote organogenesis<br />

where sucrose will not; but being more expensive<br />

than sucrose, glucose will only be preferred for<br />

micropropagation where it produces clearly<br />

advantageous results.<br />

Multiplication <strong>of</strong> Alnus crispa, A. cordata and A.<br />

rubra shoot cultures was best on glucose, while that<br />

<strong>of</strong> A. glutinosa was best on sucrose (Tremblay and<br />

Lalonde, 1984; Tremblay et al., 1984; Barghchi,<br />

1988). Direct shoot formation from Capsicum annum<br />

leaf discs in a 16 h day required the presence <strong>of</strong><br />

glucose (Phillips and Hubstenberger, 1985). Glucose<br />

is required for the culture <strong>of</strong> isolated roots <strong>of</strong> wheat<br />

(Furguson, 1967) and some other monocotyledons<br />

(Bhojwani and Razdan, 1983).<br />

Table 4.2. <strong>The</strong> main sugars which can utilized by plants.<br />

<strong>The</strong> value <strong>of</strong> as sugar for carbon nutrition is indicated by the size <strong>of</strong> the type.<br />

SUGAR Reducing Capacity Products <strong>of</strong> hydrolytic/enzymatic<br />

breakdown<br />

Monosaccharides<br />

Hexoses<br />

Glucose Reducing sugar None<br />

Fructose Reducing sugar None<br />

Galactose Reducing sugar None<br />

Mannose Reducing sugar None<br />

Pentoses<br />

Arabinose Slow reduction None<br />

Ribose Slow reduction None<br />

Xylose Slow reduction None<br />

Disaccharides<br />

Sucrose Non-reducing Glucose, fructose<br />

Maltose Reducing sugar Glucose<br />

Cellobiose Reducing sugar Glucose<br />

Trehalose Non-reducing Glucose<br />

Lactose Reducing sugar Glucose, fructose<br />

Trisaccharides<br />

Raffinose Non-reducing Glucose, galactose, fructose<br />

Some other monosaccharides such as arabinose<br />

and xylose; disaccharides such as cellobiose, maltose<br />

and trehalose; and some polysaccharides; all <strong>of</strong> which<br />

are capable <strong>of</strong> being broken down to glucose and<br />

fructose (Table 4.2), can also sometimes be used as<br />

partial replacements for sucrose (Straus and LaRue,<br />

1954; Sievert and Hildebrandt, 1965; Yatazawa et al.,<br />

1967; Smith and Stone, 1973; Minocha and Halperin,<br />

1974; Zaghmout and Torres, 1985). In Phaseolus<br />

callus, Jeffs and Northcote (1967) found that sucrose<br />

could be replaced by maltose and trehalose (all three<br />

sugars have an alpha-glucosyl radical at the nonreducing<br />

end), but not by glucose or fructose alone or<br />

125<br />

in combination, or by several other different sugars.<br />

Galactose has been said to be toxic to most plant<br />

tissues; it inhibits the growth <strong>of</strong> orchids and other<br />

plants in concentrations as low as 0.01% (0.9 mM)<br />

(Ernst et al., 1971; Arditti and Ernst, 1984).<br />

However, cells can become adapted and grown on<br />

galactose, e.g., sugar cane cells (Maretzski and<br />

Thom, 1978). <strong>The</strong> key was the induction <strong>of</strong> the<br />

enzyme galactose kinase, which converts galactose to<br />

galactose-1-phosphate. More recently, other reports<br />

on galactose use have appeared. It promoted callus<br />

growth in rugosa rose, but inhibited somatic<br />

embryogenesis (Kunitake et al., 1993). Galactose


126 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

promoted early somatic embryo maturation stages in<br />

European silver fir (Schuller and Reuther, 1993).<br />

When used instead <strong>of</strong> sucrose, it improved rooting <strong>of</strong><br />

Annona squamosa microshoots (Lemos and Blake,<br />

1996). In addition, galactose has been found to<br />

reduce or overcome hyperhydricity in shoot cultures<br />

(Druart, 1988; see Volume 2). Fructose has also been<br />

reported to be effective in preventing hyperhydricity<br />

(Rugini et al., 1987).<br />

<strong>The</strong>re are some situations where fructose supports<br />

growth just as well as sucrose or glucose (Steffen<br />

et al., 1988) and occasionally it gives better results.<br />

Some orchid species have been reported to grow<br />

better on fructose than glucose (Ernst, 1967; Ernst<br />

et al., 1971; Arditti, 1979). Fructose was the best<br />

sugar for the production <strong>of</strong> adventitious shoots from<br />

Glycine max cotyledonary nodes, especially if the<br />

concentration <strong>of</strong> nutrient salts supplied was<br />

inadequate (Wright et al., 1986). Shoot and leaf<br />

growth and axillary shoot formation in Castanea<br />

shoot cultures was stimulated when sucrose was<br />

replaced by 30 g/l fructose. <strong>The</strong> growth <strong>of</strong> basal<br />

callus was reduced and it was possible to propagate<br />

from mature explants <strong>of</strong> C. crenata, although this was<br />

not possible on the same medium supplemented with<br />

sucrose (Chauvin and Salesses, 1988). However,<br />

fructose was reported to be toxic to carrot tissue if, as<br />

the sole source <strong>of</strong> carbon, it was autoclaved with<br />

White (1943a) A medium. When filter sterilized,<br />

fructose supported the growth <strong>of</strong> callus cultures<br />

which had a final weight 70% <strong>of</strong> those grown on<br />

sucrose (Pollard et al., 1961).<br />

Sucrose in culture media is usually hydrolysed<br />

totally, or partially, into the component<br />

monosaccharides glucose and fructose (see below)<br />

and so it is logical to compare the efficacy <strong>of</strong><br />

combinations <strong>of</strong> these two sugars with that <strong>of</strong><br />

sucrose. Kromer and Kukulczanka (1985) found that<br />

meristem tips <strong>of</strong> Canna indica survived better on a<br />

mixture <strong>of</strong> 25 g/l glucose plus 5 g/l fructose, than on<br />

30 g/l sucrose. Germination <strong>of</strong> Paphiopedilum orchid<br />

seeds was best on a medium containing 5g/l fructose<br />

plus 5 g/l glucose; a mixture <strong>of</strong> 7.5 g/l <strong>of</strong> each sugar<br />

was optimal for further growth <strong>of</strong> the seedlings<br />

(Pierik et al., 1988). In spite <strong>of</strong> its rapid hydrolysis to<br />

glucose and fructose, sucrose appears to have a<br />

specific stimulatory effect on embryo development in<br />

Douglas fir, that was not observed when it was<br />

replaced by the monosaccharides (Taber et al., 1998).<br />

<strong>The</strong> general superiority <strong>of</strong> sucrose over glucose<br />

for the culture <strong>of</strong> organised plant tissues such as<br />

isolated roots may be on account <strong>of</strong> the more<br />

effective translocation <strong>of</strong> sucrose to apical meristems<br />

(Butcher and Street, 1964). In addition, there could<br />

be an osmotic effect, because, from an equal weight<br />

<strong>of</strong> compound, a solution <strong>of</strong> glucose has almost twice<br />

the molarity <strong>of</strong> a sucrose solution, and will thus, in<br />

the absence <strong>of</strong> inversion <strong>of</strong> the disaccharide, induce a<br />

more negative water potential (see below).<br />

Maltose. <strong>Plant</strong> species vary in their ability to<br />

utilise unusual sugars. For instance, although<br />

Gautheret (1945) could grow carrot callus on<br />

maltose, Mathes et al. (1973) obtained only minimal<br />

growth <strong>of</strong> Acer tissue on media supplemented with<br />

this sugar. Similarly, growth <strong>of</strong> soybean tissue on<br />

maltose is normally very slow, but variant strains <strong>of</strong><br />

cells have been selected which can utilise it (Limberg<br />

et al., 1979), perhaps because the new genotypes<br />

possessed an improved capacity for its active<br />

transport. Later studies have given a more prominent<br />

role to maltose as a component <strong>of</strong> tissue culture<br />

media. Maltose serves as both a carbon source and as<br />

an osmoticum. Compared to sucrose there is a slower<br />

rate <strong>of</strong> extracellular hydrolysis, it is taken up more<br />

slowly, and hydrolysed intracellularly more slowly.<br />

Maltose led to a substantial increase in somatic<br />

embryos from Petunia anthers (Raquin, 1983). It<br />

also led to an increase in callus induction and plantlet<br />

regeneration during in vitro androgenesis <strong>of</strong><br />

hexaploid winter triticale and wheat (Karsai et al.,<br />

1994). Maltose also increased callus induction in rice<br />

microspore culture, with an acceleration <strong>of</strong> initial cell<br />

divisions (Xie et al., 1995). For barley microspore<br />

culture, the inclusion <strong>of</strong> maltose led to a higher<br />

frequency <strong>of</strong> green plants (Finnie et al., 1989).<br />

Maltose has been reported to equal or surpass sucrose<br />

in supporting embryogenesis in a number <strong>of</strong> species,<br />

including carrot (Verma and Dougall, 1977;<br />

Kinnersley and Henderson, 1988), alfalfa (Strickland<br />

et al., 1987), wild cherry (Reidiboym-Talleux et al.,<br />

1999), Malus (Daigny et al., 1996), Abies (Norgaard<br />

1997) and loblolly pine (Li et al., 1998). <strong>The</strong> number<br />

<strong>of</strong> plants regenerated from indica (Biswas and<br />

Zapata, 1993), and japonica (Jain et al., 1997) rice<br />

varieties was also greater when protoplasts were<br />

cultured with maltose rather than sucrose. Transfer<br />

from a medium containing sucrose or glucose to one<br />

supplemented with maltose has been used by Stuart<br />

et al. (1986) and Redenbaugh et al. (1987) to enhance<br />

the conversion <strong>of</strong> alfalfa embryos. Similarly, maltose<br />

led to a much higher germination rate from asparagus<br />

somatic embryos than sucrose (Kunitake et al., 1997).<br />

Lactose. <strong>The</strong> disaccharide lactose has been<br />

detected in only a few plants. When added to tissue


culture media it has been found to induce the activity<br />

<strong>of</strong> β-galactosidase enzyme which can be secreted into<br />

the medium. <strong>The</strong> hydrolysis <strong>of</strong> lactose to galactose<br />

and glucose then permits the growth <strong>of</strong> Nemesia<br />

strumosa and Petunia hybrida callus, cucumber<br />

suspensions (Hess et al., 1979; Callebaut and Motte,<br />

1988), cotton callus and cell suspensions (Mitchell<br />

et al., 1980), and Japanese morning glory callus<br />

(Hisajima and Thorpe, 1981). <strong>The</strong> key to lactose<br />

utilization in Japanese morning glory was not only<br />

the extracellular hydrolysis <strong>of</strong> this disaccharide, but<br />

the induction <strong>of</strong> galactose kinase, which prevented<br />

the accumulation <strong>of</strong> toxic galactose (Hisajima and<br />

Thorpe, 1985). Rodriguez and Lorenzo Martin<br />

(1987) found that adding 30 g/l lactose to MS<br />

medium instead <strong>of</strong> sucrose increased the number <strong>of</strong><br />

shoots produced by a Musa accuminata shoot culture,<br />

but no new shoots were produced on subsequent<br />

subculture, although they were when sucrose was<br />

present.<br />

In addition to lactose, plant cells have been shown<br />

to become adapted and then to grow on other<br />

galactose-containing oligosaccharides, including<br />

melibiose (Nickell and Maretzki, 1970; Gross et al.,<br />

1981), raffinose (Wright and Northcote, 1972;<br />

Thorpe and Laishley, 1974; Gross et al., 1981), and<br />

stachyose (Verma and Dougall, 1977; Gross et al.,<br />

1981).<br />

Corn syrups. Kinnersley and Henderson (1988)<br />

have shown that certain corn syrups can be used as<br />

carbon sources in plant culture media and that they<br />

may induce morphogenesis which is not provoked by<br />

supplementing with sucrose. Embryogenesis was<br />

induced in a 10-year old non-embryogenic cell line <strong>of</strong><br />

Daucus carota and plantlets were obtained from<br />

Nicotiana tabacum anthers by using syrups. Those<br />

used contained a mixture <strong>of</strong> glucose, maltose,<br />

maltotriose and higher polysaccharides. <strong>The</strong>ir<br />

stimulatory effect was reproduced by mixtures <strong>of</strong><br />

maltose and glucose.<br />

3.2.2. Sugar alcohols.<br />

Sugar alcohols were thought not usually to be<br />

metabolised by plant tissues and therefore<br />

unavailable as carbon sources. For this reason,<br />

mannitol and sorbitol have been frequently employed<br />

as osmotica to modify the water potential <strong>of</strong> a culture<br />

medium. In these circumstances, sufficient sucrose<br />

must also be present to supply the energy requirement<br />

<strong>of</strong> the tissues. Adding either mannitol or sorbitol to<br />

the medium may make boron unavailable (See<br />

Chapter 3).<br />

Chapter 4<br />

127<br />

Mannitol was found to be metabolised by<br />

Fraxinus tissues (Wolter and Skoog, 1966). Later,<br />

studies with carrot and tobacco suspensions and<br />

cotyledon cultures <strong>of</strong> radiata pine showed that<br />

although mannitol was taken up very slowly, it was<br />

readily metabolized (Thompson et al., 1986). Thus,<br />

this sugat alcohol is only <strong>of</strong> value as a short-term<br />

osmotic agent. In contrast, sorbitol is readily taken<br />

up and metabolized in some species. It has been<br />

found to support the growth <strong>of</strong> apple callus (Chong<br />

and Taper, 1972, 1974a,b) and that <strong>of</strong> other rosaceous<br />

plants (C<strong>of</strong>fin et al., 1976), occasionally giving rise<br />

to more vigorous growth than can be obtained on<br />

sucrose. <strong>The</strong> ability <strong>of</strong> Rosaceae to use sorbitol as a<br />

carbon source is reported to be variety dependent.<br />

Albrecht (1986) found that shoot cultures <strong>of</strong> one<br />

crabapple variety required sorbitol for growth and<br />

would not grow on sucrose; another benefited from<br />

being grown on a mixture <strong>of</strong> sorbitol and sucrose and<br />

the growth <strong>of</strong> a third suffered if any sucrose was<br />

replaced by sorbitol. <strong>The</strong> apple rootstock ‘Ottawa 3’<br />

produced abnormal shoots on sorbitol (Chong and<br />

Pua, 1985). Evidence is accumulating to show that<br />

sugar alcohols generally exhibit non-osmotic roles in<br />

regulating morphogenesis and metabolism in plants<br />

that do not produce polyols as primary photosynthetic<br />

products (Steinitz, 1999). In addition to being<br />

metabolised to varying degrees in heterotrophic<br />

cultures, such as tobacco, maize, rice, citrus and<br />

chichory, sugar alcohols stimulate specific molecular<br />

and physiological responses, where they apparently<br />

act as chemical signals.<br />

<strong>The</strong> cyclic hexahydric alcohol myo-inositol does<br />

not seem to provide a source <strong>of</strong> energy (Smith and<br />

Stone, 1973) and its beneficial effect on the growth <strong>of</strong><br />

cultured tissues when used as a supplementary<br />

nutrient must depend on its participation in<br />

biosynthetic pathways (see vitamins above).<br />

3.2.3. Starch.<br />

<strong>Culture</strong>d cells <strong>of</strong> a few plants are able to utilise<br />

starch in the growth medium and appear to do so by<br />

release <strong>of</strong> extracellular amylases (Nickell and<br />

Burkholder, 1950). Growth rates <strong>of</strong> these cultures are<br />

increased by the addition <strong>of</strong> gibberellic acid, probably<br />

because it increases the synthesis or secretion <strong>of</strong><br />

amylase enzymes (Maretzki et al., 1971, 1974).<br />

3.3. HYDROLYSIS OF SUCROSE.<br />

<strong>The</strong> remainder <strong>of</strong> this section on sugars is devoted<br />

to the apparent effects <strong>of</strong> sucrose concentration on<br />

cell differentiation and morphogenesis. Reports on


128 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

the subject should be tempered with the knowledge<br />

that some or all <strong>of</strong> the sucrose in the medium is liable<br />

to be broken down into its constituent hexose sugars,<br />

and that such inversion will also occur within plant<br />

tissues, where reducing sugar levels <strong>of</strong> at least 0.5 per<br />

cent are likely to occur (Helgeson et al., 1972).<br />

3.3.1. Autoclaving.<br />

A partial hydrolysis <strong>of</strong> sucrose takes place during<br />

the autoclaving <strong>of</strong> media (Ball, 1953; Wolter and<br />

Skoog, 1966) the extent being greater when the<br />

compound is dissolved together with other medium<br />

constituents than when it is autoclaved in pure<br />

aqueous solution (Ferguson et al., 1958). In a fungal<br />

medium, not dissimilar to a plant culture medium,<br />

Bretzl<strong>of</strong>f (1954) found that sucrose inversion during<br />

autoclaving (15 min at 15 lbs/in 2 ) was dependent on<br />

pH in the following way:<br />

pH 3.0 100%<br />

pH 3.4 75%<br />

pH 3.8 40%<br />

pH 4.2 25%<br />

pH 4.7 12.5%<br />

pH 5.0 10%<br />

pH 6.0 0%<br />

<strong>The</strong>se results suggest that the proportion <strong>of</strong><br />

sucrose hydrolysed by autoclaving media at<br />

conventional pH levels (5.5-5.8) should be negligible.<br />

Most evidence suggests that this is not the case and<br />

that 10-15% sucrose can be converted into glucose<br />

and fructose. <strong>Culture</strong>s <strong>of</strong> some plants grow better in<br />

media containing autoclaved (rather than filtersterilised)<br />

sucrose (Ball, 1953; Guha and Johri, 1966;<br />

Johri and Guha, 1963; Verma and Van Huystee,<br />

1971; White, 1932) suggesting that the cells benefit<br />

from the availability <strong>of</strong> glucose and/or fructose.<br />

However, Nitsch and Nitsch (1956) noted that<br />

glucose only supported the growth <strong>of</strong> Helianthus<br />

callus if it had been autoclaved, and Romberger and<br />

Tabor (1971) that the growth <strong>of</strong> Picea shoot apices<br />

was less when the medium contained sucrose<br />

autoclaved separately in water (or sucrose autoclaved<br />

with only the organic constituents <strong>of</strong> the medium)<br />

than when all the constituents had been autoclaved<br />

together. <strong>The</strong>y suggested that a stimulatory<br />

substance might be released when sugars are<br />

autoclaved with agar.<br />

In other species there has been no difference<br />

between the growth <strong>of</strong> cultures supplied with<br />

autoclaved or filter sterilised sucrose (Mathes et al.,<br />

1973) or growth has been less on a medium<br />

containing autoclaved instead <strong>of</strong> filter sterilised<br />

sucrose (Stehsel and Caplin, 1969).<br />

3.3.2. Enzymatic breakdown.<br />

Sucrose in the medium is also inverted into<br />

monosaccharides during the in vitro culture <strong>of</strong> plant<br />

material. This occurs by the action <strong>of</strong> invertase<br />

located in the plant cell walls (Burstrom, 1957;<br />

Yoshida et al., 1973) or by the release <strong>of</strong> extracellular<br />

enzyme (King and Street, 1977). In most cultures,<br />

inversion <strong>of</strong> sucrose into glucose and fructose takes<br />

place in the medium; but, because the secretion <strong>of</strong><br />

invertase enzymes varies, the degree to which it<br />

occurs differs from one kind <strong>of</strong> plant to another.<br />

After 28 days on a medium containing either 3 or 4%<br />

sucrose, single explants <strong>of</strong> Hemerocallis and<br />

Delphinium had used 20-30% <strong>of</strong> the sugar. Of that<br />

which remained in the medium which had supported<br />

Hemerocallis, about 45% was sucrose, while only 5%<br />

was sucrose in the media in which Delphinium had<br />

been grown. In both cases the rest <strong>of</strong> the sucrose had<br />

been inverted (Lumsden et al., 1990).<br />

<strong>The</strong> sucrose-inverting capacity <strong>of</strong> tomato root<br />

cultures was greatest in media <strong>of</strong> pH 3.6-4.7.<br />

Activity sharply declined in less acid media (Weston<br />

and Street, 1968). Helgeson et al. (1972) found that<br />

omission <strong>of</strong> IAA auxin from the medium in which<br />

tobacco callus was cultured, caused there to be a<br />

marked rise in reducing sugar due to the progressive<br />

hydrolysis <strong>of</strong> sucrose, both in the medium and in the<br />

tissue. A temporary increase in reducing sugars also<br />

occurred at the end <strong>of</strong> the lag phase when newly<br />

transferred callus pieces started to grow rapidly. It is<br />

interesting to note that cell wall invertase possessed<br />

catalytic activity in situ, whether or not tobacco tissue<br />

was grown on sucrose (Obata-Sasamoto and Thorpe,<br />

1983).<br />

In cultures <strong>of</strong> some species, uptake <strong>of</strong> sugar may<br />

depend on the prior extracellular hydrolysis <strong>of</strong> sugar.<br />

This is the case in sugar cane (Komor et al., 1981);<br />

and possibly also in Dendrobium orchids (Hew et al.,<br />

1988) and carrot (Kanabus et al., 1986). Nearly all<br />

the sucrose in suspension cultures <strong>of</strong> sugar cane and<br />

sugar beet was hydrolysed in 3 days (Zamski and<br />

Wyse, 1985) and Daucus carota suspensions have<br />

been reported to hydrolyse all <strong>of</strong> the sucrose in the<br />

medium (Thorpe, 1982) into the constituent hexoses<br />

within 3 days (Kanabus et al., 1986) or within 24 h<br />

(Dijkema et al., 1990). However, most species can<br />

take up sucrose directly as was shown through studies<br />

with asymetrically labeled 14 C-sucrose (Parr and<br />

Edelman, 1975), and metabolise it intracellularly.<br />

Within the cell, soluble invertase, sucrose phosphate<br />

synthetase and sucrose synthetase serve to hydrolyse<br />

sucrose (Thorpe, 1982). Thus, in Acer (Copping and


Street, 1972) the soluble invertase activity paralleled<br />

growth rate, while in tobacco (Thorpe and Meier,<br />

1973) and Japanese morning glory (Hisajima et al.,<br />

1978) sucrose synthetase was more important. In the<br />

last species, the change in activity <strong>of</strong> sucrose<br />

synthetase was greater than that <strong>of</strong> sucrose phosphate<br />

synthetase, an enzyme not extensively examined in<br />

cultured cells.<br />

<strong>The</strong> growth <strong>of</strong> shoots from non-dormant buds <strong>of</strong><br />

mulberry is not promoted by sucrose, only by<br />

maltose, glucose or fructose. Even though mulberry<br />

tissue hydrolysed sucrose into component<br />

monosaccharides, shoots did not develop. In the<br />

presence <strong>of</strong> 3% fructose, sucrose was actually<br />

inhibitory to shoot development at concentrations as<br />

low as 0.2% (Oka and Ohyama, 1982).<br />

3.4. UPTAKE.<br />

<strong>The</strong> uptake <strong>of</strong> sugar molecules into plant tissues<br />

appears to be partly through passive permeation and<br />

partly through active transport. <strong>The</strong> extent <strong>of</strong> the two<br />

mechanisms may vary. Active uptake is associated<br />

with the withdrawal <strong>of</strong> protons (H + ) from the<br />

medium. Charge compensation is effected by the<br />

excretion <strong>of</strong> a cation (H + or K + ) (Komor et al., 1977,<br />

1981). Glucose was taken up preferentially by carrot<br />

suspensions during the first 7 days <strong>of</strong> a 14 day<br />

passage; fructose uptake followed during days 7-9<br />

(Dijkema et al., 1990). At concentrations below 200<br />

mM, glucose was taken up more rapidly into<br />

strawberry fruit discs and protoplasts than either<br />

sucrose or fructose (Scott and Breen, 1988).<br />

3.5. EFFECTIVE CONCENTRATIONS.<br />

In most <strong>of</strong> the comparisons between the<br />

nutritional capabilities <strong>of</strong> sugars discussed above, the<br />

criterion <strong>of</strong> excellence has been the most rapid<br />

growth <strong>of</strong> unorganised callus or suspension-cultured<br />

cells. For this purpose 2-4% sucrose w/v is usually<br />

optimal. Similar concentrations are also used in<br />

media employed for micropropagation, but<br />

laboratories probably pay insufficient attention to the<br />

effects <strong>of</strong> sucrose on morphogenesis (see below) and<br />

plantlet development. Sucrose levels in culture<br />

media which result in good callus growth may not be<br />

optimal for morphogenesis, and either lower or<br />

higher levels may be more effective.<br />

<strong>The</strong> optimum concentration <strong>of</strong> sucrose to induce<br />

morphogenesis or growth differs between different<br />

genotypes, sometimes even between those which are<br />

closely related. For instance, Damiano et al. (1987)<br />

found that the concentration <strong>of</strong> sucrose necessary to<br />

Chapter 4<br />

129<br />

produce the best rate <strong>of</strong> shoot proliferation in<br />

Eucalyptus gunnii shoot cultures varied between<br />

clones. <strong>The</strong> influence <strong>of</strong> sucrose concentration on<br />

direct shoot formation from Chrysanthemum explants<br />

varied with plant cultivar (Fig 4.1).<br />

Experiments <strong>of</strong> Molnar (1988) have shown that<br />

the optimum level <strong>of</strong> sucrose may depend upon the<br />

other amendments added to a culture medium. <strong>The</strong><br />

most rapid growth <strong>of</strong> Brassica nigra suspensions on<br />

one containing MS salts (but less iron and B5<br />

vitamins) occurred when 2% sucrose was added.<br />

However, when it was supplemented with 1-4 g/l<br />

casein hydrolysate or a mixture <strong>of</strong> 3 defined amino<br />

acids, growth was increased on up to 6% sucrose. A<br />

similar result was obtained if, instead <strong>of</strong> the amino<br />

acids, 15 mM sodium succinate was added. <strong>The</strong>re<br />

was an extended growth period and the harvested dry<br />

weight <strong>of</strong> the culture was 2.8 times that on the<br />

original medium with 2% sucrose.<br />

<strong>The</strong> level <strong>of</strong> sucrose in the medium may have a<br />

direct effect on the type <strong>of</strong> morphogenesis. Thus,<br />

sucrose (87 mM) favored organogenesis, while a<br />

higher level (350 mM) favoured somatic<br />

embryogenesis from immature zygotic embryos <strong>of</strong><br />

sunflower (Jeannin et al., 1995). In vitro minicrowns<br />

<strong>of</strong> asparagus developed short, thickened storage roots<br />

at high frequencies when the sucrose concentration in<br />

the medium was increased to 6% (Conner and<br />

Falloon, 1993). Lower sucrose concentrations, even<br />

with the addition <strong>of</strong> non- or poorly metabolised<br />

carbohydrates, such as cellobiose, maltose, mannose,<br />

melibiose and sorbitol produced thin fibrous roots,<br />

indicating that the additional sucrose was nutritional<br />

rather than osmotic.<br />

<strong>The</strong> respiration rate <strong>of</strong> cultured plant tissues rises<br />

as the concentration <strong>of</strong> added sucrose or glucose is<br />

increased. In wheat callus, it was found to reach a<br />

maximum when 90 g/1 (0.263 M) was added to the<br />

medium, even though 20 g/l produced the highest rate<br />

<strong>of</strong> growth and number <strong>of</strong> adventitious shoots (Galiba<br />

and Erdei, 1986). <strong>The</strong> uptake <strong>of</strong> inorganic ions can<br />

be dependent on sugar concentration and the benefit<br />

<strong>of</strong> adding increased quantities <strong>of</strong> nutrients to a<br />

medium may not be apparent unless the amount <strong>of</strong><br />

sugar is increased at the same time (Gamborg et al.,<br />

1974).<br />

3.5.1. Cell differentiation<br />

Formation <strong>of</strong> vascular elements. Although<br />

sugars are clearly involved in the differentiation <strong>of</strong><br />

xylem and phloem elements in cultured cells, it is still<br />

uncertain whether they have a regulatory role apart<br />

from providing a carbon energy source necessary for


130 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

active cell metabolism. Sucrose is generally required<br />

to be present in addition to IAA before tracheid<br />

elements are differentiated in tissue cultures. <strong>The</strong><br />

number <strong>of</strong> both sieve and xylem elements formed<br />

[and possibly the proportion <strong>of</strong> each kind - Wetmore<br />

and Rier (1963); Rier and Beslow (1967)] depends on<br />

sucrose concentration (Aloni, 1980). In Helianthus<br />

tuberosus tuber slices, although sucrose, glucose and<br />

trehalose were best for supporting cell division and<br />

tracheid formation, maltose was only a moderately<br />

effective carbon source (Minocha and Halperin,<br />

1974). Shininger (1979) has concluded that only<br />

carbohydrates which enable significant cell division<br />

are capable <strong>of</strong> promoting tracheary element<br />

formation. <strong>The</strong> occurrence <strong>of</strong> lignin in cultured cells<br />

is not invariably associated with thickened cell walls.<br />

Sycamore suspension cultures produced large<br />

amounts <strong>of</strong> lignin when grown on a medium with<br />

abnormally high sucrose (more than 6%) and 2,4-D<br />

levels. It was deposited within the cells and released<br />

into the medium (Carceller et al., 1971).<br />

Fig. 4.1 <strong>The</strong> effect <strong>of</strong> sucrose concentration on direct adventitious shoot formation from flower pedicels <strong>of</strong> two chrysanthemum cultivars<br />

[from data <strong>of</strong> Roest and Bokelmann, 1975].<br />

Chlorophyll formation. Levels <strong>of</strong> sucrose<br />

normally used to support the growth <strong>of</strong> tissue cultures<br />

are <strong>of</strong>ten inhibitory to chlorophyll synthesis (Rier and<br />

Chen, 1964; Edelman and Hanson, 1972) but the<br />

degree <strong>of</strong> inhibition does vary according to the<br />

species <strong>of</strong> plant from which the tissue was derived.<br />

In experiments <strong>of</strong> Hildebrandt et al. (1963) and<br />

Fukami and Hildebrandt (1967) for example, carrot<br />

and rose callus had a high chlorophyll content on<br />

Hildebrandt et al. (1946) tobacco medium with 2-8%<br />

sucrose; but tissue <strong>of</strong> endive, lettuce and spinach only<br />

produced large amounts <strong>of</strong> the pigment on a medium<br />

with no added sucrose (although a small amount <strong>of</strong><br />

sugar was probably supplied by 15-16% coconut<br />

milk).<br />

Cymbidium protocorms contain high chlorophyll<br />

levels only if they are cultured on media containing<br />

0.2-0.5% sucrose. <strong>The</strong>ir degree <strong>of</strong> greening declines<br />

rapidly when they are grown on sucrose<br />

concentrations higher than this (Vanséveren-Van<br />

Espen, 1973). Likewise, orchid protocorm-like<br />

bodies will not become green and cannot develop into<br />

plantlets if sucrose is present in the medium beyond<br />

the stage <strong>of</strong> their differentiation from the explants.<br />

Where added sucrose does reduce chlorophyll<br />

formation, it is thought that the synthesis <strong>of</strong> 5aminolaevulinic<br />

acid (ALA - a precursor <strong>of</strong> the<br />

porphyrin molecules <strong>of</strong> which chlorophyll is<br />

composed) is reduced due to an inhibition <strong>of</strong> the<br />

activity <strong>of</strong> the enzyme ALA synthase (Pamplin and


Chapman, 1975). Sugars apart from sucrose are not<br />

inhibitory (Edelman and Hanson, 1972; El Hinnawiy,<br />

1974). It has been said that cells grown on sucrose<br />

for prolonged periods may permanently lose the<br />

ability to synthesise chlorophyll (Van Huystee,<br />

1977). Plastids become converted to amyloplasts<br />

packed with starch and this may change the<br />

expression <strong>of</strong> plastid DNA (Gunning and Steer, 1975)<br />

or result in a reduction <strong>of</strong> plastid RNA (Rosner et al.,<br />

1977).<br />

A small amount <strong>of</strong> photosynthesis may be carried<br />

out by cultured shoots providing they are not<br />

maintained on media containing a high concentration<br />

<strong>of</strong> sucrose. Photosynthesis increased in Rosa shoot<br />

cultures when they were grown initially on 20 or 40<br />

g/l sucrose which was decreased to 10 g/l in<br />

successive subcultures (Langford and Wainwright,<br />

1987). An increase in photosynthesis occurs when<br />

sucrose is omitted from the medium in which rooted<br />

plantlets are growing (Short et al., 1987), but these<br />

treatments are not successful in ensuring a greater<br />

survival <strong>of</strong> plantlets when they are transferred extra<br />

vitrum. A more recent study also showed the relative<br />

contribution <strong>of</strong> autotrophic and heterotrophic carbon<br />

metabolism in cultured potato plants (Wolf et al.,<br />

1998). With 8% sucrose in the medium 90% <strong>of</strong> the<br />

tissue carbon was <strong>of</strong> heterotrophic origin in lightgrown<br />

plants; while on 3% sucrose, only 50% was <strong>of</strong><br />

heterotrophic origin.<br />

3.6 STARCH ACCUMULATION AND MORPHOGSNESIS<br />

3.6.1. Starch deposition preceeding morphogenesis.<br />

Cells <strong>of</strong> callus and suspension cultures commonly<br />

accumulate starch in their plastids and it is<br />

particularly prevalent in cells at the stationary phase.<br />

Starch in cells <strong>of</strong> rice suspensions had different<br />

chemical properties to that in the endosperm <strong>of</strong> seeds<br />

(Landry and Smyth, 1988). In searching for features<br />

which might be related to later morphogenetic events<br />

in Nicotiana callus, Murashige and co-workers<br />

(Murashige and Nakano, 1968; Thorpe and<br />

Murashige, 1968a, b) noticed that starch accumulated<br />

preferentially in cells sited where shoot primordia<br />

ultimately formed. <strong>The</strong> starch is produced from<br />

sucrose supplied in the culture medium (Thorpe et al.,<br />

1986). As a result <strong>of</strong> this work, it was suggested that<br />

starch accumulation might be a prerequisite <strong>of</strong><br />

morphogenesis. In tobacco, starch presumably acts<br />

as a direct cellular reserve <strong>of</strong> the energy required for<br />

morphogenesis, because it disappears rapidly as<br />

meristenoids and shoot primordia are formed (Thorpe<br />

and Meier, 1974, 1975). Morphogenesis is an<br />

Chapter 4<br />

131<br />

energy-demanding process and callus maintained on<br />

a shoot-inducing medium does have a greater<br />

respiration rate than similar tissue kept on a noninductive<br />

medium (Thorpe and Murashige, 1970;<br />

Thorpe and Meier, 1972). Organ-forming callus <strong>of</strong><br />

tobacco has been found to accumulate starch prior to<br />

shoot or root formation, whereas callus not capable <strong>of</strong><br />

morphogenesis did not do so (Kavi Kishor and<br />

Mehta, 1982).<br />

3.6.2. Morphogenesis without starch deposition<br />

Cells <strong>of</strong> other plants which become committed to<br />

initiate organs do not necessarily accumulate starch<br />

as a preliminary to morphogenesis and it seems likely<br />

that the occurrence <strong>of</strong> this phenomenon is speciesrelated.<br />

<strong>The</strong> deposition <strong>of</strong> starch was observed as an<br />

early manifestation <strong>of</strong> organogenesis in Pinus<br />

coulteri embryos (Patel and Berlyn, 1983), but not in<br />

those <strong>of</strong> Picea abies (Von Arnold, 1987). Although<br />

zygotic embryos <strong>of</strong> the latter species immediately<br />

began to accumulate starch (particularly in the<br />

chloroplasts <strong>of</strong> cells in the cortex) when they were<br />

placed on a medium containing sucrose. It was never<br />

observed in meristematic cells from which<br />

adventitious buds developed (Von Arnold, 1987).<br />

However, if a major role for the accumulation <strong>of</strong><br />

starch prior to the initiation <strong>of</strong> organized development<br />

is for energy production, this role would be satisfied<br />

by the lipid reserves in zygotic embryos <strong>of</strong> conifers<br />

(Thorpe, 1982). Indeed, the rapid and nearly linear<br />

degradation <strong>of</strong> triglycerides during the period <strong>of</strong> high<br />

respiration during shoot initiation in excised<br />

cotyledons <strong>of</strong> radiata pine (Biondi and Thorpe, 1982;<br />

Douglas et al., 1982) would support this view.<br />

Meristematic centres in bulb scales <strong>of</strong> Nerine<br />

bowdenii can be detected as groups <strong>of</strong> cells from<br />

which starch is absent (Grootaarts et al., 1981).<br />

Starch was not accumulated in caulogenic callus <strong>of</strong><br />

Rosa persica x R. xanthina initiated from recentlyinitiated<br />

shoot cultures, but cells did accumulate<br />

starch when the shoot forming capacity <strong>of</strong> the callus<br />

was lost after more than three passages (Lloyd et al.,<br />

1988). Callus derived from barley embryos was<br />

noted to accumulate starch very rapidly and this was<br />

accompanied by a reduction in osmotic pressure<br />

within the cells (Granatek and Cockerline, 1978).<br />

Gibberellic acid, which in this plant could be used to<br />

induce shoot formation, brought about an increase in<br />

cell osmolarity.<br />

<strong>The</strong>re are some further examples where a<br />

diminution <strong>of</strong> the amount <strong>of</strong> stored carbohydrate in<br />

cultured tissues has restored or improved their


132 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

organogenetic capacity. A salt-tolerant line <strong>of</strong> alfalfa<br />

cells which showed no ability for shoot regeneration<br />

after three and a half years in culture on 3% sucrose<br />

was induced to form shoots and plantlets by being<br />

cultured for one passage <strong>of</strong> 24 days on 1% sucrose,<br />

before being returned to a medium containing 3%<br />

sucrose and a high 2,4-D level (Rains et al., 1980,<br />

and personal communication). Cells which in 3%<br />

sucrose were full <strong>of</strong> starch became starch-depleted<br />

during culture on a lower sucrose level. <strong>The</strong> number<br />

<strong>of</strong> somatic embryos formed by embryogenic<br />

‘Shamouti’ orange callus, was increased when<br />

sucrose was omitted from the medium for one<br />

passage, before being returned to Murashige and<br />

Tucker (1969) medium with 5-6% sucrose (Kochba<br />

and Button, 1974).<br />

3.6.3. Unusual sugars<br />

In some plants, unusual sugars are able to regulate<br />

morphogenesis and differentiation. Galactose stimulates<br />

embryogenesis in Citrus cultures (Kochba et al.,<br />

1978) and can enhance the maturation <strong>of</strong> alfalfa<br />

embryos. Callus <strong>of</strong> Cucumis sativus grew most<br />

rapidly on raffinose and was capable <strong>of</strong> forming roots<br />

when grown on this sugar; somatic embryos were<br />

only differentiated when the callus was cultured on<br />

sucrose (88-175 mM), but if a small amount <strong>of</strong><br />

stachyose (0.3 mM) was added to 88 mM sucrose the<br />

callus produced adventitious shoots instead (Kim and<br />

Janick, 1989). Stachyose is the major translocated<br />

carbohydrate in cucurbits.<br />

4. OSMOTIC EFFECT OF MEDIA INGREDIENTS<br />

Besides having a purely nutritive effect, solutions<br />

<strong>of</strong> inorganic salts and sugars, which compose tissue<br />

culture media, influence plant cell growth through<br />

their osmotic properties. A discussion is most<br />

conveniently accommodated at this point, as many <strong>of</strong><br />

the papers published on the subject stress the osmotic<br />

effects <strong>of</strong> added sugars.<br />

4.1. OSMOTIC AND WATER POTENTIALS: A GENERAL<br />

INTRODUCTION<br />

Water movement into and out <strong>of</strong> a plant cell is<br />

governed by the relative concentrations <strong>of</strong> dissolved<br />

substances in the external and internal solutions, and<br />

by the pressure exerted by its restraining cell wall.<br />

<strong>The</strong> manner <strong>of</strong> defining the respective forces has<br />

changed in recent years, and as both old and new<br />

terminology are found in the tissue culture literature,<br />

the following brief description may assist the reader.<br />

More detailed explanations can be found in many text<br />

books on plant physiology.<br />

In the older concept, cells were considered to take<br />

up water by suction (i.e. by exerting a negative<br />

pressure) induced by the osmotically active<br />

concentration <strong>of</strong> dissolved substances within the cell.<br />

<strong>The</strong> suction force or suction pressure (SP) was<br />

defined as that resulting from the osmotic pressure <strong>of</strong><br />

the cell sap (OPcs) minus the osmotic pressure <strong>of</strong> the<br />

external solution (OPext), and the pressure exerted on,<br />

and stretching the cell wall, turgor pressure (TP) - so<br />

called because it is at a maximum when the cells are<br />

turgid. This may be represented by the equations:<br />

SP = (OPcs - OPext) – TP or<br />

SP = OPcs - (OPext + TP).<br />

<strong>The</strong>se definitions devised by botanists, are not<br />

satisfactory thermodynamically because water should<br />

be considered to move down an energy gradient,<br />

losing energy as it does so. <strong>The</strong> term water potential<br />

(Ψ - Greek capital letter psi) is now used, and<br />

solutions <strong>of</strong> compounds in water are said to exert an<br />

osmotic potential. As the potential <strong>of</strong> pure water is<br />

defined as being zero, and dissolved substances cause<br />

it to be reduced, solutions have osmotic potentials, Ψs<br />

(or Ψπ - Greek small letter pi) which are negative in<br />

value. Water is said to move from a region <strong>of</strong> high<br />

potential (having a less negative value) to one that is<br />

lower (having a more negative value). Both osmotic<br />

pressure and osmotic potential are used as terms in<br />

tissue culture literature and are equivalent except that<br />

they are opposite in sign (i.e. an OP <strong>of</strong> +6 bar equals<br />

a Ψs <strong>of</strong> -6 bar). <strong>The</strong>rmodynamically, the cell’s turgor<br />

pressure is defined as a positive pressure potential<br />

(Ψp). <strong>The</strong> water potential <strong>of</strong> a cell (Ψcell ) is then<br />

equal to the sum <strong>of</strong> its osmotic and pressure<br />

potentials plus the force holding water in<br />

microcapillaries or bound to the cell wall matrix<br />

(Ψm):<br />

Ψ cell = Ψs + Ψp + Ψm<br />

Modern statements <strong>of</strong> the older suction pressure<br />

concept are therefore that the difference in water<br />

potential (i.e. the direction <strong>of</strong> water movement)<br />

between a cell and solution outside is given by:<br />

ΔΨ = (Ψsinside cell – Ψsoutside cell) – Ψp or<br />

ΔΨ = Ψ cell – Ψπoutside<br />

and when ΔΨ = 0 (at equilibrium)<br />

Ψπoutside = Ψ cell<br />

<strong>The</strong> force <strong>of</strong> water movement between two cells, `


‘A’ and ‘B’, <strong>of</strong> different water potential, is given<br />

by:<br />

ΔΨ = Ψ cellA – ΨcellB<br />

the direction <strong>of</strong> movement being towards the<br />

more negative water potential.<br />

<strong>The</strong> osmotic potential (pressure) <strong>of</strong> solutions is<br />

determined by their molar concentration and by<br />

temperature. <strong>The</strong> water potential <strong>of</strong> a plant tissue<br />

culture medium (Ψtcm) is equivalent to the osmotic<br />

potential <strong>of</strong> the dissolved compounds (Ψs). <strong>The</strong>re is<br />

no pressure potential but, if they are added,<br />

substances such as agar and Gelrite contribute a<br />

matric potential (Ψm):<br />

Ψtcm = Ψs + Ψm<br />

Osmotic pressure and water potential are<br />

measured in standard pressure units thus:<br />

1 bar = 0.987 atm<br />

= 10 6 dynes cm –2<br />

= 10 5 Pa (0.1 MPa = 1 bar).<br />

Whereas molarity is defined as number <strong>of</strong> gram<br />

moles <strong>of</strong> a substance in one litre <strong>of</strong> a solution (i.e. one<br />

litre <strong>of</strong> solution requires less than one litre <strong>of</strong><br />

solvent), molality is the number <strong>of</strong> gram moles <strong>of</strong><br />

solute per kilogram <strong>of</strong> solvent, and thus, unlike<br />

osmotic potential (osmolality, measured in pressure<br />

units) is independent <strong>of</strong> temperature. It is therefore<br />

more convenient to give measurements <strong>of</strong> osmotic<br />

pressure in osmolality units. <strong>The</strong> osmole (Osm) is<br />

defined as:<br />

<strong>The</strong> unit <strong>of</strong> the osmolality <strong>of</strong> a solution exerting<br />

an osmotic pressure equal to that <strong>of</strong> an ideal nondissociating<br />

substance which has a concentration<br />

<strong>of</strong> one mole <strong>of</strong> solute per kilogram <strong>of</strong> solvent.<br />

<strong>The</strong> osmolality <strong>of</strong> a very dilute solution <strong>of</strong> a<br />

substance which does not dissociate into ions, will be<br />

the same as its molality (i.e. g moles per kilogram <strong>of</strong><br />

solvent). <strong>The</strong> osmolality <strong>of</strong> a weak solution <strong>of</strong> a salt,<br />

Chapter 4<br />

or salts, which has completely dissociated into ions,<br />

will equal that <strong>of</strong> the total molality <strong>of</strong> the ions.<br />

<strong>The</strong> osmotic potential <strong>of</strong> dilute solutions<br />

approximates to Van’t H<strong>of</strong>f’s equation: Ψ = -cRT;<br />

where<br />

c = concentration <strong>of</strong> solutes in mol/litre;<br />

R = the gas constant and<br />

T = temperature in °K<br />

From the above equation, at 0°C one litre <strong>of</strong> a<br />

solution containing 1 mole <strong>of</strong> an undissociated<br />

compound, or 1 mole <strong>of</strong> ions, could be expected to<br />

have an osmolality <strong>of</strong> 1 Osm/kg, and an osmotic<br />

(water) potential <strong>of</strong>:<br />

Ψ= -1 (mole) x 0.082054 (atm /mole/°C)<br />

x 273.16 (°K) = - 22.414 atm<br />

Thus, although in practice the Van’t H<strong>of</strong>f<br />

equation must be corrected by the osmotic<br />

coefficient, Φ:<br />

Ψ = − Φ cRT (Lang, 1967),<br />

it is possible to give approximate figures for<br />

converting osmolality into osmotic potential pressure<br />

units. <strong>The</strong>se are shown in Table 4.3. This table can<br />

also be used to estimate the osmotic potential <strong>of</strong> nondissociating<br />

molecules such as sugars or mannitol.<br />

Thus at 25°C, 30 g/l sucrose (molecular wt. 342.3)<br />

should exert an osmotic potential <strong>of</strong>:<br />

30<br />

-2.4789 x = −0.217 MPa<br />

342.<br />

3<br />

<strong>The</strong> observed potential is -0.223 Mpa (Table 4.4).<br />

A definition. <strong>The</strong> osmotic properties <strong>of</strong> solutions<br />

can be difficult to describe without confusion. In this<br />

book, the addition <strong>of</strong> solutes to a solvent (which<br />

makes the osmotic or water potential more negative,<br />

but makes the osmolality <strong>of</strong> the solution increase to a<br />

larger positive value), has been said to reduce<br />

Table 4.3 Factors by which osmolality (Osm/kg) should be multiplied to estimate an equivalent osmotic potential<br />

in pressure units.<br />

Multiply Osm/kg by the factor shown for an equivalent osmotic potential in<br />

pressure units 1<br />

For<br />

conversion to<br />

15 ºC 20 ºC 0 ºC 25 ºC 30 ºC<br />

Atm -22.414 -23.645 -24.055 -24.465 -24.875<br />

Bar -22.711 -23.958 -24.374 -24.789 -25.205<br />

Dyne/cm 2<br />

-22.711<br />

x 10 6<br />

-23.958<br />

x 10 6<br />

-24.374<br />

x 10 6<br />

-24.789<br />

x 10 6<br />

-25.205<br />

x 10 6<br />

Mpa -2.2711 -2.3958 -2.4374 -2.4789 -2.5205<br />

* Divide pressure units by the figures shown to find approximate osmolality,<br />

-223<br />

e.g. − 223 kPa =<br />

1000<br />

x<br />

1<br />

- 2.<br />

4789<br />

= 0.090 Osm/kg.<br />

133


134 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

Table 4.4 <strong>The</strong> osmolality and osmotic potential <strong>of</strong> sucrose solutions at different concentrations.<br />

Sucrose concentration Osmolality Osmotic potential at<br />

(% , w/v) (mM) (Osm/kg) 25 ºC (MPa)<br />

0.5 14.61 0.015 -0.037<br />

1.0 29.21 0.030 -0.074<br />

1.5 43.82 0.045 -0.112<br />

2.0 58.43 0.060 -0.149<br />

2.5 73.04 0.075 -0.186<br />

3.0 87.64 0.090 -0.223<br />

4.0 116.86 0.121 -0.300<br />

5.0 175.28 0.186 -0.461<br />

6.0 233.71 0.253 -0.627<br />

8.0 292.14 0.324 -0.803<br />

10.0 350.57 0.396 -0.982<br />

(decrease) the osmotic potential <strong>of</strong> solutions. MS<br />

medium containing 3% sucrose (osmolality, ca. 186<br />

mOsm/kg; ca. -461 kPa at 25°C) is thus described as<br />

having a lower water potential than the medium <strong>of</strong><br />

White (1954) supplemented with 2% sucrose<br />

(osmolality, ca. 78 mOsm/kg; ca. −193 kPa at 25°C).<br />

4.2. THE OSMOTIC POTENTIAL OF TISSUE CULTURE<br />

MEDIA<br />

4.2.1. <strong>The</strong> total osmotic potential <strong>of</strong> solutes<br />

<strong>The</strong> approximate total osmotic potential <strong>of</strong> a<br />

medium due to dissolved substances, can be<br />

estimated from: Ψs = Ψsmacronutrients + Ψssugars<br />

When 3% w/v sucrose is added to Murashige and<br />

Skoog (1962) medium, the osmolality <strong>of</strong> a filter<br />

sterilised preparation rises from 0.096 to 0.186<br />

Osm/kg and at 25°C, the osmotic potential <strong>of</strong> the<br />

medium decreases from -0.237 to -0.460 MPa.<br />

Sugars are thus responsible for much <strong>of</strong> the osmotic<br />

potential <strong>of</strong> normal plant culture media. Even<br />

without any inversion to monosaccharides, the<br />

addition <strong>of</strong> 3% w/v sucrose is responsible for over four<br />

fifths <strong>of</strong> the total osmotic potential <strong>of</strong> White (1954)<br />

medium, 60% <strong>of</strong> that <strong>of</strong> Schenk and Hildebrandt<br />

(1972), and for just under one half that <strong>of</strong> MS.<br />

<strong>The</strong> contribution <strong>of</strong> the gelling agent. <strong>The</strong> water<br />

potential <strong>of</strong> media solidified with gels is more<br />

negative than that <strong>of</strong> a liquid medium, due to their<br />

matric potential, but this component is probably<br />

relatively small (Amador and Stewart, 1987). In the<br />

following sections, the matric potential <strong>of</strong> semi-solid<br />

media containing ca. 6 g/l agar has been assumed to<br />

be -0.01 MPa at 25°C, but as adding extra agar to<br />

media helps to prevent hyperhydricity, it is possible<br />

that this is an underestimate.<br />

<strong>The</strong> contribution <strong>of</strong> nutrient salts. Inorganic salts<br />

dissociate into ions when they are dissolved in water,<br />

so that the water potential <strong>of</strong> their solutions<br />

(especially weak solutions) does not depend on the<br />

molality (or molarity) <strong>of</strong> undissociated compounds,<br />

but on the molality (or molarity) <strong>of</strong> their ions. Thus a<br />

solution <strong>of</strong> KCl with a molality <strong>of</strong> 0.1, will have a<br />

theoretical osmolality <strong>of</strong> 0.2, because in solution it<br />

dissociates into 0.1 mole K + and 0.1 mole Cl – .<br />

Osmolality <strong>of</strong> a solution <strong>of</strong> mixed salts is dependent<br />

on the total molality <strong>of</strong> ions in solution.<br />

Dissociation may not be complete, especially<br />

when several different compounds are dissolved<br />

together as in plant culture media, which is a further<br />

reason why calculated predictions <strong>of</strong> water potential<br />

may be imprecise. In practice, osmotic potentials<br />

should be determined by actual measurement with an<br />

osmometer. Clearly though, osmotic potential <strong>of</strong> a<br />

culture medium is related to the concentration <strong>of</strong><br />

solutes, particularly that <strong>of</strong> the macronutrients and<br />

sugar.<br />

Of the inorganic salts in nutrient media, the<br />

macronutrients contribute most to the final osmotic<br />

(water) potential because <strong>of</strong> their greater<br />

concentration. <strong>The</strong> osmolality <strong>of</strong> these relatively<br />

dilute solutions is very similar to the total osmolarity<br />

<strong>of</strong> the constituent ions at 0°C, and can therefore be<br />

estimated from the total molarity <strong>of</strong> the macronutrient<br />

ions. Thus based on its macronutrient composition, a<br />

liquid Murashige and Skoog (1962) medium (without<br />

sugar) with a total macronutrient ion concentration <strong>of</strong><br />

95.75 mM, will have an osmolality <strong>of</strong> ca. 0.0958<br />

osmoles (Osm) per kilogram <strong>of</strong> water solvent (95.8<br />

mOsm/kg), at 25°C, an osmotic potential <strong>of</strong> ca. -<br />

0.237 MPa (237 kPa). Estimates <strong>of</strong> osmolality


derived in this way agree closely to actual<br />

measurements <strong>of</strong> osmolality or osmolarity for named<br />

media given in the papers <strong>of</strong> Yoshida et al. (1973),<br />

Kavi Kishor and Reddy (1986) and Lazzeri et al.<br />

(1988) (see Table 4.5).<br />

<strong>The</strong> contribution <strong>of</strong> sugars. <strong>The</strong> osmolality and<br />

osmotic potential <strong>of</strong> sucrose solutions can be read<br />

from Table 4.6. Those <strong>of</strong> mannitol and sorbitol<br />

solutions <strong>of</strong> equivalent molarities will be<br />

approximately comparable. At concentrations up to<br />

3% w/v, the osmolality <strong>of</strong> sucrose is close to<br />

molarity. It will be seen that the osmotic potential (in<br />

MPa) <strong>of</strong> sucrose solutions at 25°C can be roughly<br />

estimated by multiplying the % weight/volume<br />

concentration by -0.075: that <strong>of</strong> the monosaccharides<br />

fructose, glucose, mannitol and sorbitol (which have<br />

a molecular weight approximately 0.52 times that <strong>of</strong><br />

sucrose), by multiplying by -0.14.<br />

If any <strong>of</strong> the sucrose in a medium becomes<br />

hydrolysed into monosaccharides, the osmotic<br />

potential <strong>of</strong> the combined sugar components (sucrose<br />

+ glucose + fructose), (Ψssugars), will be lower (more<br />

negative) than would be estimated from Table 4.5.<br />

<strong>The</strong> effect can be seen in Table 4.6. From this data it<br />

seems that 40-50% <strong>of</strong> the sucrose added to MS<br />

medium by Lazzeri et al. (1988) was broken down<br />

into monosaccharides during autoclaving. Hydrolysis<br />

<strong>of</strong> sucrose by plant-derived invertase enzymes will<br />

also have a similar effect on osmotic potential. In<br />

some suspension cultures, all sucrose remaining in<br />

the medium is inverted within 24 h. A fully inverted<br />

sucrose solution would have almost double the<br />

negative potential <strong>of</strong> the original solution, but as the<br />

appearance <strong>of</strong> glucose and fructose by enzymatic<br />

hydrolysis usually occurs concurrently with the<br />

uptake <strong>of</strong> sugars by the tissues, it will tend to have a<br />

stabilizing effect on osmotic potential during the<br />

passage <strong>of</strong> a culture. Reported measurements <strong>of</strong> the<br />

Chapter 4<br />

osmolality <strong>of</strong> MS medium containing 3% sucrose,<br />

after autoclaving, are:<br />

230 mOsm/kg (0.65% Phytagar; Lazzeri et al.,<br />

1988)<br />

240 ± 20 mOsm/kg (0.6-0.925% agar; Scherer<br />

et al., 1988)<br />

230 ± 50 mOsm/kg (0.2-0.4% Gelrite; Scherer<br />

et al., 1988)<br />

4.2.2. Decreasing osmotic potential with other<br />

osmotica<br />

By adding soluble substances in place <strong>of</strong> some <strong>of</strong><br />

the sugar in a medium, it can be shown that sugars<br />

not only act as a carbohydrate source, but also as<br />

osmoregulants. Osmotica employed for the<br />

deliberate modification <strong>of</strong> osmotic potential, should<br />

be largely lacking other biological effects. Those<br />

most frequently selected are the sugar alcohols<br />

mannitol and sorbitol. It is assumed that plants that<br />

do not have a native pathway for sugar alcohol<br />

biosynthesis are also deficient in pathways to<br />

assimilate them. Sugar alcohols, though, are usually<br />

translocated, and may be metabolised and utilized to<br />

various degrees (Steinitz, 1999; for mannitol<br />

Lipavska and Vreugedenhil, 1996 Tian and Russell,<br />

1999; for sorbitol Pua et al., 1984). Polyethylene<br />

glycol may be more helpful as an inert<br />

nonpenetrating osmolyte although it may contain<br />

toxic contaminants (Chazen et al., 1995). Mannitol<br />

can easily penetrate cell walls, but the plasmalemma<br />

is considered to be relatively impermeable to it<br />

(Rains, 1989), whereas high-molecular-weight<br />

polyethylene glycol 4000 is too large to penetrate cell<br />

walls (Carpita et al., 1979; Rains, 1989). Thus, a<br />

nonpenetrating osmolyte cannot penetrate into the<br />

plant cells, but inhibits water uptake. Sodium<br />

sulphate and sodium chloride have also been used in<br />

some experiments.<br />

Table 4.5. Predicted osmolality and the osmolality actually observed after autoclaving (data from Lazzeri et al., 1988).<br />

Predicted osmolality<br />

(no sucrose hydrolysis)<br />

Oberved total<br />

Salts Agar Sucrose Total osmolality<br />

(mOsm/kg) (mOsm/kg) (mOsm/kg) (mOsm/kg) (mOsm/kg)<br />

MS + agar 96 4 - 100 -<br />

MS + agar + 0.5% sucrose 96 4 15 115 115<br />

MS + agar + 1.0% sucrose 96 4 30 130 140<br />

MS + agar + 2.0% sucrose 96 4 60 160 184†<br />

MS + agar + 4.0% sucrose<br />

† 161 mOsm/kg by Brown et al. (1989)<br />

96 4 121 221 276<br />

135


136 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

Table 4.6. <strong>The</strong> osmolality and osmotic potential <strong>of</strong> sucrose solutions <strong>of</strong> different concentrations.<br />

Sucrose Concentration Osmolality Osmotic potential at<br />

25°C<br />

(% w/v) mM Osm/Kg MPa<br />

0.5 14.61 0.015 -0.037<br />

1.0 29.21 0.030 -0.074<br />

1.5 43.82 0.045 -0.112<br />

2.0 58.43 0.060 -0.149<br />

2.5 73.04 0.075 -0.186<br />

3.0 87.64 0.090 -0.223<br />

4.0 116.86 0.121 -0.300<br />

6.0 175.28 0.186 -0.461<br />

8.0 233.71 0.253 -0.627<br />

10.0 292.14 0.324 -0.803<br />

12.0 350.57 0.396 -0.982<br />

4.3. EFFECTS AND USES OF OSMOLYTES IN TISSUE<br />

CULTURE MEDIA<br />

4.3.1. Protoplast isolation and culture<br />

<strong>The</strong> osmotic potential <strong>of</strong> a plant cell is counterbalanced<br />

by the pressure potential exerted by the cell<br />

wall. To safely remove the cell wall during<br />

protoplast isolation without damaging the plasma<br />

membrane, it has been found necessary to plasmolyse<br />

cells before wall-degrading enzymes are used. This is<br />

done by placing the cells in a solution <strong>of</strong> lower water<br />

potential than that <strong>of</strong> the cell.<br />

Glucose, sucrose and especially mannitol and<br />

sorbitol, are usually added to protoplast isolation<br />

media for this purpose, either singly or in<br />

combination, at a total concentration <strong>of</strong> 0.35-0.7 M.<br />

<strong>The</strong>se addenda are then retained in the subsequent<br />

protoplast culture medium, their concentration being<br />

progressively reduced as cell colonies start to grow.<br />

Tobacco cell suspensions take up only a very small<br />

amount <strong>of</strong> mannitol from solution (Thompson and<br />

Thorpe, 1981) and its effect as an osmotic agent<br />

appears to be exerted outside the cell (Thorpe, 1982).<br />

When protoplasts were isolated from Pseudotsuga<br />

and Pinus suspensions, which required a high<br />

concentration <strong>of</strong> inositol to induce embryogenesis , it<br />

was found to be essential to add 60 g/l (0.33 M) myoinositol<br />

(plus 30 g/l sucrose, 20 g/l glucose and 10 g/l<br />

sorbitol) to the isolation and culture media (Gupta<br />

et al., 1988). Mannitol and further amounts <strong>of</strong><br />

sorbitol could not serve as substitutes.<br />

4.3.2. Osmotic effects on growth<br />

Solutions <strong>of</strong> different concentrations partly exert<br />

their effect on growth and morphogenesis by their<br />

nutritional value, and partly through their varying<br />

osmotic potential. Lapeña et al (1988) estimated that<br />

three quarters <strong>of</strong> the sucrose necessary to promote the<br />

optimum rate <strong>of</strong> direct adventitious shoot formation<br />

from Digitalis obscura hypocotyls, was required to<br />

supply energy, while the surplus regulated<br />

morphogenesis osmotically.<br />

How osmotic potential influences cellular<br />

processes is still far from clear. Cells maintained in<br />

an environment with low (highly negative) osmotic<br />

potential, lose water and in consequence the water<br />

potential <strong>of</strong> the cell decreases. This brings about<br />

changes in metabolism and cells accumulate high<br />

levels <strong>of</strong> proline (Rabe, 1990). <strong>The</strong> activity <strong>of</strong> the<br />

main respiratory pathway <strong>of</strong> cells (the cytochrome<br />

pathway) is reduced in conditions <strong>of</strong> osmotic stress,<br />

in favour <strong>of</strong> an alternative oxidase system (De Klerk-<br />

Kiebert and Van der Plas, 1985). <strong>The</strong> increase <strong>of</strong> the<br />

concentration <strong>of</strong> osmolytes, may also result in high<br />

levels <strong>of</strong> the plant hormone abscisic acid, both extra<br />

vitrum and in vitro (recent reviews Zhu, 2002; Riera<br />

et al., 2005).<br />

Equilibrium between the water potential <strong>of</strong> the<br />

medium and that <strong>of</strong> Echinopsis callus, only occurred<br />

when the callus was dead. Normally the water<br />

potential <strong>of</strong> the medium was greater so that water<br />

flowed into the callus (Kirkham and Holder, 1981).<br />

Clearly this situation could not occur in media which<br />

were too concentrated, and Cleland (1977) proposed<br />

that a critical water potential needs to be established<br />

within a cell before cell expansion and cell division,<br />

can occur. <strong>The</strong> osmotic concentration <strong>of</strong> culture<br />

media could therefore be expected to influence the<br />

rate <strong>of</strong> cell division or the success <strong>of</strong> morphogenesis<br />

<strong>of</strong> the cells or tissues they support. Both the<br />

inorganic and organic components will be


contributory. <strong>The</strong> cells <strong>of</strong> many plants which are<br />

natives <strong>of</strong> sea-shores or deserts (e.g. many cacti)<br />

characteristically have a low water potential (Ψcell)<br />

and in consequence may need to be cultured in media<br />

<strong>of</strong> relatively low (highly negative) osmotic potentials<br />

(Lassocinski, 1985). Sucrose concentrations <strong>of</strong> 4.5-<br />

6% have sometimes been found to be beneficial for<br />

such plants (Sachar and Iyer, 1959; Johnson and<br />

Emino, 1979; Mauseth, 1979; Lassocinski, 1985).<br />

Above normal sucrose concentrations can <strong>of</strong>ten be<br />

beneficial in media for anther culture [e.g 13%<br />

sucrose in Gamborg et al. (1968) B5 medium -<br />

Chuong and Beversdorf, 1985], and for the culture <strong>of</strong><br />

immature embryos [e.g. 10% sucose in MS medium -<br />

Stafford and Davies (1979); 12.5% in Phillips and<br />

Collins (1979) L2. medium - Phillips et al. (1982)].<br />

If the osmotic potential <strong>of</strong> the medium does<br />

indeed influence the growth <strong>of</strong> tissue cultures, one<br />

might expect the sucrose concentration, which is<br />

optimal for growth, to vary from one medium to<br />

another, more sucrose being required in dilute media<br />

than in more concentrated ones. Evans et al. (1976)<br />

found this was so with cultures <strong>of</strong> soybean tissue.<br />

Maximum rates <strong>of</strong> callus growth were obtained in<br />

media containing either:<br />

1. 50-75% <strong>of</strong> MS basal salts + 3-4% sucrose, or,<br />

2. 75-100% <strong>of</strong> MS basal salts + 2% sucrose.<br />

Similarly Yoshida et al. (1973) obtained equally<br />

good growth rates <strong>of</strong> Nicotiana glutinosa callus with<br />

nutrient media in which<br />

1. the salts exerted -0.274 MPa and sucrose -0.223<br />

MPa, or<br />

2. the salts exerted -0.365 MPa and sucrose -0.091<br />

MPa.<br />

It was essential to add 60 g/l sucrose (i.e.<br />

Ψssucrose= -0.461 MPa at 25°C) to the ‘MEDIUM’ salts<br />

<strong>of</strong> de Fossard et al. (1974) (Ψ = -0.135 MPa) to<br />

obtain germination and seedling growth from<br />

immature zygotic embryos <strong>of</strong> tomato. However, if<br />

the ‘HIGH’ salts (Ψsmedium = -0.260 MPa) were used,<br />

60 g/l sucrose gave only slightly better growth than<br />

2.1 g/l (Ψssucrose = -0.156 MPa) (Neal and Topoleski,<br />

1983).<br />

A detailed examination <strong>of</strong> osmotic effects <strong>of</strong><br />

culture media on callus cultures was conducted by<br />

Kimball et al. (1975). Various organic substances<br />

were added to a modified Miller (1961) medium<br />

(which included 2% sucrose), to decrease the osmotic<br />

potential (Ψs) from -0.290 MPa. Surprisingly, the<br />

greatest callus growth was said to occur at -1.290 to<br />

-1.490 MPa (unusually low potentials which normally<br />

inhibit growth — see below) in the presence <strong>of</strong><br />

Chapter 4<br />

137<br />

mannitol or sorbitol, and between -1.090 to -1.290<br />

MPa when extra sucrose or glucose were added. On<br />

the standard medium, many cells <strong>of</strong> the callus were<br />

irregularly shaped; as the osmotic potential <strong>of</strong> the<br />

solution was decreased there were fewer irregularities<br />

and at about Ψs = -1.090 MPa all the cells were<br />

spherical. <strong>The</strong> percentage dry matter <strong>of</strong> cultures also<br />

increased as Ψs was decreased.<br />

Doley and Leyton (1970) found that decreasing<br />

the water (osmotic) potential <strong>of</strong> half White (1963)<br />

medium by -0.100 or -0.200 MPa through adding<br />

more sucrose (and/or polyethylene glycol), caused the<br />

rate <strong>of</strong> callus growth from the cut ends <strong>of</strong> Fraxinus<br />

stem sections to be lower than on a standard medium.<br />

At the reduced water potential, callus had suberised<br />

surfaces and grew through the activity <strong>of</strong> a vascular<br />

cambium. It also contained more lignified xylem and<br />

sclereids. At each potential there was an optimal<br />

IAA concentration for xylem differentiation.<br />

When the concentration <strong>of</strong> sucrose in a high salt<br />

medium such as MS is increased above 4-5 per cent,<br />

there begins to be a progressive inhibition <strong>of</strong> cell<br />

growth in many types <strong>of</strong> culture. This appears to be<br />

an osmotic effect because addition <strong>of</strong> other<br />

osmotically-active substances (such as mannitol and<br />

polyethyleneglycol) to the medium causes a similar<br />

response (Maretzki et al., 1972). Usually, high<br />

concentrations <strong>of</strong> sucrose are not toxic, at least not in<br />

the short term, and cell growth resumes when tissues<br />

or organs are transferred to media containing normal<br />

levels <strong>of</strong> sugar. Increase <strong>of</strong> Ψs is one method <strong>of</strong><br />

extending the shelf life <strong>of</strong> cultures. Pech and Romani<br />

(1979) found that the addition <strong>of</strong> 0.4 M mannitol to<br />

MS medium (modified organics), was able to prevent<br />

the rapid cell lysis and death which occurred when<br />

2,4-D was withdrawn from pear suspension cultures.<br />

Decreasing the osmotic potential (usually by<br />

adding mannitol) together with lowering <strong>of</strong> the<br />

temperature has been used to reduce the growth rate<br />

for preservation <strong>of</strong> valuable genotypes in vitro. This<br />

has been reported, among others, for potato (Gopal<br />

et al., 2002; Harding et al., 1997), Dioscorea alata<br />

(Borges et al., 2003) and enset (Negash et al., 2001)<br />

4.3.3. <strong>Tissue</strong> water content<br />

<strong>The</strong> water content <strong>of</strong> cultured tissues decreases as<br />

the level <strong>of</strong> sucrose in the medium is increased.<br />

Isolated embryos <strong>of</strong> barley were grown by Dunwell<br />

(1981) on MS medium in sucrose concentrations up<br />

to 12 per cent. Dry weight increased as the sucrose<br />

concentration was raised to 6 or 9 per cent and shoot<br />

length <strong>of</strong> some varieties was also greater than on 3<br />

per cent sucrose. Water content <strong>of</strong> the developing


138 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

plantlets was inversely proportional to the sucrose<br />

level.<br />

<strong>The</strong> dry weight <strong>of</strong> Lilium auratum bulbs and roots<br />

on MS medium increased as the sucrose<br />

concentration was augmented to 90 g/l but decreased<br />

dramatically with 150 g/l because growth was<br />

inhibited. <strong>The</strong> fresh weight/dry weight ratio <strong>of</strong> both<br />

kinds <strong>of</strong> tissue once again declined progressively as<br />

sucrose was added to the medium (0 - 150 g/l). In a<br />

medium containing 30 mg/l sucrose, the number and<br />

fresh weight <strong>of</strong> bulblets increased as MS salt strength<br />

was raised from one eighth to two times its normal<br />

concentration. An interaction between salt and<br />

sucrose concentrations was demonstrated in that<br />

optimum dry weight <strong>of</strong> bulbs could be obtained in<br />

either single strength MS + 120 g/l sucrose<br />

(osmolality, without sucrose inversion = ca. 492<br />

mOsm/kg; Ψs = ca. -1.22 MPa), or double strength<br />

MS + 60 g/l sucrose (osmolality, without sucrose<br />

inversion = ca. 378 mOsm/kg; Ψs = ca. -0.94 MPa)<br />

(Takayama and Misawa, 1979).<br />

4.3.4. Morphogenesis<br />

<strong>The</strong> osmotic effect <strong>of</strong> sucrose in culture solutions<br />

was well demonstrated by a series <strong>of</strong> experiments on<br />

tobacco callus by Brown et al., (1979): rates <strong>of</strong> callus<br />

growth and shoot regeneration which were optimal on<br />

culture media containing 3 per cent sucrose, could be<br />

maintained when the sugar was replaced partially by<br />

osmotically equivalent levels <strong>of</strong> mannitol. <strong>The</strong><br />

optimal (medium plus sucrose) here was between<br />

-0.4 and -0.6 MPa, and increasing sucrose levels<br />

above 3 per cent brought a progressive decrease in<br />

shoot regeneration. Similar results were obtained by<br />

Barg and Umiel (1977), but when they kept the<br />

osmotic potential <strong>of</strong> the culture solution roughly<br />

constant by additions <strong>of</strong> mannitol, the sucrose<br />

concentrations optimal for tobacco callus growth or<br />

morphogenesis were not the same (Fig. 4.2)<br />

Brown and Thorpe (1980) subsequently found<br />

that callus <strong>of</strong> Nicotiana capable <strong>of</strong> forming shoots,<br />

had a water potential (Ψ) <strong>of</strong> -0.8 MPa, while nonshoot-forming<br />

callus had a Ψ <strong>of</strong> -0.4 MPa. <strong>The</strong> two<br />

relationships were:<br />

Shoot forming callus<br />

Ψcell = Ψs +Ψp + Ψm<br />

–0.8 =–1 + 0.4 + 0 MPa<br />

Non-shoot-forming callus<br />

Ψcell = Ψs +Ψp + Ψm<br />

–0.4 = –0 + 0.3 + 0 MPa<br />

Correct water potential. An optimum rate <strong>of</strong><br />

growth and adventitious shoot formation <strong>of</strong> wheat<br />

callus occurred on MS medium containing 2%<br />

sucrose. Only a small number <strong>of</strong> shoots were<br />

produced on the medium supplemented with 1%<br />

sucrose, but if mannitol was added so that the total<br />

Ψcell = Ψs was the same as when 2% sucrose was<br />

present, the formation <strong>of</strong> adventitious shoots was<br />

stimulated (Galiba and Erdei, 1986). Very similar<br />

results were obtained by Lapeña et al. (1988). A<br />

small number <strong>of</strong> shoot buds were produced from<br />

Digitalis obscura hypocotyls on MS medium<br />

containing 1% sucrose: more than twice as many if<br />

the medium contained 2% sucrose (total Ψs given as<br />

-0.336 MPa), or 1% sucrose plus mannitol to again<br />

give a total Ψs equal to -0.336 MPa.<br />

Water potential can modify commitment.<br />

Morphogenesis can also be regulated by altering the<br />

water potential <strong>of</strong> media. Shepard and Totten (1977)<br />

found that very small (ca. 1-2 mm) calluses formed<br />

from potato mesophyll protoplasts were unable to<br />

survive in 1 or 2% sucrose, and the base <strong>of</strong> larger (5-<br />

10 mm) ones turned brown, while the upper portions<br />

turned green but formed roots and no shoots. <strong>The</strong><br />

calli became fully green only on 0.2-0.5% sucrose.<br />

At these levels shoots were formed in the presence <strong>of</strong><br />

0.2-0.3 M mannitol. When the level <strong>of</strong> mannitol was<br />

reduced to 0.05 M, the proportion <strong>of</strong> calli<br />

differentiating shoots fell from 61% to 2%. <strong>The</strong><br />

possibility <strong>of</strong> an osmotic affect was suggested<br />

because equimolar concentrations <strong>of</strong> myo-inositol<br />

were just as effective in promoting shoot<br />

regeneration.<br />

Another way to modify morphogenesis is to<br />

increase the ionic concentration <strong>of</strong> the medium. Pith<br />

phloem callus <strong>of</strong> tobacco proliferates on Zapata et al.<br />

(1983) MY1 medium supplemented with 10 –5 M IAA<br />

and 2.5 x 10 –6 M kinetin, but forms shoots on<br />

Murashige et al. (1972) medium containing 10 –5 M<br />

IAA and 10 –5 M kinetin. <strong>The</strong>se two media contain<br />

very similar macronutrients (total ionic concentrations,<br />

respectively 96 and 101mM), yet adding 0.5-<br />

1.0% sodium sulphate (additional osmolality 89-130<br />

mOsm/kg) decreased the shoot formation <strong>of</strong> callus<br />

grown on the shoot-forming medium, but increased it<br />

on the medium which previously only supported<br />

callus proliferation (Pua et al., 1985a). Callus<br />

cultured in the presence <strong>of</strong> sodium sulphate retained<br />

its shoot-producing capacity over a long period,<br />

although the effect was not permanent (Pua et al.,<br />

1985b; Chandler et al., 1987). In these experiments<br />

shoot formation was also enhanced by sodium<br />

chloride and mannitol.<br />

Increasing the level <strong>of</strong> sucrose from 1 to 3 per<br />

cent in MS medium containing 0.3 mg/l IAA,


induced tobacco callus to form shoots, while further<br />

increasing it to 6 per cent resulted in root<br />

differentiation (Rawal and Mehta, 1982; Mehta,<br />

1982). <strong>The</strong> formation <strong>of</strong> adventitious shoots from<br />

Nicotiana tabacum pith callus is inhibited on a<br />

medium with MS salts if 10-15% sucrose is added.<br />

Preferential zones <strong>of</strong> cell division and meristemoids<br />

produced in 3% sucrose then become disorganised<br />

into parenchymatous tissue (Hammersley-Straw and<br />

Thorpe, 1988).<br />

Chapter 4<br />

It should be noted that auxin which has a very<br />

important influence on the growth and<br />

morphogenesis <strong>of</strong> cultured plant cells, causes their<br />

osmotic potential to be altered (Van Overbeek, 1942;<br />

Hackett, 1952; Ketellapper, 1953). When tobacco<br />

callus is grown on a medium which promotes shoot<br />

regeneration, the cells have a greater osmotic<br />

pressure (or more negative water potential) than<br />

callus grown on a non-inductive medium (Brown and<br />

Thorpe, 1980; Brown, 1982).<br />

Fig. 4.2 <strong>The</strong> effect <strong>of</strong> sucrose concentration on the growth and morphogenesis <strong>of</strong> tobacco callus.<br />

[Drawn from data for four lines <strong>of</strong> tobacco callus in Barg and Umiel, 1977]. Callus growth = solid line. Morphogenesis = line <strong>of</strong> dashes.<br />

Scale was, 1 = No differentiation, 2 = Dark green callus with meristemoids, 3 = leafy shoots<br />

Apogamous buds on ferns. Whittier and Steeves<br />

(1960) found a very clear effect <strong>of</strong> glucose<br />

concentration on the formation <strong>of</strong> apogamous buds on<br />

prothalli <strong>of</strong> the fern Pteridium. (Apogamous buds<br />

give rise to the leafy and spore-producing generation<br />

<strong>of</strong> the plant which has the haploid genetic<br />

constitution <strong>of</strong> the prothallus). Bud formation was<br />

greatest between 2-3% glucose (optimum 2.5%).<br />

Results which confirm this observation were obtained<br />

by Menon and Lal (1972) in the moss Physcomitrium<br />

pyriforme. Here apogamous sporophytes were<br />

formed most freely in low sucrose concentrations<br />

(0.5-2%) and low light conditions (50-100 lux), and<br />

were not produced at all when prothalli were cultured<br />

in 6% sucrose or high light (5000-6000 lux). Whittier<br />

and Steeves (loc. cit.) noted that they could not obtain<br />

the same rate <strong>of</strong> apogamous bud production by using<br />

0.25% glucose plus mannitol or polyethyleneglycol;<br />

on the other hand, adding these osmotica to 2.5%<br />

glucose (so that the osmotic potential <strong>of</strong> the solution<br />

139<br />

was equivalent to one with 8% sucrose) did reduce<br />

bud formation. It therefore appeared that the<br />

stimulatory effect <strong>of</strong> glucose on morphogenesis was<br />

mainly due to its action as a respiratory substrate, but<br />

that inhibition might be caused by an excessively<br />

depressed osmotic potential.<br />

Differentiation <strong>of</strong> floral buds. Pieces <strong>of</strong> coldstored<br />

chicory root were found by Margara and<br />

Rancillac (1966) to require more sucrose (up to 68<br />

g/l; 199 mM) to form floral shoots than to produce<br />

vegetative shoots (as little as 17 g/l; 50 mM). Tran<br />

Thanh Van and co-workers (Tran Thanh Van and<br />

Trinh, 1978) have similarly shown that the specific<br />

formation <strong>of</strong> vegetative buds, flower buds, callus or<br />

roots by thin cell layers excised from tobacco stems,<br />

could be controlled by selecting appropriate<br />

concentrations <strong>of</strong> sugars and <strong>of</strong> auxin and cytokinin<br />

growth regulators.<br />

Root formation and root growth. <strong>Media</strong> <strong>of</strong> small<br />

osmotic potential are usually employed for the


140 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

induction and growth <strong>of</strong> roots on micropropagated<br />

shoots. High salt levels are frequently inhibitory to<br />

root initiation. Where such levels have been used for<br />

Stage <strong>II</strong> <strong>of</strong> shoot cultures, it is common to select a<br />

low salts medium (e.g. ¼ or ½ MS), when detached<br />

shoots are required to be rooted at Stage <strong>II</strong>I. By<br />

testing four concentrations <strong>of</strong> MS salts (quarter, half,<br />

three quarters and full strength) against four levels <strong>of</strong><br />

sucrose (1, 2, 3 and 4%), Harris and Stevenson<br />

(1979) found that correct salt concentration (½ or ¼<br />

MS) was more important than sucrose concentration<br />

for root induction on grapevine cuttings in vitro. <strong>The</strong><br />

benefit <strong>of</strong> low salt levels for root initiation may be<br />

due more to the need for a low nitrogen level, than<br />

for an increased osmotic potential. Dunstan (1982)<br />

showed that microcuttings <strong>of</strong> several tree-fruit<br />

rootstocks rooted best on MS salts, but that in media<br />

<strong>of</strong> these concentrations, the amount <strong>of</strong> added sugar<br />

was not critical, although it was essential for there to<br />

be some present. For the best rooting <strong>of</strong> Castanea, it<br />

was important to place shoots in Lloyd and McCown<br />

(1981) WPM medium containing 4% sucrose (Serres,<br />

1988).<br />

<strong>The</strong>re are reports that an excessive sugar<br />

concentration can inhibit root formation. Green<br />

cotyledons <strong>of</strong> Sinapis alba and Raphanus sativus<br />

were found by Lovell et al. (1972) to form roots in<br />

2% sucrose in the dark, but not in light <strong>of</strong> 5500 lux<br />

luminous intensity. In the light, rooting did occur if<br />

the explants were kept in water, or (to a lesser extent)<br />

if they were treated with DCMU (a chemical inhibitor<br />

<strong>of</strong> photosynthesis) before culture in 2% sucrose. <strong>The</strong><br />

authors <strong>of</strong> this paper suggested that sugars were<br />

produced within the plant tissues during<br />

photosynthesis, which, added to the sucrose absorbed<br />

from the medium, provided too great a total sugar<br />

concentration for rooting. Rahman and Blake (1988)<br />

reached the same conclusion in experiments on<br />

Artocarpus heterophyllus. When shoots <strong>of</strong> this plant<br />

were kept on a rooting medium in the dark, the<br />

number and weight <strong>of</strong> roots formed on shoots,<br />

increased with the inclusion <strong>of</strong> up to 80 g/l sucrose.<br />

<strong>The</strong> optimum sucrose concentration was 40 g/l if the<br />

shoots were grown in the light.<br />

Root formation on avocado cuttings in 0.3 MS<br />

salts [plus Linsmaier and Skoog (1965) vitamins] was<br />

satisfactory with 1.5, 3 or 6% sucrose, and only<br />

reduced when 9% sucrose was added (Pliego-Alfaro,<br />

1988).<br />

Although 4% (and occasionally 8%) sucrose has<br />

been used in media for isolated root culture, 2% has<br />

been used in the great majority <strong>of</strong> cases (Butcher and<br />

Street, 1964). In an investigation into the effects <strong>of</strong><br />

sucrose concentration on the growth <strong>of</strong> tomato roots,<br />

Street and McGregor (1952) found that although<br />

sucrose concentrations <strong>of</strong> between 1.5 and 2.5%<br />

caused the same rate <strong>of</strong> increase <strong>of</strong> root fresh weight,<br />

1.5% sucrose was optimal. It produced the best rate<br />

<strong>of</strong> growth <strong>of</strong> the main root axis, and the greatest<br />

number and total length <strong>of</strong> lateral roots.<br />

Somatic embryogenesis. <strong>The</strong> osmotic potential<br />

<strong>of</strong> a medium can influence whether somatic<br />

embryogenesis can occur and can regulate the proper<br />

development <strong>of</strong> embryos. As will be shown below, a<br />

low osmotic potential is <strong>of</strong>ten favourable, but this is<br />

not always the case. For instance immature<br />

cotyledons <strong>of</strong> Glycine max produced somatic<br />

embryos on Phillips and Collins (1979) L2 medium<br />

containing less than 2% sucrose, but not if the<br />

concentration <strong>of</strong> sugar was increased above this level<br />

(Lippmann and Lippmann, 1984).<br />

Placing tissues in solutions with high osmotic<br />

potential will cause cells to become plasmolysed,<br />

leading to the breaking <strong>of</strong> cytoplasmic<br />

interconnections between adjacent cells<br />

(plasmodesmata). Wetherell (1984) has suggested<br />

that when cells and cell groups <strong>of</strong> higher plants are<br />

isolated by this process, they become enabled to<br />

develop independently, and express their totipotency.<br />

He pointed out that the isolation <strong>of</strong> cells <strong>of</strong> lower<br />

plants induces regeneration, and plasmolysis has long<br />

been known to initiate regeneration in multicellular<br />

algae, the leaves <strong>of</strong> mosses, fern prothallia and the<br />

gemmae <strong>of</strong> liverworts (Narayanaswami and LaRue,<br />

1955; Miller, 1968). Carrot cell cultures preplasmolysed<br />

for 45 min in 0.5-1.0 M sucrose or 1.0<br />

M sorbitol gave rise to many more somatic embryos<br />

when incubated in Wetherell (1969) medium with 0.5<br />

mg/l 2,4-D than if they had not been pre-treated in<br />

this way. Moreover embryo formation was more<br />

closely synchronized. Ikeda-Iwai et al. (2003) found<br />

that in Arabidopsis a 6-12 hour treatment with 0.7 M<br />

sucrose, sorbitol or mannitol resulted in somatic<br />

embryogenesis.<br />

Callus derived from hypocotyls <strong>of</strong> Albizia<br />

richardiana, produced the greatest numbers <strong>of</strong><br />

adventitious shoots on B5 medium containing 4%<br />

sucrose, but somatic embryos grew most readily<br />

when 2% sucrose was added. At least 1% sucrose<br />

was necessary for any kind <strong>of</strong> morphogenesis to take<br />

place (Tomar and Gupta, 1988). A similar result was<br />

obtained by Ćulafić et al. (1987) with callus from<br />

axillary buds <strong>of</strong> Rumex acetosella: adventitious<br />

shoots were produced on a medium containing MS


salts and 2% sucrose (Ψs = ca. -0.39 MPa at 25°C),<br />

but embryogenesis occurred when the sucrose<br />

concentration was increased to 6% (Ψssucrose = -0.46<br />

MPa, Ψs = ca. -0.70 MPa at 25°C) or if the medium<br />

was supplemented with 2% sucrose plus 21.3 g/l<br />

mannitol or sorbitol (which together have the same<br />

osmolality as 6% sucrose).<br />

A low (highly negative) osmotic potential helps to<br />

induce somatic embryogenesis in some other plants.<br />

Adding 10-30 g/l sorbitol to Kumar et al. (1988) L-6<br />

medium (total macronutient ions 64.26 mM; 20 g/l<br />

sucrose), caused there to be a high level <strong>of</strong><br />

embryogenesis in Vigna aconitifolia suspensions and<br />

the capacity for embryogenesis to be retained in longterm<br />

cultures. <strong>The</strong> formation <strong>of</strong> somatic embryos in<br />

ovary callus <strong>of</strong> Fuchsia hybrida was accelerated by<br />

adding 5% sucrose to B5 medium (Dabin and Beguin,<br />

1987), and the induction <strong>of</strong> embryogenic callus <strong>of</strong><br />

Euphorbia longan required the culture <strong>of</strong> young<br />

leaflets on B5 medium with 6% sucrose (Litz, 1988).<br />

<strong>The</strong>re are exceptions, particularly with regard to<br />

embryo growth. <strong>The</strong> proportion <strong>of</strong> Ipomoea batatas<br />

somatic embryos forming shoots was greatest when a<br />

medium containing MS inorganics contained 1.6%,<br />

rather than 3% sucrose (Chée et al., 1990). Protocorm<br />

proliferation <strong>of</strong> orchids is most rapid when tissue is<br />

cultured in high concentrations <strong>of</strong> sucrose, but for<br />

plantlet growth, the level <strong>of</strong> sucrose must be reduced<br />

(Homès and Vanseveran-Van Espen, 1973).<br />

<strong>The</strong> induction <strong>of</strong> embryogenic callus from<br />

immature seed embryos <strong>of</strong> Zea mays was best on MS<br />

medium with 12% sucrose (Lu et al., 1982), and Ho<br />

and Vasil (1983) used 6-10% sucrose in MS medium<br />

to promote the formation <strong>of</strong> pro-embryoids from<br />

young leaves <strong>of</strong> Saccharum <strong>of</strong>ficinarum. However,<br />

in the experiments <strong>of</strong> Ahloowalia and Maretski<br />

(1983), somatic embryo formation from callus <strong>of</strong> this<br />

plant was best on MS medium with 3% sucrose, but<br />

growth <strong>of</strong> the embryos into complete plantlets<br />

required that the embryos should be cultured first on<br />

MS medium with 6% sucrose and then on MS with<br />

3% sucrose.<br />

Polyethylene glycol 4000 (PEG 4000) improves<br />

root and shoot emergence without limiting embryo<br />

histodifferentiation in soybean somatic embryos<br />

(Walker and Parrott, 2001). Likewise in spruce, it<br />

was reported that polyethylene glycol might improve<br />

the quality <strong>of</strong> somatic embryos by promoting normal<br />

differentiation <strong>of</strong> the embryonic shoot and root (e.g.<br />

Stasolla et al., 2003). Non-penetrating osmotica like<br />

polyethylene glycol cannot enter plant cells, but<br />

restrict water uptake and provide a simulated drought<br />

Chapter 4<br />

141<br />

stress during embryo development. A combination <strong>of</strong><br />

ABA and an osmoticum prevents precocious<br />

germination in white spruce (Attree et al., 1991)<br />

and allows embryo development to proceed.<br />

Advantageous effects <strong>of</strong> polyethylene glycol and<br />

ABA have been reported in a number <strong>of</strong> species<br />

(Hevea brasiliensis, Linossier et al., 1997; Picia<br />

abies, Bozhkov and Von Arnold, 1998; white spruce,<br />

Stasolla et al., 2003, Corydalis yanhusuo, Sagare<br />

et al., 2000; and Panax ginseng, Langhansová et al.,<br />

2004).<br />

When cultured on Sears and Deckard (1982)<br />

medium, embryogenesis in callus initiated from<br />

immature embryos <strong>of</strong> ‘Chinese Spring’ and some<br />

other varieties <strong>of</strong> wheat was incomplete, because<br />

shoot apices germinated and grew before embryos<br />

had properly formed. More typical somatic embryos<br />

could be obtained by adding 40 mM sodium or<br />

potassium chloride to the medium. <strong>The</strong> salts had to<br />

be removed to allow plantlets to develop normally<br />

(Galiba and Yamada, 1988). Transferring somatic<br />

embryos to a medium <strong>of</strong> lower (more negative) water<br />

potential is <strong>of</strong>ten necessary to ensure their further<br />

growth and/or germination. High sucrose levels are<br />

<strong>of</strong>ten required in media for the culture <strong>of</strong> zygotic<br />

embryos if they are isolated when immature. <strong>The</strong> use<br />

<strong>of</strong> 50-120 g/l sucrose in media is then reported, the<br />

higher concentrations usually being added to very<br />

weak salt mixtures. Embryos which are more fully<br />

developed when excised, grow satisfactorily in a<br />

medium with 10-30 g/l sugar.<br />

Storage organ formation. At high<br />

concentration, sucrose promotes the formation <strong>of</strong><br />

tubers, bulbs and corms (e.g. Xu et al., 1998;<br />

Vreugdenhil et al., 1998; Ziv 2005, Gerrits and de<br />

Klerk 1992). This promotion might be mediated by<br />

ABA since osmotic stress induces ABA synthesis<br />

(Riera et al., 2005) and ABA promotes bulb (Kim<br />

et al., 1994) and tuber (Xu et al., 1998) formation.<br />

<strong>The</strong> situation is, however, more complex. Suttle and<br />

Hultstrand (1994) did not find a reduction <strong>of</strong> tuber<br />

formation in potato by adding fluridone, an ABAsynthesis<br />

inhibitor, and Xu et al. (1998) did not<br />

observe an increased ABA-level at high sucrose<br />

concentration. So in potato, the effect <strong>of</strong> sucrose is<br />

not likely to be mediated by ABA. Exogenous ABA<br />

does not promote Gladiolus corm formation (Dantu<br />

and Bhojwani, 1995) but bulb formation <strong>of</strong> lily was<br />

completely inhibited by fluridone and restored by<br />

simultaneous addition <strong>of</strong> ABA (Kim et al., 1994).<br />

To establish the effect <strong>of</strong> osmoticum directly,<br />

experiments have been carried out with addition <strong>of</strong>


142 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

mannitol instead <strong>of</strong> sucrose. Mannitol did not<br />

promote corm production <strong>of</strong> Gladiolus (De Bruyn<br />

and Ferreira, 1992) or bulb production in onion<br />

(Kahane and Rancillac, 1996), but data for lily<br />

suggested that, although it had a toxic effect, in this<br />

plant mannitol did stimulate bulb formation (Gerrits<br />

and De Klerk, 1992).<br />

Anther culture. <strong>The</strong> use <strong>of</strong> high concentrations<br />

<strong>of</strong> sucrose is commonly reported in papers on anther<br />

culture where the addition <strong>of</strong> 5-20% sucrose to the<br />

culture medium is found to assist the development <strong>of</strong><br />

somatic embryos from pollen microspores. This<br />

appears to be due to an osmotic regulation <strong>of</strong><br />

morphogenesis (Sunderland and Dunwell, 1977), for<br />

once embryoid development has commenced, such<br />

high levels <strong>of</strong> sucrose are no longer required, or may<br />

be inhibitory. A high concentration <strong>of</strong> mannitol has<br />

been used for pretreatment before the culture <strong>of</strong><br />

barley anthers (Roberts-Oehlschlager and Dunwell<br />

1990) and pollen (Wei et al., 1986); tobacco anthers<br />

(Imamura and Harada 1980) and pollen (Imamura<br />

et al., 1982); and before wheat microspore culture<br />

(Hu et al., 1995). A high concentration <strong>of</strong> mannitol<br />

has also been used to induce osmotic stress in<br />

microspore derived embryos <strong>of</strong> Brassica napus<br />

(Huang et al., 1991) and before anther culture <strong>of</strong><br />

Brassica campestris (Hamaoka et al., 1991). Isolated<br />

microspores <strong>of</strong> Brassica napus cultured on a high<br />

concentration <strong>of</strong> mannitol and at a low concentration<br />

<strong>of</strong> sucrose (0.08–0.1%) yield no embryos whereas on<br />

high polyethyleneglycol 4000 the embryo yield is<br />

comparable to that <strong>of</strong> the sucrose control (Ikeda-Iwai,<br />

2003). <strong>The</strong>se results demonstrate that in microspore<br />

embryogenesis <strong>of</strong> Brassica napus the level <strong>of</strong><br />

metabolizable carbohydrate required for microspore<br />

embryo induction and formation may be very low and<br />

that an appropriate osmoticum (polyethylene glycol<br />

4000 or sucrose) is required.<br />

<strong>The</strong> temporary presence <strong>of</strong> high sucrose<br />

concentrations is said to prevent the proliferation <strong>of</strong><br />

callus from diploid cells <strong>of</strong> the anther that would<br />

otherwise swamp the growth <strong>of</strong> the pollen-derived<br />

embryoids. <strong>The</strong> concentration <strong>of</strong> macronutrient ions<br />

generally used in anther culture media is not<br />

especially low. In a sample <strong>of</strong> reports it was found to<br />

be 68.7 mM (George et al., 1987), and so the total<br />

osmotic potential, Ψs, (salts plus sucrose) <strong>of</strong> many<br />

anther culture media is in the range -0.55 to -1.15<br />

MPa.<br />

4.3.5. Relative humidity.<br />

<strong>The</strong> vapour pressure <strong>of</strong> water is reduced by<br />

dissolving substances in it. This means that the<br />

relative humidity <strong>of</strong> the air within closed culture<br />

vessels is dependent on the water potential <strong>of</strong> the<br />

medium according to the equation:<br />

1000RT<br />

⎛ p<br />

Ψ = lne ⎜<br />

W ⎜<br />

0 ⎝ p0<br />

(after Glasstone, 1947)<br />

⎞ ⎛ dp ⎞<br />

⎟ .<br />

⎟ ⎜ p − c ⎟<br />

⎠ ⎝ dc ⎠<br />

where: Ψ, R and T are as in the Van’t H<strong>of</strong>f equation,<br />

W0 is the molecular weight <strong>of</strong> water, c is the<br />

concentration <strong>of</strong> the solution in moles per litre, and p0<br />

and p are respectively the vapour pressures <strong>of</strong> water<br />

and the solution.<br />

If the change in vapour pressure dependent on the<br />

density <strong>of</strong> the solution, p – c (dp/dc), is treated as<br />

unity (legitimate perhaps for very dilute solutions, or<br />

when the equation is expressed in molality, rather<br />

than molarity), it is possible to estimate relative<br />

humidity (100 × p/p0) above tissue culture media <strong>of</strong><br />

known water potentials, from:<br />

1000RT ⎛<br />

Ψ = lne<br />

18.<br />

016 ⎜<br />

⎝<br />

p ⎞<br />

⎟ , (Lang, 1967)<br />

⎠<br />

p0<br />

<strong>The</strong> relative humidity above most plant tissue<br />

cultures in closed vessels is thus calculated to be in<br />

the range 99.25-99.75% (Table 4.7), the osmolality <strong>of</strong><br />

some typical media being<br />

• White (1963), liquid, 2% sucrose, 106 mOsm/kg<br />

• MS, agar, 3% sucrose, 230 mOsm/kg<br />

• MS, agar, 6% sucrose, 359 mOsm/kg<br />

• MS, agar, 12% sucrose (unusual), 659 mOsm/kg<br />

(in these cases, 50% hydrolysis <strong>of</strong> sucrose into<br />

monosaccharides is assumed to have taken place<br />

during autoclaving)<br />

Relative humidity can be reduced below the levels<br />

indicated above by, for example, covering vessels<br />

with gas-permeable closures while using nongelatinous<br />

support systems (see Section 6.3.1).


Table 4.7 <strong>The</strong> relative humidity within culture vessels which would be expected to result from the use <strong>of</strong><br />

plant culture media <strong>of</strong> various osmolalities or water potentials<br />

<strong>The</strong> relative acidity or alkalinity <strong>of</strong> a solution is<br />

assessed by its pH. This is a measure <strong>of</strong> the hydrogen<br />

ion concentration; the greater the concentration <strong>of</strong> H +<br />

ions (actually H3O + ions), the more acid the solution.<br />

As pH is defined as the negative logarithm <strong>of</strong><br />

hydrogen ion concentration, acid solutions have low<br />

pH values (0-7) and alkaline solutions, high values<br />

(7-14). Solutions <strong>of</strong> pH 4 (the concentration <strong>of</strong> H + is<br />

10 –4 mol.l –1 ) are therefore more acid than those <strong>of</strong> pH<br />

5 (where the concentration <strong>of</strong> H + is 10 –5 mol.l –1 );<br />

solutions <strong>of</strong> pH 9 are more alkaline than those <strong>of</strong> pH<br />

8. Pure water, without any dissolved gases such as<br />

CO2, has a neutral pH <strong>of</strong> 7. To judge the effect <strong>of</strong><br />

medium pH, it is essential to discriminate between<br />

the various sites where the pH might have an effect:<br />

(1) in the explant, (2) in the medium and (3) at the<br />

interface between explant and medium.<br />

<strong>The</strong> pH <strong>of</strong> a culture medium must be such that it<br />

does not disrupt the plant tissue. Within the<br />

acceptable limits the pH also:<br />

• governs whether salts will remain in a soluble<br />

form;<br />

• influences the uptake <strong>of</strong> medium ingredients and<br />

plant growth regulator additives;<br />

• has an effect on chemical reactions (especially<br />

those catalysed by enzymes); and<br />

• affects the gelling efficiency <strong>of</strong> agar.<br />

This means that the effective range <strong>of</strong> pH for<br />

media is restricted. As will be explained, medium pH<br />

is altered during culture, but as a rule <strong>of</strong> thumb, the<br />

initial pH is set at 5.5 – 6.0. In culture media,<br />

detrimental effects <strong>of</strong> an adverse pH are generally<br />

related to ion availability and nutrient uptake rather<br />

than cell damage.<br />

5.1. THE pH OF MEDIA<br />

5.1.1. Buffering<br />

<strong>The</strong> components <strong>of</strong> common tissue culture media<br />

have only little buffering capacity. Vacin and Went<br />

(1949) investigated the effect on pH <strong>of</strong> each<br />

Chapter 4 143<br />

Relative humidity in vessel<br />

Water potential (kPa) at temperature:<br />

(%)<br />

Osmolality<br />

(mOsm/kg) 15 ºC 20 ºC 25 ºC 30 ºC<br />

98 1121 -2686 -2733 -2780 -2826<br />

99 558 -1337 -1360 -1383 -1406<br />

99.5 278 -667 -678 -690 -701<br />

99.75 139 -333 -339 -344 -350<br />

5. pH OF TISSUE CULTURE MEDIA<br />

compound in their medium. <strong>The</strong> chemicals which<br />

seemed to be most instrumental in changing pH were<br />

FeSO4.7H2O and Ca(NO3)2.4H2O. Replacing the<br />

former with ferric tartrate at a weight which<br />

maintained the original molar concentration <strong>of</strong> iron,<br />

and substituting Ca3(PO4)2 and KNO3 for<br />

Ca(NO3)2.4H2O, they found that the solution was<br />

more effectively buffered. While amino acids also<br />

showed promise as buffering agents, KH2PO4 was<br />

ineffective unless it was at high concentration. In<br />

some early experiments attempts were made to<br />

stabilise pH by incorporating a mixture <strong>of</strong> KH2PO4<br />

and K2HPO4 into a medium (see Kordan, 1959, for<br />

example), but Street and Henshaw (1966) found that<br />

significant buffering was only achieved by soluble<br />

phosphates at levels inhibitory to plant growth. For<br />

this reason Sheat et al. (1959) proposed the buffering<br />

<strong>of</strong> plant root culture media with sparingly-soluble<br />

calcium phosphates, but unfortunately if these<br />

compounds are autoclaved with other medium<br />

constituents, they absorb micronutrients which then<br />

become unavailable.<br />

A buffer is a compound which can poise the pH<br />

level at a selected level: effective buffers should<br />

maintain the pH with little change as culture<br />

proceeds. As noted before, plant tissue culture media<br />

are normally poorly buffered. However, pH is<br />

stabilised to a certain extent when tissues are cultured<br />

in media containing both nitrate and ammonium ions.<br />

Agar and Gelrite gelling agents may have a slight<br />

buffering capacity (Scherer, 1988).<br />

Organic acids: Many organic acids can act as<br />

buffers in plant culture media. By stabilizing pH at<br />

ca. pH 5.5, they can facilitate the uptake <strong>of</strong> NH4 +<br />

when this is the only source <strong>of</strong> nitrogen, and by their<br />

own metabolism, assist the conversion <strong>of</strong> NH4 + into<br />

amino acids. <strong>The</strong>re can be improvement to growth<br />

from adding organic acids to media containing both<br />

NH4 + and NO3 – , but this is not always the case.<br />

Norstog and Smith (1963) noted that 0.75 mM malic


144 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

acid was an effective buffer and appeared to enhance<br />

the effect <strong>of</strong> the glutamine and alanine which they<br />

added to their medium. Vyskot and Bezdek (1984)<br />

found that the buffering capacity <strong>of</strong> MS medium was<br />

increased by adding either 1.25 mM sodium citrate or<br />

1.97 mM citric acid plus 6.07 mM dibasic sodium<br />

phosphate. Citric acid and some other organic acids<br />

have been noted to enhance the growth <strong>of</strong> Citrus<br />

callus when added to the medium (presumably that <strong>of</strong><br />

Murashige and Tucker, 1969) (Erner and Reuveni,<br />

1981). For the propagation <strong>of</strong> various cacti from<br />

axillary buds, Vyskot and Jara (1984) added sodium<br />

citrate to MS medium to increase its buffering<br />

capacity.<br />

Recognized biological buffers: Unlike organic<br />

acids, conventional buffers are not metabolised by the<br />

plant, but can poise pH levels very effectively.<br />

Compounds which have been used in plant culture<br />

media for critical purposes such as protoplast<br />

isolation and culture, and culture <strong>of</strong> cells at very low<br />

inoculation densities, include:<br />

• TRIS, Tris(hydroxymethyl)aminomethane;<br />

• Tricine, N-tris(hydroxymethyl)methylglycine;<br />

• MES, 2-(N-morpholino)ethanesulphonic acid;<br />

• HEPES, 4-(2-hydroxyethyl)-1-piperazine(2ethanesulphonic<br />

acid); and<br />

• CAPS, 3-cyclohexylamino-1-propanesulphonic<br />

acid.<br />

Such compounds may have biological effects<br />

which are unrelated to their buffering capacity.<br />

Depending on the plant species, they have been<br />

known to kill protoplasts or greatly increase the rate<br />

<strong>of</strong> cell division and/or plant regeneration (Conrad<br />

et al., 1981). <strong>The</strong> ‘biological’ buffers MES and<br />

HEPES have been developed for biological research<br />

(Good et al., 1966).<br />

MES: MES is one <strong>of</strong> the few highly effective and<br />

commercially available buffers with significant<br />

buffering capacity in the pH range 5-6 to which plant<br />

culture media are usually adjusted and has only a low<br />

capacity to complex with micronutrients. It is not<br />

toxic to most plants, although there are some which<br />

are sensitive. Ramage and Williams (2002) report<br />

that shoot regeneration from tobacco leaf discs was<br />

not affected by MES when increasing the<br />

concentration up to 100 mM. De Klerk et al (2007),<br />

though, observed a decrease <strong>of</strong> rooting from apple<br />

stem slices with increasing MES concentration (see<br />

below; Fig. 4.4). This effect <strong>of</strong> MES was not<br />

understood. It only occurred during the first days <strong>of</strong><br />

the rooting process and was not observed during the<br />

outgrowth phase after the meristems had been<br />

formed.<br />

During the culture <strong>of</strong> thin cell layers <strong>of</strong> Nicotiana,<br />

Tiburcio et al. (1989) found that the pH <strong>of</strong> LS<br />

medium could be kept close to 5.8 for 28 days by<br />

adding 50 mM MES, whereas without the buffer pH<br />

gradually decreased to 5.25. Regulating pH with<br />

MES alters the type <strong>of</strong> morphogenesis which occur in<br />

this (and other) tissues (see below).<br />

Parfitt et al. (1988) found 10 mM MES to be an<br />

effective buffer in four different media used for<br />

tobacco, carrot and tomato callus cultures, and peach<br />

and carnation shoot cultures, although stabilizing pH<br />

did not result in superior growth. <strong>The</strong> tobacco, peach<br />

and carnation cultures were damaged by 50 mM<br />

MES. Tris was toxic at all concentrations tested,<br />

although Klein and Manos (1960) had found that the<br />

addition <strong>of</strong> only 0.5 mM Tris effectively increased<br />

the fresh weight <strong>of</strong> callus which could be grown on<br />

White (1954) medium when iron was chelated with<br />

EDTA.<br />

MES has also been used successfully to buffer<br />

many cultures initiated from single protoplasts<br />

(Müller et al., 1983) and its inclusion in the culture<br />

medium can be essential for the survival <strong>of</strong> individual<br />

cells and their division to form callus colonies (e.g.<br />

those <strong>of</strong> Datura innoxia - Koop et al., 1983). MES<br />

was found to be somewhat toxic to single protoplasts<br />

<strong>of</strong> Brassica napus, but ‘Polybuffer 74’ (PB-74, a<br />

mixture <strong>of</strong> polyaminosulphonates), allowed excellent<br />

microcolony growth in the pH range 5.5-7.0<br />

(Spangenberg et al., 1986). Used at 1/100 <strong>of</strong> the<br />

commercially available solution, it has a buffering<br />

capacity <strong>of</strong> a 1.3 mM buffer at its pK value (Koop<br />

et al., 1983).<br />

Banana homogenate is widely used in orchid<br />

micropropagation media. Ernst (1974) noted that it<br />

appeared to buffer the medium in which slipper<br />

orchid seedlings were being grown.<br />

5.1.2. <strong>The</strong> uptake <strong>of</strong> ions and molecules<br />

<strong>The</strong> pH <strong>of</strong> the medium has an effect on the<br />

availability <strong>of</strong> many minerals (Scholten and Pierik,<br />

1998). In general, the uptake <strong>of</strong> negatively charged<br />

ions (anions) is favoured at acid pH, while that <strong>of</strong><br />

cations (positively charged) is best when the pH is<br />

increased. As mentioned before, the relative uptake <strong>of</strong><br />

nutrient cations and anions will alter the pH <strong>of</strong> the<br />

medium. <strong>The</strong> release <strong>of</strong> hydroxyl ions from the plant<br />

in exchange for nitrate ions results in media<br />

becoming more alkaline; when ammonium ions are


taken up in exchange for protons, media become<br />

more acid (Table 4.8).<br />

Nitrate and ammonium ions: <strong>The</strong> uptake <strong>of</strong><br />

ammonium and nitrate ions is markedly affected by<br />

pH. Excised plant roots can be grown with NH4 + as<br />

the sole source <strong>of</strong> nitrogen providing the pH is<br />

maintained within the range 6.8 to 7.2 and iron is<br />

available in a chelated form. At pH levels below 6.4<br />

the roots grow slowly on ammonium alone and have<br />

an abnormal appearance (Sheat et al.,1959). This<br />

accords with experiments on intact plants where<br />

ammonium as the only source <strong>of</strong> nitrogen is found to<br />

be poorly taken up at low pH. It was most effectively<br />

utilised in Asparagus in a medium buffered at pH 5.5.<br />

At low pH (ca. 4 or less) the ability <strong>of</strong> the roots <strong>of</strong><br />

most plants to take up ions <strong>of</strong> any kind may be<br />

impaired, there may be a loss <strong>of</strong> soluble cell<br />

constituents and growth <strong>of</strong> both the main axis and<br />

that <strong>of</strong> laterals is depressed (Asher, 1978).<br />

Nitrate on the other hand, is not readily absorbed<br />

by plant cells at neutral pH or above (Martin and<br />

Rose, 1976). Growth <strong>of</strong> tumour callus <strong>of</strong> Rumex<br />

acetosa on a nitrate-containing medium was greater<br />

at pH 3.5 than pH 5.0 (Nickell and Burkholder,<br />

1950), while Chevre et al. (1983) reported that<br />

axillary bud multiplication in shoot cultures <strong>of</strong><br />

Castanea was most satisfactory when the pH <strong>of</strong> MS<br />

was reduced to 4 and the Ca 2+ and Mg 2+<br />

concentrations were doubled.<br />

pH Stabilization: One <strong>of</strong> the chief advantages <strong>of</strong><br />

having both NO3 – and NH4 + ions in the medium is<br />

that uptake <strong>of</strong> one provides a better pH environment<br />

for the uptake <strong>of</strong> the other. <strong>The</strong> pH <strong>of</strong> the medium is<br />

thereby stabilized. Uptake <strong>of</strong> nitrate ions by plant<br />

cells leads to a drift towards an alkaline pH, while<br />

NH4 + uptake results in a more rapid shift towards<br />

acidity (Street, 1969; Behrend and Mateles, 1975;<br />

Hyndman et al., 1982). For each equivalent <strong>of</strong><br />

ammonium incorporated into organic matter, about<br />

0.8-1 H + (proton) equivalents are released into the<br />

external medium; for each equivalent <strong>of</strong> nitrate<br />

assimilated, 1-1.2 proton equivalents are removed<br />

from the medium (Fuggi et al., 1981). Raven (1986)<br />

calculated that there should be no change <strong>of</strong> pH<br />

Table 4.8 <strong>The</strong> uptake <strong>of</strong> ions and its consequences in plant culture media<br />

Uptake anion (e.g. NH4 + )<br />

Rate Best in alkaline or weakly acid solutions<br />

Consequence Protons (H + ) extruded by plant<br />

Medium becomes more ACID<br />

Chapter 4 145<br />

Uptake cation (e.g. NO3 ⎯ )<br />

Best in relatively acid solutions<br />

Hydroxyl (OH ⎯ ) ions extruded by plant<br />

Medium becomes more ALKALINE<br />

resulting from NO3 – or NH4 + uptake, when the ratio<br />

<strong>of</strong> the two is 2 to 1.<br />

<strong>The</strong> pH shifts caused by uptake <strong>of</strong> nitrate and<br />

ammonium during culture (see above) can lead to a<br />

situation <strong>of</strong> nitrogen deficiency if either is used as the<br />

sole nitrogen source without the addition <strong>of</strong> a buffer<br />

(see later) (Hyndman et al., 1982). <strong>The</strong> uptake <strong>of</strong><br />

NH4 + , when this is the sole source <strong>of</strong> nitrogen is only<br />

efficient when a buffer or an organic acid is also<br />

present in the medium. In media containing both<br />

NO3 – and NH4 + with an initial pH <strong>of</strong> 5-6, preferential<br />

uptake <strong>of</strong> NH4 + causes the pH to drop during the early<br />

growth <strong>of</strong> the culture. This results in increased NO3 –<br />

utilisation (Martin and Rose, 1976) and a gradual pH<br />

rise. <strong>The</strong> final pH <strong>of</strong> the medium depends on the<br />

relative proportions <strong>of</strong> NO3 – and NH4 + which are<br />

provided (Gamborg et al., 1968). After 7 days <strong>of</strong> root<br />

culture, White (1943a) medium (containing only<br />

nitrate), adjusted to pH 4.8-4.9, had a pH <strong>of</strong> 5.8-6.0<br />

(Street et al., 1951, 1952), but Sheat et al. (1959)<br />

could stabilize the medium at pH 5.8 by having one<br />

fiftieth <strong>of</strong> the total amount <strong>of</strong> nitrogen as ammonium<br />

ion, the rest as nitrate. Changes in the pH <strong>of</strong> a<br />

medium do however vary from one kind <strong>of</strong> plant to<br />

another. Ramage and Williams (2002) observed a<br />

decrease in pH when tobacco leaf discs were cultured<br />

with only NH4 + nitrogen whereas no such decrease<br />

was observed on medium with both NH4 + and NO3 – .<br />

No organogenesis occurred when the medium with<br />

only NH4 + was unbuffered but the inclusion <strong>of</strong> MES<br />

resulted in the formation <strong>of</strong> meristems (but no<br />

shoots).<br />

<strong>Media</strong> differing in total nitrogen levels (but all<br />

having the same ratio <strong>of</strong> nitrate to ammonium as MS<br />

medium), had a final pH <strong>of</strong> ca. 4.5 after being used<br />

for Stage <strong>II</strong>I root initiation on rose shoots, whereas<br />

those containing ammonium alone had a final pH <strong>of</strong><br />

ca. 4.1 (Hyndman et al., 1982). A similar observation<br />

with MS medium itself was made by Delfel and<br />

Smith (1980). No matter what the starting value in<br />

the range 4.5-8.0, the final pH after culture <strong>of</strong><br />

Cephalotaxus callus was always 4.2. <strong>The</strong> medium <strong>of</strong><br />

De Jong et al. (1974) always had a pH <strong>of</strong> 4.8-5.0 after<br />

Begonia buds had been cultured, whatever the


146 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

starting pH in the range 4.0-6.5. This is not to say that<br />

the initial pH was unimportant, because there was an<br />

optimum for growth and development (see later)<br />

(Berghoef and Bruinsma, 1979).<br />

Other Compounds: <strong>The</strong> availability and uptake<br />

<strong>of</strong> other inorganic ions and organic molecules is also<br />

affected by pH. As explained above, the uptake <strong>of</strong><br />

phosphate is most efficient from acid solutions.<br />

Petunia cells took up phosphate most rapidly at pH 4<br />

and its uptake declined as the pH was raised (Chin<br />

and Miller, 1982). Vacin and Went (1949) noted the<br />

formation <strong>of</strong> iron phosphate complexes in their<br />

medium by pH changes. Insoluble iron phosphates<br />

can also be formed in MS medium at pH 6.2 or above<br />

unless the proportion <strong>of</strong> EDTA to iron is increased<br />

(Dalton et al., 1983).<br />

<strong>The</strong> uptake <strong>of</strong> Cl – ions into barley roots is<br />

favoured by low pH, but Jacobson et al. (1971)<br />

noticed that it was only notably less at high pH in<br />

solutions strongly stirred by high aeration. <strong>The</strong>y<br />

therefore suggested that H + ions secreted from plant<br />

roots as a result <strong>of</strong> the uptake <strong>of</strong> anions, can maintain<br />

a zone <strong>of</strong> reduced pH in the Nernst layer, a stationary<br />

film <strong>of</strong> water immediately adjacent to cultured plant<br />

tissue (Nernst, 1904). In plant tissue culture, uptake<br />

<strong>of</strong> ions and molecules may therefore more liable to be<br />

affected by adverse pH in agitated liquid media, than<br />

in media solidified by agar.<br />

Lysine uptake into tobacco cells was found by<br />

Harrington and Henke (1981) to be stimulated by low<br />

pH. <strong>The</strong> effect <strong>of</strong> pH on uptake is especially relevant<br />

for auxins (e.g., Edwards and Goldsmith 1980).<br />

Depending on the pH and their pKa, auxins are either<br />

present as an undissociated molecule or as an anion.<br />

For influx, the undissociated auxin molecule may<br />

pass through the membrane by diffusion whereas the<br />

anion is taken up by a carrier (Delbarre et al., 1996;<br />

Morris 2000). Dissociation dependens upon the pH.<br />

In the apoplast, the pH is low, ca. 5. When taken up,<br />

the auxin enters the cytoplasm with pH 7. At this pH,<br />

most auxin is present as anion and cannot diffuse out.<br />

Efflux <strong>of</strong> the anion is brought about by an efflux<br />

carrier system. Thus, the net uptake into cells <strong>of</strong> plant<br />

growth regulators which are weak lipophilic acids<br />

(such as IAA, NAA, 2,4-D and abscisic acid) will be<br />

greater, the more acid the medium and the greater the<br />

difference between its pH and that <strong>of</strong> the cell<br />

cytoplasm (Rubery, 1980). Shvetsov and Gamburg<br />

(1981) did in fact find that the rate <strong>of</strong> 2,4-D uptake<br />

into cultured corn cells increased as the pH <strong>of</strong> the<br />

medium fell from 5.5 to 4. Increased uptake <strong>of</strong> auxin<br />

at low pH was also found in apple microcuttings,<br />

both for IBA (Harbage et al., 1998) and IAA (De<br />

Klerk et al., 2007). As previously mentioned, auxins<br />

can themselves modify intra- and extra-cellular pH.<br />

Adding 2,4-D to a medium increased the uptake <strong>of</strong><br />

nicotine into culture <strong>of</strong> Acer pseudoplatanus cells<br />

(Kurkdjian et al., 1982).<br />

5.1.3. Choosing the pH <strong>of</strong> culture media: Starting pH<br />

Many plant cells and tissues in vitro, will tolerate<br />

pH in the range <strong>of</strong> about 4.0-7.2; those inoculated<br />

Fig. 4.3 Development <strong>of</strong> pH during tissue culture. <strong>The</strong> pH was set before autoclaving as is usually done, and<br />

measured directly after autoclaving and after 5 days <strong>of</strong> culture with 1-mm stem-slices cut from apple microshoots.<br />

<strong>The</strong> medium was in Petri dishes with BBL agar and modified MS medium. Per Petri dish, 30 ml <strong>of</strong> medium was<br />

added and 30 slices were cultured. (Data from de Klerk et al., 2007).


into media adjusted to pH 2.5-3.0 or 8.0 will probably<br />

die (Butenko et al., 1984). Best results are usually<br />

obtained in slightly acid conditions. In a random<br />

sample <strong>of</strong> papers on micropropagation, the average<br />

initial pH adopted for several different media was<br />

found to be 5.6 (mode 5.7) but adjustments to as low<br />

as 3.5 and as high as 7.1 had been made.<br />

<strong>The</strong> pH for most plant cultures is thus lower than<br />

that which is optimum for hydroponic cultures, where<br />

intact plants with their roots in aerated solution<br />

usually grow most rapidly when the pH <strong>of</strong> the<br />

solution is in the range 6.0-7.3 (De Capite, 1948;<br />

Sholto Douglas, 1976; Cooper, 1979).<br />

Suspension cultured cells <strong>of</strong> Ipomoea, grew<br />

satisfactorily in Rose and Martin (1975) P2 medium<br />

adjusted initially to pH 5.6 or pH 6.3, but the yield <strong>of</strong><br />

dry cells was less at two extremes (pH 4.8 and pH<br />

7.1). Martin and Rose (1976) supposed that a low<br />

yield <strong>of</strong> cells from a culture started at pH 7.1 was due<br />

to the inability <strong>of</strong> the culture to utilise NO3 – , but the<br />

cause <strong>of</strong> a reduction in growth at pH 4.8 was less<br />

obvious. It might have resulted from the plant having<br />

to expend energy to maintain an appropriate<br />

physiological pH internally.<br />

Kartha (1981) found that pH 5.6-5.8 supported the<br />

growth <strong>of</strong> most meristem tips in culture and that<br />

cassava meristems did not grow for a prolonged<br />

period on a medium adjusted to pH 4.8. <strong>The</strong> optimum<br />

pH (before autoclaving) for the growth <strong>of</strong> carnation<br />

shoot tips on Linsmaier and Skoog (1965) medium,<br />

was 5.5-6.5. When the medium was supplemented<br />

with 4 mg/l adenine sulphate and 2 g/l casein<br />

hydrolysate, the optimum pH was 5, and on media<br />

adjusted to 6.0 and 6.5, there was significantly less<br />

growth (Davis et al., 1977). Shoot proliferation in<br />

Camellia sasanqua shoot cultures was best when the<br />

pH <strong>of</strong> a medium with MS salts was adjusted to 5-5.5.<br />

Only in these flasks was the capacity <strong>of</strong> juvenile<br />

explants to produce more shoots than adult ones<br />

really pronounced (Torres and Carlisi, 1986).<br />

Norstog and Smith (1963) suggested that the pH<br />

<strong>of</strong> the medium used for the culture <strong>of</strong> isolated zygotic<br />

embryos, should not be greater than 5.2.<br />

5.1.4. pH adjustment<br />

Because there is then no need to take special<br />

aseptic precautions, and it is impractical to adjust pH<br />

once medium has been dispensed into small lots, the<br />

pH <strong>of</strong> a medium is usually adjusted with acid or<br />

alkali before autoclaving. According to Krieg and<br />

Gerhardt (1981), agar is partially hydrolysed if<br />

autoclaved at pH 6 or less and will not solidify so<br />

effectively when cooled. <strong>The</strong>y recommend that agar<br />

Chapter 4 147<br />

media for bacteriological purposes should be<br />

sterilised at a pH greater than 6 and, if necessary,<br />

should be adjusted to acid conditions with sterile acid<br />

after autoclaving, when they have cooled to 45-50°C.<br />

<strong>The</strong> degree <strong>of</strong> hydrolysis in plant culture media<br />

autoclaved at pH 5.6-5.7, is presumably small.<br />

<strong>The</strong> effect <strong>of</strong> autoclaving: Autoclaving changes<br />

the pH <strong>of</strong> media (Fig. 4.3). In media without sugars,<br />

the change is usually small unless the phosphate<br />

concentration is low, when more significant<br />

fluctuations occur. <strong>Media</strong> autoclaved with sucrose<br />

generally have a slightly lower pH than those<br />

autoclaved without it, but if maltose, glucose, or<br />

fructose have been added instead <strong>of</strong> sucrose, the post-<br />

autoclave pH is significantly reduced (Owen et al.,<br />

1991).<br />

<strong>The</strong> pH <strong>of</strong> liquid media containing MS salts [e.g.<br />

Linsmaier and Skoog (1965) or Skirvin and Chu<br />

(1979) media] containing 3-3.4% sucrose, has been<br />

found to fall during autoclaving from an adjusted<br />

level <strong>of</strong> 5.7, to pH 5.17 (Singha, 1982), to pH 5.5<br />

(Owen et al., 1991), or to pH 4.6 (Skirvin et al.,<br />

1986). <strong>The</strong> drop in pH may vary according to the pH<br />

to which the medium was initially adjusted. In the<br />

experiments <strong>of</strong> Skirvin et al. (loc. cit.), the pH <strong>of</strong> a<br />

medium adjusted to 5.0, fell to 4.2; one adjusted to<br />

6.4, fell to 5.1; that set at pH 8.5, fell to 8.1.<br />

Most agars cause the pH <strong>of</strong> media to increase<br />

immediately they are dissolved. Knudson (1946)<br />

noticed that the pH <strong>of</strong> his medium shifted from 4.6-<br />

4.7 to 5.4-5.5 once agar had been added and<br />

dissolved; and Singha (1982) discovered that the<br />

unadjusted pH <strong>of</strong> MS medium rose from pH 4 to pH<br />

5.2, depending on the amount <strong>of</strong> agar added.<br />

However if a medium containing agar was adjusted to<br />

pH 5.7, and then autoclaved, the medium became<br />

more acid than if agar had not been added, the fall in<br />

pH being generally in proportion to the amount <strong>of</strong><br />

agar present.<br />

<strong>The</strong> effect <strong>of</strong> storage. <strong>The</strong> pH <strong>of</strong> autoclaved plant<br />

media tends to fall if they are stored. Vacin and<br />

Went (1949) noted that autoclaving just accelerated a<br />

drop in the pH <strong>of</strong> Knudson (1946) C medium, as<br />

solutions left to stand showed similar changes.<br />

Sterilisation by filtration (see below) was not a<br />

satisfactory alternative, as it too effected pH changes.<br />

<strong>The</strong> compounds particularly responsible were thought<br />

to be unchelated ferrous sulphate (when the pH had<br />

initially been set between 3 and 6), and calcium<br />

nitrate (when the original pH was 6 to 9). Complex<br />

iron phosphates, unavailable to plants, were produced<br />

from the ferrous sulphate, but if iron was added as


148 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

ferric citrate or ferric tartrate (chelates), no significant<br />

pH changes resulted from autoclaving, and plants<br />

showed no iron-deficiency symptoms.<br />

Skirvin et al. (1986) found that both with and<br />

without agar, autoclaved MS medium tended to<br />

become more acid after 6 weeks storage, for example:<br />

MS medium<br />

Time Liquid With 0.6% Difco<br />

Bacto Agar<br />

Starting pH 5.7 5.7<br />

After autoclaving 4.6 4.6<br />

6 weeks later 4.1 4.4<br />

To minimise a change in the pH <strong>of</strong> stored media,<br />

it is suggested that they are kept in the dark: Owen<br />

et al., (1991) reported that the pH <strong>of</strong> MS containing<br />

0.1M sucrose or 0.8% Phytagar, remained relatively<br />

stable after autoclaving if kept in the dark at 4°C, but<br />

fell by up to 0.8 units if stored in the light at 25°C. De<br />

Klerk et al. (2007) using BBL agar, also observed a<br />

shift <strong>of</strong> pH, but this was negligible when 10 mM<br />

MES was added (Fig. 4.3).<br />

Hydrolysis: Some organic components <strong>of</strong> culture<br />

media are liable to be hydrolysed by autoclaving in<br />

acid media. <strong>The</strong> degree <strong>of</strong> hydrolysis <strong>of</strong> different<br />

brands <strong>of</strong> agar may be one factor influencing the<br />

incidence <strong>of</strong> hyperhydricity in plant cultures. Agar<br />

media may not solidify satisfactorily when the initial<br />

pH has been adjusted to 4-4.5. <strong>The</strong> reduction in pH<br />

which occurs in most media during autoclaving may<br />

also cause unsatisfactory gelling <strong>of</strong> agar which has<br />

been added in low concentrations. Part <strong>of</strong> the sucrose<br />

added to plant culture media adjusted to pH 5.5 is<br />

also hydrolysed during autoclaving: the proportion<br />

degraded increases if the pH <strong>of</strong> the medium is much<br />

less than this. Hydrolysis <strong>of</strong> sucrose is, however, not<br />

necessarily detrimental.<br />

Activated charcoal: <strong>The</strong> presence <strong>of</strong> activated<br />

charcoal can alter the pH <strong>of</strong> a medium. As in the<br />

production <strong>of</strong> activated charcoal it is, at one stage<br />

washed with HCl, the pH <strong>of</strong> a medium can be<br />

lowered by acid residues (Owen et al., 1995; Wann<br />

et al., 1997).<br />

4.1.5. pH changes during culture<br />

Due to the differential uptake <strong>of</strong> anions and<br />

cations into plant tissues, the pH <strong>of</strong> culture media<br />

does not usually stay constant, but changes as ions<br />

and compounds are absorbed by the plant. It is usual<br />

for media containing nitrate and ammonium ions to<br />

decline slowly in pH during a passage, after being<br />

adjusted initially to pH 5.4-5.7, Sometimes after a<br />

preliminary decrease, the pH may begin to rise and<br />

return to a value close to, or even well above that at<br />

which the culture was initiated. <strong>The</strong> pH <strong>of</strong> White<br />

(1954) medium, which contains only NO3 – nitrogen,<br />

drifted from an initial 5.0 to 5.5 towards neutrality as<br />

callus was cultured on it (Klein and Manos, 1960).<br />

With some cultures, the initial pH <strong>of</strong> the medium<br />

may have little effect. Cell suspension cultures <strong>of</strong><br />

Dioscorea deltoides in the medium <strong>of</strong> Kaul and Staba<br />

(1968) [containing MS salts], adjusted to a pH either<br />

3.5, 4.3, 5.8 or 6.3, all had a pH <strong>of</strong> 4.6-4.7 within 10<br />

h <strong>of</strong> inoculation. <strong>The</strong>re was a further fall to pH 4.0-<br />

4.2 in the next 2-3 days, but during the following 15-<br />

17 days the pH was 4.7-5.0, finally rising to pH 6.0-<br />

6.3 on about day 19 (Butenko et al., 1984). Similarly,<br />

Skirvin et al. (1986) found that MS medium adjusted<br />

to pH 3.33, 5.11, 6.63, or 7.98 before autoclaving,<br />

and then used for the culture <strong>of</strong> Cucumis melo callus,<br />

had a pH after 48 h in the range 4.6-5.0. Visseur<br />

(1987) also reported that although the pH <strong>of</strong> his<br />

medium (similar macronutrients to MS but more Ca 2+<br />

and PO4 3– ) decreased if it was solidified with agar,<br />

but on a 2-phase medium, the final pH was 6.9 ± 0.4<br />

irrespective <strong>of</strong> whether it was initially 4.8, 5.5 or 6.2.<br />

Despite the above remarks, it should be noted that<br />

the nature <strong>of</strong> the pH drift which occurs in any one<br />

medium, differs widely, according to the species <strong>of</strong><br />

plant grown upon it. <strong>The</strong> pH <strong>of</strong> the medium<br />

supporting shoot cultures <strong>of</strong> Disanthus cercidifolius<br />

changed from 5.5 to 6.5 over a 6 week period,<br />

necessitating frequent subculturing to prevent the<br />

onset <strong>of</strong> senescence, whereas in the medium in which<br />

shoots <strong>of</strong> the calcifuge Lapageria rosea were grown,<br />

the pH, initially set to 3.5-5.0, only changed to 3.8-<br />

4.1 (Howard and Marks, 1987). Note also that the<br />

reversion <strong>of</strong> media to a homeostatic pH, may be due<br />

to the presence <strong>of</strong> both NO3 – and NH4 + ions (Dougall,<br />

1980). Adjustment <strong>of</strong> media containing only one <strong>of</strong><br />

these nitrogen sources to a range <strong>of</strong> pH levels, would<br />

be expected to result in a more variable set <strong>of</strong> final<br />

values.<br />

As the pH <strong>of</strong> media deviate from the original<br />

titration level, simple unmonitored cultures may not<br />

provide the most favourable pH for different phases<br />

<strong>of</strong> growth and differentiation. In Rosa ‘Paul’s Scarlet’<br />

suspensions, the optimum pH for the cell division<br />

phase was 5.2-5.4: pH 5.5-6.0 was best for the cell<br />

expansion phase (Nesius and Fletcher, 1973). <strong>The</strong><br />

maximum growth rate <strong>of</strong> Daucus carota habituated<br />

callus on White (1954) medium with iron as Fe-<br />

EDTA, occurred at pH 6.0 (Klein and Manos, 1960).


It has been suggested that acidification <strong>of</strong> media is<br />

partly due to the accumulation <strong>of</strong> carbon dioxide in<br />

tightly closed culture flasks (Leva et al., 1984), but<br />

the decrease in pH associated with incubating anther<br />

cultures with 5% CO2 was found by Johansson and<br />

Eriksson (1984) to be only ca. 0.1 units. Removal <strong>of</strong><br />

CO2 from an aerated cell suspension culture <strong>of</strong><br />

Poinsettia resulted in an increase <strong>of</strong> about 0.2 pHunits<br />

(Preil, 1991).<br />

Auxin plant growth regulators promote cell<br />

growth by inducing the efflux <strong>of</strong> H + ions through the<br />

cell wall. Hydrogen ion efflux from the cell is<br />

accompanied by potassium ion influx. When cultures<br />

are incubated in a medium containing an auxin, the<br />

medium will therefore become more acid while the<br />

pH <strong>of</strong> the cell sap will rise. <strong>The</strong> extent <strong>of</strong> these<br />

changes was found by Kurkdjian et al. (1982) to be<br />

proportional to auxin (2,4-D) concentration.<br />

5.2. pH CONTROL WITHIN THE PLANT<br />

<strong>The</strong> various compartments <strong>of</strong> cells have a<br />

different pH and this pH is maintained (Felle, 2001).<br />

In the symplasm, the pH <strong>of</strong> the cytoplasm is ca. 7 and<br />

<strong>of</strong> the vacuole ca. 5. <strong>The</strong> apoplasm has a pH <strong>of</strong> ca. 5.<br />

<strong>Plant</strong> cells typically generate an excess <strong>of</strong> acidic<br />

compounds during metabolism which have to be<br />

neutralised (Felle, 1998). One <strong>of</strong> the most important<br />

ways by which this is accomplished is for H + , or K +<br />

to be pumped out <strong>of</strong> the cell, in exchange for anions<br />

(e.g. OH – ), thereby decreasing the extracellular pH.<br />

<strong>Plant</strong> cells also compensate for an excess <strong>of</strong> H + by the<br />

degradation <strong>of</strong> organic acids. Synthesis <strong>of</strong> organic<br />

acids, such as malate, from neutral precursors is used<br />

to increase H + concentration when the cytoplasmic<br />

pH rises, for instance if plants are grown in alkaline<br />

soils (Raven and Smith, 1976; Findenegg et al.,<br />

1986). In intact plants, there is usually a downwards<br />

gradient from the low pH external to the cell, to<br />

higher pH levels in more mature parts, and this<br />

enables the upwards transport <strong>of</strong> non-electrolytic<br />

compounds such as sugars and amino acids (Böttger<br />

and Luthen, 1986).<br />

Altering the pH <strong>of</strong> the external solution<br />

surrounding roots or cells can alter the pH <strong>of</strong> the cell<br />

(Smith and Raven, 1979). Because <strong>of</strong> necessary<br />

controls, the pH <strong>of</strong> the cytoplasm may be only<br />

slightly altered, that <strong>of</strong> vacuoles may show a more<br />

marked change. Changing the pH <strong>of</strong> the medium in<br />

which photo-autotrophic Chenopodium rubrum<br />

suspensions were cultured from 4.5 to 6.3, caused the<br />

pH <strong>of</strong> the cytoplasm to rise from 7.4 to 7.6 and that <strong>of</strong><br />

the cell vacuoles to increase from 5.3 to 6.6. <strong>The</strong><br />

increase in cytoplasmic pH caused there to be a<br />

marked diversion <strong>of</strong> carbon metabolism, away from<br />

sugar and starch, into the production <strong>of</strong> lipids, amino<br />

acids and proteins (Hüsemann et al., 1990). <strong>The</strong><br />

maintenance <strong>of</strong> the pH is also illustrated in an<br />

experiment with detached leaves <strong>of</strong> Vicia faba (Felle<br />

and Hanstein, 2002). When the leaves were placed in<br />

a 10 mM MES-TRIS buffer and transferred to buffer<br />

with another pH, changes in the pH <strong>of</strong> the apoplasm<br />

were small: with an initial buffer pH = 4.1 and<br />

transfer to buffer pH = 6.8, the apoplastic pH <strong>of</strong><br />

substomatal cavities increased from 4.71 to 5.13 and<br />

in the reverse transfer decreased from 5.13 to 4.70.<br />

This indicates that the pH <strong>of</strong> the apoplasm is not<br />

strongly influenced by the medium but stays close to<br />

the ‘natural’ pH <strong>of</strong> ca. 5.0. No exact details are given<br />

but in this experiment the distance between the site <strong>of</strong><br />

the pH measurements and the MES-TRIS solution in<br />

which the leaves had been placed was probably large.<br />

<strong>The</strong> symplasm has a much larger capacity to buffer<br />

(Felle, 2001) so that its pH will be even less<br />

influenced by the medium pH. Thus, within the<br />

explant the pH <strong>of</strong> both apoplasm and symplasm will<br />

be affected only little by the medium pH. <strong>The</strong><br />

situation may be different at the interface between<br />

explant and medium. <strong>The</strong> influence <strong>of</strong> medium pH<br />

will extend towards inner tissues <strong>of</strong> the explant as the<br />

buffering capacity <strong>of</strong> the medium is increased (and<br />

thus overcomes buffering by the tissue). Inside the<br />

explant, the pH will also greatly influence movement<br />

Table 4.9. <strong>The</strong> influence <strong>of</strong> the initial pH <strong>of</strong> Linsmaier Skoog (1965) medium on morphogenesis in thin cell layers <strong>of</strong> Nicotiana<br />

(data <strong>of</strong> Mutaftschiev et al., 1987)<br />

Initial pH<br />

<strong>of</strong><br />

the medium<br />

Mean number <strong>of</strong><br />

organs per<br />

explant<br />

Chapter 4 149<br />

Mean percentage <strong>of</strong> explants forming:<br />

Callus Roots Vegetative buds Flowers<br />

3.8 4 ± 2 60 ± 10 40 ± 10 0 0<br />

5.0 2 ± 1 90 ± 10 10 ± 10 0 0<br />

6.1 20 ± 10 0 0 100 0<br />

6.8 6 ± 2 0 0 20 ± 10 80 ± 10


150 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

through membranes, i.e. uptake in cells, as this <strong>of</strong>ten<br />

depends on the dissociation <strong>of</strong> compounds which is<br />

pH dependent.<br />

Changing the pH <strong>of</strong> the medium can thus have a<br />

regulatory role on plant cultures which is similar to<br />

that <strong>of</strong> plant growth regulating chemicals, one <strong>of</strong> the<br />

actions <strong>of</strong> which is to modify intracellular pH and the<br />

quantity <strong>of</strong> free calcium ions. Auxins can modify<br />

cytoplasmic pH by triggering the release <strong>of</strong> H + from<br />

cells. In plant tissue culture these ions can acidify the<br />

medium (Kurkdjian et al., 1982). Proton release is<br />

thought to be the first step in acid-triggered and<br />

turgor-triggered growth (Schubert and Matzke, 1985).<br />

It should be noted that pH changes themselves may<br />

act as a signal (Felle, 2001).<br />

5.3. THE EFFECT OF pH ON CULTURES<br />

5.3.1. Initiating cultures at low pH<br />

<strong>Plant</strong>s <strong>of</strong> the family Ericaceae which only grow<br />

well on acid soils (e.g. rhododendrons and<br />

blueberries), have been said to grow best on media<br />

such as Anderson (1975), Anderson (1978; 1980) and<br />

Lloyd and McCown (1981) WPM, when the pH is<br />

first set to ca. 4.5 (Anderson, 1975; Skirvin, 1981),<br />

but for highly calcifuge species such as Magnolia<br />

soulangiana, a starting pH <strong>of</strong> 3.5 can result in the<br />

highest rate <strong>of</strong> shoot proliferation in shoot cultures<br />

(Howard and Marks, 1987). Chevre et al. (1983) state<br />

that chestnut shoot cultures grew and proliferated best<br />

at pH 4 provided MS medium was modified by<br />

doubling the usual levels <strong>of</strong> calcium and magnesium.<br />

De Jong et al. (1974), using a specially developed<br />

medium, found that a low pH value favoured the<br />

growth <strong>of</strong> floral organs. A similar result was seen by<br />

Berghoef and Bruinsma (1979c) with Begonia buds.<br />

Growth was greatest when the pH <strong>of</strong> the medium was<br />

initially adjusted to acid, 4.5-5.0 being optimal. At<br />

pH 4.0 the buds became glassy.<br />

Before the discovery <strong>of</strong> effective chelating agents<br />

for plant cultures, root cultures were grown on media<br />

with a low pH. Tomato roots were, for instance,<br />

unable to grow on media similar to those <strong>of</strong> White<br />

(1943a), when the pH rose to 5.2 (Street et al., 1951).<br />

Boll and Street (1951) were able to show that the<br />

depression <strong>of</strong> growth at high pH was due to the loss<br />

<strong>of</strong> Fe from the medium and that it could be overcome<br />

by adding a chelated form <strong>of</strong> iron (see chapter 3).<br />

Using FeEDTA, Torrey (1956) discovered that<br />

isolated pea roots [grown on a medium containing<br />

Bonner and Devirian (1939) A macronutrients, which<br />

do not contain NH4 + ], grew optimally at pH 6.0-6.4<br />

but were clearly inhibited at pH 7.0 or greater; and<br />

Street (1969) reported that growth <strong>of</strong> tomato roots<br />

could be obtained between pH 4.0 and pH 7.2, if<br />

EDTA was present in the medium. Because agar does<br />

not gel properly when the initial pH <strong>of</strong> the medium is<br />

adjusted to 4, it is necessary to use liquid media for<br />

low pH cultures; or employ another gelling agent, or<br />

a mechanical support.<br />

Fig. 4.4 Effect <strong>of</strong> medium pH on adventitious rooting from apple stem disks. <strong>The</strong> pH at the x-axis is the pH as measured at the<br />

start <strong>of</strong> culture after autoclaving (cf. Fig. 4.3). (from de Klerk et al., 2007).


Some other cultures may also be beneficially<br />

started at low pH, which may indicate that the tissues<br />

have an initial requirement for NO3 – . Embryogenesis<br />

<strong>of</strong> Pelargonium was induced more effectively if MS,<br />

or other media, were adjusted to pH 4.5-5.0 before<br />

autoclaving (rather than pH 5.5 and above)<br />

(Marsolais et al., 1991).<br />

5.3.2. Differentiation and Morphogenesis<br />

Differentiation and morphogenesis are frequently<br />

found to be pH-dependent. Xylogenesis depends on<br />

the medium pH (Khan et al., 1986). <strong>The</strong> growth <strong>of</strong><br />

callus and the formation <strong>of</strong> adventitious organs from<br />

thin cell layers excised from superficial tissues <strong>of</strong> the<br />

inflorescence rachis <strong>of</strong> Nicotiana, depended on the<br />

initial pH <strong>of</strong> Linsmaier and Skoog (1965) medium<br />

containing 0.5 μM IBA and 3 μM kinetin (Table 4.9)<br />

(Mutaftschiev et al., 1987). Pasqua et al. (2002)<br />

reported many quantitative effects <strong>of</strong> pH during<br />

regeneration from tobacco thin cell layers. <strong>The</strong> types<br />

<strong>of</strong> callus produced from the plumules <strong>of</strong> Hevea<br />

seedlings differed according to the pH <strong>of</strong> the medium<br />

devised by Chua (1966). S<strong>of</strong>t and spongy callus<br />

formed at acid (5.4) or alkaline (8.0) pH. A compact<br />

callus was obtained between pH 6.2 and 6.8.<br />

5.3.3. Adventitious root formation<br />

<strong>The</strong>re are several reports in the literature which<br />

show that the pH <strong>of</strong> the medium can influence root<br />

formation <strong>of</strong> some plants in vitro. A slightly acid pH<br />

seems to be preferred by most species. Zatkyo and<br />

Molnar (1986), who showed a close correlation<br />

between the acidity <strong>of</strong> the medium (pH 7.0 to 3.0)<br />

and the rooting <strong>of</strong> Vitis, Ribes nigrum and Aronia<br />

melancarpa shoots, suggested that this was because<br />

acidity is necessary for auxin action.<br />

Sharma et al. (1981) found it advantageous to<br />

reduce the pH <strong>of</strong> the medium to 4.5 to induce rooting<br />

<strong>of</strong> Bougainvillea shoots and a reduction <strong>of</strong> the pH <strong>of</strong><br />

MS medium to 4.0 (accompanied by incubation in the<br />

dark) was required for reliable root formation <strong>of</strong> two<br />

Santalum species (Barlass et al., 1980) and Correa<br />

decumbens and Prostanthera striatifolia (Williams<br />

et al., 1984; 1985). Other Australian woody species<br />

rooted satisfactorily at pH 5.5 and pH 4 was<br />

inhibitory (Williams et al., loc. cit.). Shoots from<br />

carnation meristem tips rooted more readily at pH 5.5<br />

than pH 6.0 (Stone, 1963), and rooting <strong>of</strong> excised<br />

potato buds was best at pH 5.7, root formation being<br />

inhibited at pH 4.8 and at pH 6.2 or above (Mellor<br />

and Stace-Smith, 1969). Direct root formation on<br />

Nautilocalyx leaf segments was retarded if a modified<br />

MS medium containing IAA was adjusted initially to<br />

Chapter 4 151<br />

an acid pH (3.5 or 4.0) or a neutral pH (6.5). Good<br />

and rapid root formation occurred when the medium<br />

was adjusted to between pH 5.0 to 6.3 (Venverloo,<br />

1976). De Klerk et al (2007), working with apple<br />

stem slices, found only a small effect <strong>of</strong> pH on<br />

rooting (Fig. 4.4): when the pH was set before<br />

autoclaving at 4.5 (after autoclaving the pH was<br />

4.54), the number <strong>of</strong> roots was 4.5, and with the pH<br />

set at 8.0 (after autoclaving the pH had dropped to<br />

5.65), the number <strong>of</strong> roots per slice increased to 7. In<br />

medium buffered with MES, the maximum number<br />

<strong>of</strong> roots was formed at pH 4.4 (measured after autoclaving).<br />

In these experiments, the dose-response<br />

curve for root number did not correspond with the<br />

effect <strong>of</strong> pH on IAA uptake. Such discrepancy<br />

between the effects <strong>of</strong> the pH on uptake and root<br />

number, was also reported by Harbage, Stimart and<br />

Auer (1998).<br />

Direct formation <strong>of</strong> roots from Lilium auratum<br />

bulb scales occurred when MS medium was adjusted<br />

within the range 4-7 but was optimal at pH 6. <strong>The</strong> pH<br />

range for adventitious bulblet formation in this plant<br />

was from 4 to 8, but most bulblets were produced<br />

when the initial pH was between 5 and 7 (Takayama<br />

and Misawa, 1979).<br />

Substrates which are to acid or too alkaline can<br />

adversely affect rooting ex vitro.<br />

5.3.4. Embryogenesis<br />

Smith and Krikorian (1989) discovered that preglobular<br />

stage pro-embryos (PGSP) <strong>of</strong> carrot could be<br />

made to proliferate from tissues capable <strong>of</strong> direct<br />

embryo formation, with no auxin, on a medium<br />

containing 1-5 mM NH4 + (and no nitrate). Somatic<br />

embryos were formed when this tissue was moved to<br />

MS medium. <strong>The</strong> pH <strong>of</strong> the ‘ammonium-nitrogen’<br />

medium fell from 5.5 to 4.0 in each subculture<br />

period, and it was later found (Smith and Krikorian,<br />

1990a,b) that culture on a medium <strong>of</strong> low pH was<br />

essential for the maintenance <strong>of</strong> PGSP cultures.<br />

Sustained culture at a pH equal or greater than 5.7,<br />

with no auxin, allowed somatic embryo development.<br />

A similar observation to that <strong>of</strong> Smith and Krikorian<br />

had been made by (Stuart et al., 1987). Although the<br />

pH <strong>of</strong> a suspension culture <strong>of</strong> alfalfa ‘Regen-S’ cells<br />

in a modified Schenk and Hildebrandt (1972)<br />

medium with 15 mM NH4 + was adjusted to 5.8, it<br />

quickly fell back to pH 4.4-5.0 in a few hours. <strong>The</strong><br />

pH then gradually increased as somatic embryos were<br />

produced, until at day 14 it was 5.0. In certain<br />

suspension cultures, the pH was titrated daily to 5.5,<br />

but on each occasion soon returned to nearly the same<br />

pH as that in flasks which were untouched. Even so,


152 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

the pH-adjusted suspensions produced more embryos<br />

than the controls.<br />

<strong>The</strong> ammonium ion has been found to be essential<br />

for embryogenesis. Is one <strong>of</strong> its functions to reduce<br />

the pH <strong>of</strong> the medium through rapid uptake and<br />

metabolism, thereby facilitating the uptake <strong>of</strong> nitrate,<br />

upon which embryogenesis is really dependent?<br />

Embryogenesis from leaf explants <strong>of</strong> Ostericum<br />

koreanum, was found to depend strongly on pH (Cho<br />

et al 2003). As the explants were cultured continuously<br />

with NAA, it is possible that the observed relationship<br />

was caused by differential NAA uptake. This is also<br />

suggested by the slower rate <strong>of</strong> embryo development<br />

seen at low pH, because this would be expected where<br />

there is a high internal NAA concentration.<br />

6. LIQUID MEDIA AND SUPPORT SYSTEMS<br />

<strong>The</strong> nutritional requirements <strong>of</strong> plant cultures can<br />

be supplied by liquid media but growth in liquid<br />

medium may be retarded and development affected<br />

by oxygen deprivation and hyperhydration. <strong>The</strong><br />

oxygen concentration <strong>of</strong> liquid media is <strong>of</strong>ten<br />

insufficient to meet the respiratory requirements <strong>of</strong><br />

submerged cells and tissues. It can be increased<br />

either by raising the oxygen concentration <strong>of</strong> the<br />

medium or placing cells or tissues in direct contact<br />

with air. If the water potential <strong>of</strong> the medium is<br />

greater (less negative) than that <strong>of</strong> a cell, water flows<br />

into the cell and the vacuole becomes distended.<br />

Cells and tissues affected in this way are described as<br />

hyperhydric. Shoots <strong>of</strong>ten show physiological disturbances<br />

with symptoms that can be recognised<br />

visually (Chapter 13) (Debergh et al., 1992; Gaspar<br />

et al., 1987; Preece and Sutter, 1991; Ziv, 1991). <strong>The</strong><br />

term hyperhydric is preferable to the previously used<br />

term ‘vitrified’ for reasons explained by Debergh<br />

et al. (1992). Water potential is determined by<br />

osmotic potential <strong>of</strong> the solutes in the medium and, in<br />

the case <strong>of</strong> a gelled medium, by the matric potential<br />

<strong>of</strong> the gel (Section 4.1) (Beruto et al., 1995; Fujiwara<br />

and Kozai, 1995; Owens and Wozniak, 1991).<br />

Reductions in hyperhydricity can be achieved by<br />

increasing the concentrations <strong>of</strong> the solutes and the<br />

gel. Hyperhydricity may also be reduced through<br />

evaporation <strong>of</strong> water from tissues if they are placed in<br />

contact with air.<br />

Contact <strong>of</strong> cultured tissues with air, to alleviate<br />

problems <strong>of</strong> hyperhydricity and hypoxia, can be<br />

achieved by the use <strong>of</strong> either porous or semi-solid<br />

(gelled) supports. <strong>The</strong> advantage <strong>of</strong> supports, as<br />

opposed to thin layers <strong>of</strong> liquid medium, is that<br />

tissues can be placed in a sufficient volume <strong>of</strong><br />

medium to prevent depletion <strong>of</strong> nutrients and allow<br />

for the dispersion <strong>of</strong> any toxins that might be<br />

produced by the plant tissues. <strong>The</strong> relative<br />

advantages <strong>of</strong> liquid medium, solid supports or gelled<br />

media, varies with the type <strong>of</strong> material being<br />

cultured, the purpose <strong>of</strong> the culture and the scale <strong>of</strong><br />

culture, as discussed below.<br />

6.1. LIQUID MEDIA<br />

Liquid medium, without supporting structures, is<br />

used for the culture <strong>of</strong> protoplasts, cells or root<br />

systems for the production <strong>of</strong> secondary metabolites,<br />

and the propagation <strong>of</strong> somatic embryos,<br />

meristematic nodules, microtubers and shoot clusters.<br />

In liquid medium, these cultures <strong>of</strong>ten give faster<br />

growth rates than on agar-solidifed medium.<br />

<strong>Culture</strong>s may be fully or only partially immersed in<br />

the medium.<br />

Aeration <strong>of</strong> liquid medium in stationary Petri<br />

dishes is sometimes adequate for the culture <strong>of</strong><br />

protoplasts and cells because <strong>of</strong> the shallow depth <strong>of</strong><br />

the medium, but may still be suboptimal. Anthony<br />

et al. (1995) cultured protoplasts <strong>of</strong> cassava, in liquid<br />

medium in Petri dishes with an underlying layer <strong>of</strong><br />

agarose in which glass rods were embedded<br />

vertically. Sustained protoplast division was<br />

observed in the cultures with glass rods but not in the<br />

controls without glass rods. <strong>The</strong> authors suggested<br />

that the glass rods extended the liquid meniscus,<br />

where the cell colonies were clustered thus causing<br />

gaseous exchange between the liquid and the<br />

atmosphere above to be facilitated.<br />

Anthony et al. (1997) cultured protoplasts <strong>of</strong><br />

Passiflora and Petunia in 30 ml glass bottles<br />

containing protoplast suspensions in 2 ml aliquots,<br />

either alone or in the presence <strong>of</strong> the oxygen carriers<br />

Erythrogen or oxygenated Perfluorodecalin. Cell<br />

division in each <strong>of</strong> the two species was stimulated by<br />

both oxygen carriers.<br />

Laboratory-scale experimentation on immersed<br />

cultures <strong>of</strong> cells, tissues and organs, may be carried<br />

out in 125 ml or 250 ml Erlenmeyer flasks. Largescale<br />

cultures are usually carried out in bioreactors<br />

with a capacity <strong>of</strong> 1 litre or more. <strong>The</strong> concentration<br />

<strong>of</strong> oxygen in the medium is raised by oxygen in the<br />

gas phase above and air bubbles inside the liquid.<br />

Increasing the oxygen concentration and circulation<br />

<strong>of</strong> the medium is facilitated in flasks by the use <strong>of</strong><br />

gyratory shakers and in bioreactors by stirring and/or<br />

bubbling air through the medium (Chapter 1). <strong>The</strong>


use <strong>of</strong> bioreactors <strong>of</strong>ten involves the automated<br />

adjustment <strong>of</strong> the culture medium. <strong>The</strong> design <strong>of</strong><br />

bioreactors for plant cells and organs was reviewed<br />

by Doran (1993) and the use <strong>of</strong> shake-flasks and<br />

bioreactors for the scale-up <strong>of</strong> embryogenic plant<br />

suspension cultures has been reviewed by Tautorus<br />

and Dunstan (1995). <strong>The</strong> importance <strong>of</strong> oxygen<br />

concentration in bioreactors can be illustrated by an<br />

investigation into the growth <strong>of</strong> hairy roots <strong>of</strong> Atropa<br />

belladonna (Yu and Doran, 1994). <strong>The</strong>y found that<br />

no growth occurred at oxygen tensions <strong>of</strong> 50% air<br />

saturation but between 70% and 100% air saturation,<br />

total root length and the number <strong>of</strong> root tips increased<br />

exponentially. Hyperhydricity in liquid cultures may<br />

be avoided by adding to the medium osmoregulators,<br />

such as mannitol, maltose and sorbitol, and inhibitors<br />

<strong>of</strong> gibberellin biosynthesis including ancymidol and<br />

paclobutrazol (Ziv, 1989).<br />

<strong>Plant</strong>lets and microtubers can be cultured by<br />

partial submersion in liquid medium. One method <strong>of</strong><br />

aerating tissues is by the automated flooding and<br />

evacuation <strong>of</strong> tissues by liquid medium. This method<br />

has been used to produce microtubers <strong>of</strong> potato from<br />

single node cuttings (Teisson and Alvard, 1999). An<br />

alternative approach to aeration is to apply the liquid<br />

medium over the plant tissues as a nutrient mist. For<br />

example, Kurata et al. (1991) found that nodes <strong>of</strong><br />

potato grew better in nutrient mist than on agar-based<br />

cultures.<br />

6.2. SUPPORT BY SEMI-SOLID MATRICES<br />

Gelled media provide semi-solid, supporting<br />

matrices that are widely used for protoplast, cell,<br />

tissue and organ culture. Agar, agarose, gellan gums<br />

and various other products have been used as gelling<br />

agents.<br />

6.2.1. Agar<br />

Agar is very widely employed for the preparation<br />

<strong>of</strong> semi-solid culture media. It has the advantages<br />

that have made it so widely used for the culture <strong>of</strong><br />

bacteria, namely :-<br />

• it forms gels with water that melt at approx.<br />

100°C and solidify at approx. 45°C, and are thus<br />

stable at all feasible incubation temperatures;<br />

• gels are not digested by plant enzymes;<br />

• agar does not strongly react with media<br />

constituents.<br />

To ensure adequate contact between tissue and<br />

medium, a lower concentration <strong>of</strong> agar is generally<br />

used for plant cultures than for the culture <strong>of</strong> bacteria.<br />

<strong>Plant</strong> media are not firmly gelled, but only rendered<br />

Chapter 4 153<br />

semi-solid. Depending on brand, concentrations <strong>of</strong><br />

between 0.5-1.0% agar are generally used for this<br />

purpose. Agar is thought to be composed <strong>of</strong> a<br />

complex mixture <strong>of</strong> related polysaccharides built up<br />

from galactose. <strong>The</strong>se range from an uncharged<br />

neutral polymer fraction, agarose, that has the<br />

capacity to form strong gels, to highly charged<br />

anionic polysaccharides, sometimes called<br />

agaropectins, which give agar its viscosity. Agar is<br />

extracted from species <strong>of</strong> Gelidium and other red<br />

algae, collected from the sea in several different<br />

countries. It varies in nature according to country <strong>of</strong><br />

origin, the year <strong>of</strong> collection and the way in which it<br />

has been extracted and processed. <strong>The</strong> proportion <strong>of</strong><br />

agarose to total polysaccharides can vary from 50 to<br />

90% (Adrian and Assoumani, 1983). Agars contain<br />

small amounts <strong>of</strong> macro- and micro-elements;<br />

particularly calcium, sodium, potassium, and<br />

phosphate (Beruto et al., 1995; Debergh, 1983;<br />

Scherer et al., 1988), carbohydrates, traces <strong>of</strong> amino<br />

acids and vitamins (Day, 1942) that affect the<br />

osmotic and nutrient characteristics <strong>of</strong> a gel. <strong>The</strong>y<br />

also contain phenolic substances (Scherer et al.,<br />

1988) and less pure grades may contain long chain<br />

fatty acids, inhibitory to the growth <strong>of</strong> some bacteria.<br />

As agar can be the most expensive component <strong>of</strong><br />

plant media, there is interest in minimising its<br />

concentration. Concentrations <strong>of</strong> agar can be<br />

considered inadequate if they do not support explants<br />

or lead to hyperhydricity. Hyperhydricity decreases<br />

as the agar concentration is raised but there may be<br />

an accompanying reduction in the rate <strong>of</strong> growth. For<br />

example, Debergh et al. (1981) found that shoot<br />

cultures <strong>of</strong> Cynara scolymus were hyperhydric on<br />

medium containing 0.6% Difco ‘Bacto’ agar. No<br />

hyperhydricity occurred on medium containing 1.1%<br />

agar but shoot proliferation was reduced. Likewise,<br />

Hakkaart and Versluijs (1983) found that shoots <strong>of</strong><br />

carnation were hyperhydric on medium containing<br />

0.6% agar whereas growth was severely reduced on<br />

medium containing 1.2%. During studies to optimise<br />

the production <strong>of</strong> morphogenic callus from leaf discs<br />

<strong>of</strong> sugarbeet, Owens and Wozniak (1991) found large<br />

differences in the numbers <strong>of</strong> somatic embryos and<br />

shoots according to the gelling agent employed.<br />

<strong>The</strong>y found that water availability, determined by gel<br />

matric potential, was the dominant factor involved.<br />

When they adjusted the concentrations <strong>of</strong> the gelling<br />

agents to give media <strong>of</strong> equal gel matric potential,<br />

somatic embryos and shoots were found in similar<br />

numbers on Bacto agar (0.7%), HGT agarose<br />

(0.46%), Phytagar (0.62%) and Gelrite (0.12%).


154 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

Various brands and grades <strong>of</strong> agar are available<br />

commercially. <strong>The</strong>se differ in the amounts <strong>of</strong><br />

impurities they contain and their gelling capabilities.<br />

<strong>The</strong> gelling capacity <strong>of</strong> Difco brands <strong>of</strong> agar<br />

increased with increasing purity i.e. ‘Noble’ ><br />

‘Purified’ > ‘Bacto’ (Debergh, 1983). After<br />

impurities <strong>of</strong> agar were removed by sealing agar in a<br />

semi-permeable bags and washing in deionized water,<br />

the water potentials <strong>of</strong> gels <strong>of</strong> three brands were<br />

substantially lower than unwashed gels (Beruto et al.,<br />

1995). A ten-fold difference in the regeneration rate<br />

<strong>of</strong> sugar cane was observed by Anders et al. (1988)<br />

on media gelled with the best and least effective <strong>of</strong><br />

seven brands <strong>of</strong> agar. Scholten and Pierik (1998)<br />

investigated the different growth characteristics <strong>of</strong><br />

seven different agar brands on the growth <strong>of</strong> axillary<br />

shoots, adventitious shoots and adventitious roots <strong>of</strong><br />

rose, lily and cactus. <strong>The</strong>y concluded that no single<br />

bioassay could identify ‘good or bad’ agars for a<br />

large group <strong>of</strong> plant species but Merk 1614, Daishin,<br />

MC29, and BD Purified gave the best results in most<br />

experiments. ‘Daishin’ showed no batch-to-batch<br />

variations. <strong>The</strong>y found no relationship between price<br />

and quality <strong>of</strong> the brands <strong>of</strong> agar.<br />

Agarose. Agarose is the high gel strength moiety<br />

<strong>of</strong> agar. It consists <strong>of</strong> β-D(1→3) galactopyranose and<br />

3,6-anhydro-α-L(1 → 4) galactopyranose polymer<br />

chains <strong>of</strong> 20-160 monosaccharide units alternatively<br />

linked to form double helices. <strong>The</strong>re are also several<br />

different agaro-pectin products available, which have<br />

been extracted from agar and treated to remove most<br />

<strong>of</strong> the residual sulphate side groupings (Shillito et al.,<br />

1983). Because preparation involves additional<br />

processes, agarose is much more expensive than agar<br />

and its use is only warranted for valuable cultures,<br />

including protoplast and anther culture. Brands may<br />

differ widely in their suitability for these applications.<br />

Concentrations <strong>of</strong> 0.4-1.0% are used. Agarose<br />

derivatives are available which melt and gel at<br />

temperatures below 30°C, making them especially<br />

suitable for testing media ingredients that are heatlabile,<br />

or for embedding protoplasts. Low meltingpoint<br />

agarose is prepared by introducing<br />

hydroxyethyl groups into the agarose molecules<br />

(Shillito et al., 1983). Another advantage <strong>of</strong> agarose<br />

over agar lies in the removal <strong>of</strong> toxic components <strong>of</strong><br />

agar during its preparation. Bolandi et al. (1999)<br />

preferred to embed protoplasts in agarose, rather than<br />

use liquid medium, because in agarose the semi-solid<br />

matrix applies a direct pressure on the plasma<br />

membrane <strong>of</strong> the protoplasts. <strong>The</strong>y mixed protoplasts<br />

<strong>of</strong> sunflower with 0.5% agarose, pipetted 50 μl<br />

droplets <strong>of</strong> the mixture into Petri dishes and covered<br />

them with a thin layer <strong>of</strong> culture medium. A similar<br />

method was used to culture protoplasts <strong>of</strong> Dioscorea<br />

by Tor et al. (1999). <strong>The</strong> use <strong>of</strong> droplets has the<br />

advantage that a high plating density can be achieved<br />

in the droplets, while exposing the protoplasts to a<br />

larger reservoir <strong>of</strong> culture medium in the liquid phase.<br />

Bishoi et al. (2000) initially cultured anthers <strong>of</strong><br />

Basmati rice on liquid medium, then used 1.0%<br />

agarose to culture calli derived from microspores.<br />

6.2.2. Gellan gum<br />

Gellan gum is a widely used gelling agent in plant<br />

tissue culture, that is marketed under various trade<br />

names including Gelrite, Phytagel and Kelcogel. It<br />

is an exopolysaccharide that encapsulates cells <strong>of</strong> the<br />

bacterium Sphingomonas paucimobilis (= Auromonas<br />

elodea = Pseudomonas elodea), from which it is<br />

obtained by industrial fermentation. <strong>The</strong> structure,<br />

physico-chemical properties and the rheology <strong>of</strong><br />

solutions <strong>of</strong> gellan gum and related polysaccharides<br />

has been reviewed by Banik et al. (2000). Gellan<br />

gum consists <strong>of</strong> a linear repeating tetrasaccharide <strong>of</strong><br />

D-glucose, D-glucuronic acid, D-glucose and Lrhamnose.<br />

Heating solutions <strong>of</strong> gellan gum in<br />

solutions that contain cations, such as K + , Ca 2+ , Mg 2+ ,<br />

causes the polysaccharide to form a gel in which the<br />

polymers form a half-staggered parallel double helix.<br />

<strong>The</strong> commercial deacetylated and purified<br />

polysaccharide forms a firm non-elastic gel. <strong>The</strong> gel<br />

sets rapidly at a temperature, determined by the<br />

concentrations <strong>of</strong> the polysaccharide and the cation,<br />

which varies between 35-50 °C (Kang et al., 1982).<br />

<strong>The</strong> commercial product contains significant<br />

quantities <strong>of</strong> potassium, sodium, calcium and<br />

magnesium (Pasqualetto et al., 1988a,b; Scherer<br />

et al., 1988) but is said to be free <strong>of</strong> the organic<br />

impurities found in agar. It is unclear whether or not<br />

these cations remain fully available to plant cultures.<br />

Some researchers (Gawel et al., 1986; 1990;<br />

Trolinder and Goodin, 1987; Umbeck et al., 1987)<br />

add an extra 750 mg l –1 MgCl2 to a medium<br />

containing MS salts to aid the gelling <strong>of</strong> 1.6% Gelrite.<br />

In most other reports on the use <strong>of</strong> Gelrite, cations in<br />

the medium have been sufficient to produce a gel.<br />

Beruto et al. (1995) found that 0.12 % Gelrite and<br />

0.7% Bacto agar have equivalent matric potential<br />

and support equivalent adventitious regeneration in<br />

leaf discs <strong>of</strong> sugarbeet. As gellan gum is used in<br />

lower concentration than agar, the cost per litre <strong>of</strong><br />

medium is less. It produces a clear gel in which plant<br />

tissues can be more easily seen and microbial


contamination more easily detected than in agar gels.<br />

It has proved to be a suitable gelling agent for tissue<br />

cultures <strong>of</strong> many herbaceous plants and there are<br />

reports <strong>of</strong> its successful use for callus culture, the<br />

direct and indirect formation <strong>of</strong> adventitious organs<br />

and somatic embryos, shoot culture <strong>of</strong> herbaceous<br />

and semi-woody species and the rooting <strong>of</strong> plantlets.<br />

In most cases the results have been as good as, and<br />

sometimes superior to, those obtainable on agarsolidified<br />

media. Anders et al. (1988) described the<br />

regeneration <strong>of</strong> greater numbers <strong>of</strong> plants from <strong>of</strong><br />

sugar cane on Gelrite than on the most productive<br />

brand <strong>of</strong> agar, and Koetje et al. (1989) obtained more<br />

somatic embryos from callus cultures <strong>of</strong> rice on<br />

media solidified with Gelrite than with Bacto-agar.<br />

Shoot cultures, particularly <strong>of</strong> some woody species,<br />

are liable to become hyperhydric if Gelrite, like agar,<br />

is used at too low a concentration. <strong>The</strong>re are sharp<br />

differences in the response <strong>of</strong> different species to<br />

concentration <strong>of</strong> Gelrite. Turner and Singha (1990)<br />

found the highest rate <strong>of</strong> shoot proliferation in Geum<br />

occurred on 0.2% and in Malus on 0.4%. Garin et al.<br />

(2000) obtained more mature somatic embryos <strong>of</strong><br />

Pinus strobus on gellan gum at 1% concentration <strong>of</strong><br />

than at 0.6%. Pasqualetto et al. (1986a,b) used<br />

mixtures <strong>of</strong> Gelrite (0.1-0.15%) and Sigma @ agar<br />

(0.2-0.3%) to prevent the hyperhydricity that<br />

occurred in shoot cultures <strong>of</strong> Malus domestica ‘Gala’<br />

on media solidified with Gelrite alone. Nairn (1988)<br />

used a mixture <strong>of</strong> Gelrite (0.194%) and Difco<br />

‘Bacto’ agar (0.024%) to prevent the hyperhydricity<br />

that occurred in shoot cultures <strong>of</strong> Pinus radiata on<br />

medium gelled with 0.2% Gelrite alone.<br />

6.2.3. Alginates<br />

Alginic acid is a binary linear heteropolymer 1,4β-D-mannuronic<br />

acid and 1,4-α-L-guluronic acid<br />

(Larkin et al., 1988). It is extracted from various<br />

species <strong>of</strong> brown algae. When the sodium salt is<br />

exposed to calcium ions, gelation occurs. Alginates<br />

are widely used for protoplast culture and to<br />

encapsulate artificial seed.<br />

Protoplasts embedded in beads or thin films <strong>of</strong><br />

alginate can plated densely while yet exposed to a<br />

large pool <strong>of</strong> medium that dilutes inhibitors and toxic<br />

substances. Embedded protoplasts can be surrounded<br />

by nurse cells, either free in the surrounding medium<br />

or separated by filters or membranes. Alginate has<br />

the advantages over agarose that protoplasts do not<br />

have to be exposed to elevated temperatures when<br />

they are mixed with the gelling agent and the gel can<br />

be liquified by adding sodium citrate, releasing<br />

Chapter 4 155<br />

protoplasts or cell colonies for transfer to other<br />

media. <strong>The</strong> method <strong>of</strong> embedding protoplasts in<br />

beads as employed by Larkin et al. (1988) involved<br />

mixing the protoplast suspensions with an aqueous<br />

solution <strong>of</strong> sodium alginate and dropping it, through a<br />

needle, into a solution <strong>of</strong> calcium chloride. Beads<br />

containing the protoplasts were formed when the<br />

alginate made contact with the calcium ions. Beads<br />

with a final concentration <strong>of</strong> 1.5% sodium alginate<br />

were washed and cultured in liquid medium on an<br />

orbital shaker. Protoplasts may also be captured in<br />

thin layers <strong>of</strong> alginate (Dovzhenko et al., 1998).<br />

Synthetic seeds (synseeds, somatic seeds)<br />

encapsulated in alginate (Fig. 4.4), can be prepared<br />

from somatic embryos (Timbert et al., 1995), shoot<br />

tips (Maruyama et al., 1998), apical and axillary buds<br />

(Piccioni and Standardi, 1995), single nodes<br />

(Piccioni, 1997), and cell aggregates from hairy roots<br />

(Repunte et al.,1995). <strong>The</strong> uses <strong>of</strong> sythetic seeds<br />

include direct planting into soil, storage <strong>of</strong> tissues and<br />

transfer <strong>of</strong> materials between laboratories under<br />

sterile conditions. <strong>The</strong> methods <strong>of</strong> encapsulation <strong>of</strong><br />

somatic embryos <strong>of</strong> carrot were described by Timbert<br />

et al. (1995). Torpedo-shaped embryos were mixed<br />

with a 1% sodium alginate solution. <strong>The</strong> mixture was<br />

dropped into a solution 100 mM calcium chloride in<br />

10% sucrose. <strong>The</strong> beads (3-3.5 mm in diameter) were<br />

then rinsed in a 10% sucrose solution.<br />

6.2.4. Starch<br />

Sorvari (1986a,c) found that plantlets formed in<br />

higher frequencies in anther cultures <strong>of</strong> barley on a<br />

medium solidified with 5% corn starch or barley<br />

starch rather than with agar. Corn starch only formed<br />

a weak gel and it was necessary to place a polyester<br />

net on its surface to prevent the explants from<br />

sinking. Sorvari (1986b) found that it took 5-14<br />

weeks for adventitious shoots to form on potato discs<br />

on agar-solidified medium but only 3 weeks on<br />

medium containing barley starch. Henderson and<br />

Kinnersley (1988) found that embryogenic carrot<br />

callus cultures grew slightly better on media gelled<br />

with 12% corn starch than on 0.9% Difco `Bacto’<br />

agar.<br />

6.2.5. ‘Kappa’--carrageenan<br />

Carrageenan is a product <strong>of</strong> sea weeds <strong>of</strong> the<br />

genus Euchema and the kappa form has strong<br />

gelling properties. Like gellan gum, kappacarrageenan<br />

requires the presence <strong>of</strong> cations for<br />

gelation. In Linsmaier and Skoog (1965) medium at<br />

0.6% w/v, the gel strength was slightly less than that<br />

<strong>of</strong> 0.2% Gelrite or 0.8% <strong>of</strong> extra pure agar (Ichi et al.,


156 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

1986). Chauvin et al. (1999) found that regeneration<br />

from cultures <strong>of</strong> tulip, gladiolus and tobacco shoots<br />

was possible in the presence <strong>of</strong> 200 mg l –1<br />

kanamycin, whereas in several other gelling agents a<br />

concentration <strong>of</strong> 100 mg l –1 inhibited regeneration.<br />

6.2.6. Pectins<br />

A mixture <strong>of</strong> pectin and agar can be a less<br />

expensive substitute for agar. For example, a semisolid<br />

medium consisting <strong>of</strong> 0.2% agar plus 0.8-1.0%<br />

pectin, was employed for shoot culture <strong>of</strong> strawberry<br />

and some other plants (Zimmerman, 1979).<br />

6.2.7. Other gelling agents<br />

Battachary et al. (1994) found sago (from<br />

Metroxylon sagu) and isubgol (from <strong>Plant</strong>ago ovata)<br />

were satisfactory substitutes for agar at, respectively,<br />

one eighth and one tenth <strong>of</strong> the cost <strong>of</strong> Sigma purified<br />

agar A7921.<br />

6.3. POROUS SUPPORTS<br />

Aeration <strong>of</strong> the tissues on a porous substrate is<br />

usually better than it would be in agar or static liquid.<br />

Chin et al. (1988) used a buoyant polypropylene<br />

membrane floated on top <strong>of</strong> a liquid medium to<br />

culture cells and protoplasts <strong>of</strong> Asparagus. <strong>The</strong><br />

membrane (Celgard 3500, Questar Corp., Charlotte,<br />

N.C.) has a pore size <strong>of</strong> 0.04 mm and is autoclavable.<br />

Conner and Meredith (1984) found that cells grew<br />

more rapidly on filter papers laid over polyurethane<br />

foam pads saturated with medium than on agar.<br />

Young et al. (1991) supported shoots <strong>of</strong> tomato over<br />

liquid culture medium on a floating microporous<br />

polypropylene membrane and entrained the growing<br />

shoots through polypropylene netting. <strong>The</strong>y reported<br />

opportunities for the development <strong>of</strong> this method for<br />

mechanisation by mass handling. Membrane rafts<br />

were also used by Teng (1997) and Watad et al.<br />

(1995,1996).<br />

Cheng and Voqui (1977) and Cheng (1978) used<br />

polyester fleece to support cultures <strong>of</strong> Douglas fir that<br />

were irrigated with liquid media in Petri dishes.<br />

<strong>Plant</strong>lets that were regenerated from cotyledon<br />

explants were cultured on 3 mm-thick fabric. When a<br />

protoplast suspension was dispersed over 0.5 mmthick<br />

fabric, numerous colonies were produced in 12<br />

days, whereas in the absence <strong>of</strong> the support, cell<br />

colonies failed to proliferate beyond the 20 cell stage.<br />

A major advantage in using this type <strong>of</strong> fabric support<br />

is that media can be changed simply and quickly<br />

without disturbing the tissues. <strong>The</strong> system has also<br />

been used for protoplast culture <strong>of</strong> other plants (e.g.<br />

by Russell and McCown, 1986).<br />

Heller and Gautheret (1949) found that tissues<br />

could be satisfactorily cultured on pieces <strong>of</strong> ashless<br />

filter paper moistened by contact with liquid medium.<br />

Very small explants, such as meristem tips, that might<br />

be lost if placed in a rotated or agitated liquid medium,<br />

can be successfully cultured if placed on an M-shaped<br />

strip <strong>of</strong> filter paper (sometimes called a ‘Heller’<br />

support). When the folded paper is placed in a tube <strong>of</strong><br />

liquid medium, the side arms act as wicks (Goodwin,<br />

1966). This method <strong>of</strong> support ensures excellent tissue<br />

aeration but the extra time required for preparation and<br />

insertion has meant that paper wicks are only used for<br />

special purposes such as the initial cultural stages <strong>of</strong><br />

single small explants which are otherwise difficult to<br />

establish. Whether explants grow best on agar or on<br />

filter paper supports, varies from one species <strong>of</strong> plant<br />

to another. Davis et al. (1977) found that carnation<br />

shoot tips grew less well on filter paper bridges than on<br />

0.6% agar but axillary bud explants <strong>of</strong> Leucospermum<br />

survived on bridges but not on agar.<br />

Paper was also used in the construction <strong>of</strong> plugs<br />

(marketed by Ilacon Ltd, Tonbridge, UK TN9 1NR)<br />

known as Sorbarods (Roberts and Smith, 1990). <strong>The</strong>se<br />

are cylindrical (20 mm in length and 18 mm in<br />

diameter) and consist <strong>of</strong> cold-crimped cellulose,<br />

longitudinally folded, wrapped in cellulose paper. <strong>The</strong><br />

plug has a porosity (total volume – volume <strong>of</strong><br />

cellulose) <strong>of</strong> 94.2% and high capillarity, so that the<br />

culture medium is efficiently drawn up into the plug,<br />

leaving the sides <strong>of</strong> the plug in direct contact with air.<br />

Roots permeate the plugs and are protected by the<br />

cellulose during transfer to soil. <strong>Plant</strong>lets <strong>of</strong><br />

chrysanthemum in Sorbarods formed longer stems,<br />

larger leaves, more roots, and developed greater fresh<br />

mass, dry weights and fresh to dry mass ratios than<br />

plantlets in agar-solidified medium (Roberts and<br />

Smith, 1990). <strong>The</strong> greater fresh to dry mass ratio<br />

indicates that contact with liquid medium led to greater<br />

hydration <strong>of</strong> tissues. This was subsequently controlled<br />

by the inclusion <strong>of</strong> a growth retardant, paclobutrazol (1<br />

mg l –1 ), in the culture medium (Smith et al., 1990a).<br />

Other porous materials that have been used to support<br />

plant growth include rockwool (Woodward et al.,<br />

1991; Tanaka et al., 1991), polyurethane foam<br />

(Gutman and Shiryaeva, 1980; Scherer et al., 1988),<br />

vermiculite (Kirdmanee et al., 1995), a mixture <strong>of</strong><br />

vermiculite and Gelrite (Jay-Allemand et al., 1992).<br />

Afreen-Zobayed et al. (2000) cultured sweet<br />

potato, on sugar-free medium in autotrophic<br />

conditions, on mixtures <strong>of</strong> paper pulp and vermiculite<br />

in various proportions. Optimal growth was obtained<br />

on a mixture containing 70% paper pulp. On this


mixture, the fresh mass <strong>of</strong> plantlets was greater by a<br />

factor <strong>of</strong> 2.7 than on agar-solidified medium.<br />

Mixtures <strong>of</strong> paper pulp and vermiculite, in<br />

unspecified proportions, are prepared in a commercial<br />

product known as Florialite (Nisshinbo Industries,<br />

Inc., Tokyo). Afreen-Zobayed et al. (1999) found<br />

that growth rates <strong>of</strong> plantlets <strong>of</strong> sweet potato grown<br />

autotrophically on Florialite were greater, in<br />

ascending order, on agar, gellan gum, vermiculite,<br />

Sorbarods and Florialite (best). <strong>The</strong> dry mass <strong>of</strong><br />

leaves and roots were greater by factors <strong>of</strong> 2.9 and<br />

2.8, respectively, on Florialite than on an agar matrix.<br />

<strong>The</strong>se authors observed that roots spread better in<br />

Florialite than in Sorbarods. <strong>The</strong>y attributed this to<br />

the net-like orientation <strong>of</strong> fibers in Florialite that<br />

contrasted with the vertical orientation <strong>of</strong> fibres in<br />

Sorbarods. Ichimura and Oda (1995) found three<br />

substances that stimulated plant growth in extracts <strong>of</strong><br />

paper pulp. Each was characterised by low molecular<br />

weight and high polarity. It is possible that these<br />

substances contribute to the superior growth observed<br />

on substrates containing paper pulp.<br />

6.3.1. Opportunities for improved ventilation and<br />

photoautotrophy<br />

When plantlets are cultured in vessels containing<br />

air at a relative humidity (RH) <strong>of</strong> less than 100%,<br />

transpiration occurs which is an important factor in<br />

reducing hyperhydricity (Gribble, 1999). Relative<br />

humidity in culture vessels can be reduced through<br />

ventilation, but gelled substrates are then unsuitable<br />

because the absorbance <strong>of</strong> water by the roots <strong>of</strong> a<br />

transpiring plant is impeded by the gel’s low hydraulic<br />

conductivity (Fujiwara and Kozai, 1995) and this<br />

increases as the gel dries. Thus a common feature <strong>of</strong><br />

studies using ventilated vessels has been the use <strong>of</strong><br />

liquid medium supported by porous materials.<br />

For example, when plantlets <strong>of</strong> chrysanthemum<br />

were grown in Sorbarods in a culture vessel with air<br />

94% RH, a reduction in the tissue hydration was<br />

indicated by a significantly lower fresh to dry mass<br />

ratio than at 100% RH (Smith et al., 1990b) and<br />

increases in stem length and leaf area. In this<br />

ADRIAN J. & ASSOUMANI M. 1983 Gums and hydrocolloids in<br />

nutrition. pp. 301-333 in Rechcigl M. (ed.) 1983 CRC Handbook<br />

<strong>of</strong> Nutritional Supplements Vol. <strong>II</strong>. Agricultural Use. CRC Press<br />

Inc. Baton Rouge.<br />

AFREEN-ZOBAYED F., ZOBAYED S.M.A., KUBOTA C.,<br />

KOZAI T. & HASEGAWA O. 2000 A combination <strong>of</strong> vermiculite<br />

and paper pulp supporting material for the photoautotrophic<br />

micropropagation <strong>of</strong> sweet potato. <strong>Plant</strong> Sci. 157, 225-231.<br />

Chapter 4 157<br />

REFERENCES<br />

investigation, the RH was reduced to 94% by gaseous<br />

diffusion through a gas-permeable membrane that<br />

covered holes drilled in the lid <strong>of</strong> the culture vessel.<br />

<strong>The</strong> use <strong>of</strong> such ventilated culture vessels can<br />

significantly improve plant growth by reducing<br />

hyperhydration and facilitating the movement <strong>of</strong><br />

solutes to the leaves in the transpiration stream. It<br />

also provides opportunities for photoautotrophic<br />

growth in sugar-free media. When plantlets are<br />

cultured in closed vessels, carbon dioxide<br />

concentrations fall to low levels in the light period, as<br />

Kozai and Sekimoto (1988) demonstrated in cultures<br />

<strong>of</strong> strawberry plants. Photosynthesis requires an<br />

adequate supply <strong>of</strong> carbon dioxide and suitable<br />

lighting. Adequate concentrations <strong>of</strong> carbon dioxide<br />

for photoautotrophy can be maintained in ventilated<br />

culture vessels with (Afreen-Zobayed et al., 1999,<br />

2000) or without (Horan et al., 1995) elevated levels<br />

<strong>of</strong> carbon dioxide in the atmosphere outside the<br />

culture vessel. Adequate lighting can be achieved<br />

under lights delivering a photosynthetic photon flux<br />

<strong>of</strong> 150 μmol m –2 s –1 in a culture room (Afreen-<br />

Zobayed et al., 1999, 2000) or in day-light in a<br />

greenhouse (Horan et al., 1995). <strong>The</strong> environmental<br />

requirements <strong>of</strong> photoautotrophy in vitro have been<br />

reviewed by Jeong et al. (1995) and its advantages<br />

have been described by Kozai et al. (1995).<br />

6.4. IMMOBILISED CELLS<br />

Yields <strong>of</strong> secondary metabolites are usually greater<br />

in differentiated, slow-growing cells than in fast<br />

growing, undifferentiated cells. By immobilising cells<br />

in a suitable matrix, their rate <strong>of</strong> growth can be slowed<br />

and the production <strong>of</strong> secondary products enhanced.<br />

Several ingenious methods <strong>of</strong> immobilisation have<br />

been employed. Examples include immobilization in<br />

spirally wound cotton (Choi et al. (1995), glass fibre<br />

fabric reinforced with a gelling solution <strong>of</strong> hybrid SiO2<br />

precursors (Campostrini et al., 1996), lo<strong>of</strong>a sponge and<br />

polyurethane foam (Liu et al., 1999) and alginate<br />

beads (Serp et al., 2000).<br />

AFREEN-ZOBAYED F., ZOBAYED S.M.A., KUBOTA C.,<br />

KOZAI T. & HASEGAWA O. 1999 Supporting material<br />

affects the growth and development <strong>of</strong> in vitro sweet potato<br />

plantlets cultured photoautotrophically. In Vitro Cell. Dev. –<br />

Pl. 35, 470-474.<br />

AHLOOWALIA B.S. & MARETZKI A. 1983 <strong>Plant</strong> regeneration via<br />

somatic embryogenesis in sugarcane. <strong>Plant</strong> Cell Rep. 2, 21-25.<br />

ALBRECHT C. 1986 Optimization <strong>of</strong> tissue culture media. Comb.<br />

Proc. Int. <strong>Plant</strong> Prop. Soc. 1985, 35, 196-199.


158 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

ALONI R. 1980 Role <strong>of</strong> auxin and sucrose in the differentiation <strong>of</strong><br />

seive and tracheary elements in plant tissue cultures. <strong>Plant</strong>a 150,<br />

255-263.<br />

AMADOR A.M. & STEWART K.A. 1987 Osmotic potential and<br />

pH <strong>of</strong> fluid drilling gels as influenced by moisture loss and<br />

incorporation <strong>of</strong> growth regulators. J. Am. Soc. Hortic. Sci. 112,<br />

26-28.<br />

ANDERS J., LARRABEE P.L. & FAHEY J.W. 1988 Evaluation<br />

<strong>of</strong> gelrite and numerous agar sources for in vitro regeneration <strong>of</strong><br />

sugarcane. HortScience 23, 755.<br />

ANDERSON W.C. 1975 Propagation <strong>of</strong> rhododendrons by tissue<br />

culture. I. Development <strong>of</strong> a culture medium for multiplication <strong>of</strong><br />

shoots. Comb. Proc. Int. <strong>Plant</strong> Prop. Soc. 25, 129-135.<br />

ANDERSON W.C. 1978 <strong>Tissue</strong> culture propagation <strong>of</strong> rhododendrons.<br />

In Vitro 14, 334.<br />

ANDERSON W.C. 1980 <strong>Tissue</strong> culture <strong>of</strong> red raspberries. pp. 27-<br />

34 in Proceedings <strong>of</strong> the Conference on Nursery Production <strong>of</strong><br />

Fruit <strong>Plant</strong>s –Applications and Feasibility. U.S.D.A., A.R.S.,<br />

ARR-NE-11.<br />

ANON (RESEARCH GROUP 301) 1976 A sharp increase <strong>of</strong> the<br />

frequency <strong>of</strong> pollen plant induction in wheat with potato medium.<br />

(Chinese) Acta Genet. Sin. 3, 25-31.<br />

ANON 1980 Ann. Rep. and Accounts, British Petroleum Co.<br />

Ltd. 1979.<br />

ANTHONY P., DAVEY M.R., POWER J.B. & LOWE K.C., 1995<br />

An improved protocol for the culture <strong>of</strong> cassava leaf protoplasts.<br />

<strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 42, 299-302.<br />

ANTHONY P., DAVEY M.R., POWER J.B. & LOWE K.C. 1997<br />

Enhanced mitotic division <strong>of</strong> cultured Passiflora and Petunia<br />

protoplasts by oxygenated perfluorocarbon and haemoglobin.<br />

Biotechnol. Tech. 11, 581-584.<br />

ARDITTI J. & ERNST R. 1984 Physiology <strong>of</strong> germinating orchid<br />

seeds. pp. 177-222 in Arditti J (ed.) 1984 Orchid Biology -<br />

Reviews and Perspectives <strong>II</strong>I. Comstock Publishing, Cornell<br />

Univ. Press, Ithaca, London.<br />

ARDITTI J. 1979 Aspects <strong>of</strong> the physiology <strong>of</strong> orchids. Adv. Bot.<br />

Res. 7, 421-655.<br />

ARNOW P., OLESON J.J. & WILLIAMS J.H. 1953 <strong>The</strong> effect <strong>of</strong><br />

arginine in the nutrition <strong>of</strong> Chlorella vulgaris. Am. J. Bot. 40,<br />

100-104.<br />

ASANO Y., KATSUMOTO H., INOKUMA C., KANEKO S.,<br />

ITO Y. & FUJ<strong>II</strong>E A. 1996 Cytokinin and thiamine requirements<br />

and stimulative effects <strong>of</strong> rib<strong>of</strong>lavin and α-ketoglutaric acid on<br />

embryogenic callus induction from the seeds <strong>of</strong> Zoysia japonica<br />

Steud. J. <strong>Plant</strong> Physiol. 149, 413-417.<br />

ASHER C.J. 1978 Natural and synthetic culture media for<br />

spermatophytes. pp. 575-609 in Recheige M. Jr. (ed.). CRC<br />

Handbook Series in Nutrition and Food. Section G., <strong>Culture</strong><br />

<strong>Media</strong> and Food Supplements. Diets Vol 3.<br />

ATTREE S.M., MOORE D., SAWHNEY V.R. & FOWKE L.C.<br />

1991 Enhanced maturation and desiccation tolerance <strong>of</strong> white<br />

spruce [Picea glauca (Moench.) Voss] somatic embryos: effects<br />

<strong>of</strong> a non-plasmolysing water stress and abscisic acid. Ann. Bot.<br />

68, 519-525.<br />

AYABE S., UDAGAWA A. & FURUYA T. 1988 Stimulation <strong>of</strong><br />

chalcone synthase activity by yeast extract in cultured<br />

Glycorrhiza echinata cells and 5-deoxyflavone formation by<br />

isolated protoplasts. <strong>Plant</strong> Cell Rep. 7, 35-38.<br />

BAGGA S., DAS R. & SOPORY S.K. 1987 Inhibition <strong>of</strong> cell<br />

proliferation and glyoxalase-1 activity by calmodulin inhibitors<br />

and lithium in Brassica oleracea. J. <strong>Plant</strong> Physiol. 129, 149-153.<br />

BALL E. 1953 Hydrolysis <strong>of</strong> sucrose by autoclaving media, a<br />

neglected aspect in the culture <strong>of</strong> plant tissues. B. Torrey Bot.<br />

Club 80, 409-411.<br />

BANIK R.M., KANARI B. & UPADHYAY S.N., 2000.<br />

Exopolysaccharide <strong>of</strong> the gellan family: prospects and potential.<br />

World J. Microb. Biot. 16, 407-414.<br />

BARGHCHI M. 1988 Micropropagation <strong>of</strong> Alnus cordata (Loisel.)<br />

Loisel. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 15, 233-244.<br />

BARG R. & UMIEL N. 1977 Effects <strong>of</strong> sugar concentration on<br />

growth, greenery and shoot formation in callus cultures <strong>of</strong> four<br />

genetic lines <strong>of</strong> tobacco. Z. Pflanzenphysiol. 81, 161-166.<br />

BARLASS M., GRANT W.J.R. & SKENE K.G.M. 1980 Shoot<br />

regeneration in vitro from native Australian fruit-bearing trees –<br />

Quandong and Plum bush. Aust. J. Bot. 28, 405-409.<br />

BARWALE U.B., KERNS H.R. & WIDHOLM J.M. 1986 <strong>Plant</strong><br />

regeneration from callus cultures <strong>of</strong> several soybean genotypes<br />

via embryogenesis and organogenesis. <strong>Plant</strong>a 167, 473-481.<br />

BEHREND J. & MATELES R.I. 1975 Nitrogen metabolism in<br />

plant cell suspension cultures. I. Effect <strong>of</strong> amino acids on growth.<br />

<strong>Plant</strong> Physiol. 56, 584-589.<br />

BERGHOEF J. & BRUINSMA J. 1979 Flower development <strong>of</strong><br />

Begonia franconis Liebm. IV. Adventitious flower bud formation<br />

in excised inflorescence pedicels in vitro. Z. Pflanzenphysiol. 94,<br />

407- 416.<br />

BERGMANN L. 1967 Wachstum gruner Suspensionskulturen von<br />

Nicotiana tabacum Var. ‘Samsun’ mit CO2 als Kohlenst<strong>of</strong>fquelle.<br />

<strong>Plant</strong>a 74, 243-249.<br />

BERUTO D., BERUTO M., CICCARELLI C. & DEBERGH P.<br />

1995 Matric potential evaluations and measurements for gelled<br />

substrates. Physiol. <strong>Plant</strong>. 94, 151-157.<br />

BHATTACHARYA P., DEY S. & BHATTACHARYA B.C. 1994<br />

Use <strong>of</strong> low-cost gelling agents and support matrices for<br />

industrial-scale plant tissue culture. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult.<br />

37, 15-23.<br />

BHOJWANI S.S. & RAZDAN M.K. 1983 <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong>:<br />

<strong>The</strong>ory and Practice. Elsevier Science Publishers B.V.,<br />

Amsterdam.<br />

BIONDI S. & THORPE T.A. 1982 Growth regulator effects,<br />

metabolite changes and respiration during shoot initiation in<br />

cultured cotyledon explants <strong>of</strong> Pinus radiata. Bot. Gaz. 143,<br />

20-25.<br />

BISHNOI U.S., JAIN R.K., GUPTA K.R., CHOWDHURY V.K.<br />

& CHOWDHURY J.B. 2000 High frequency androgenesis in<br />

indica x Basmati rice hybrids using liquid culture media. <strong>Plant</strong><br />

Cell <strong>Tissue</strong> Organ Cult. 61, 153-159.<br />

BISWAS G.C.G. & ZAPATA F.J. 1993 High-frequency plant<br />

regeneration from protoplasts <strong>of</strong> indica rice (Oryza sativa L.)<br />

using maltose. J. <strong>Plant</strong> Physiol. 141, 470-475.<br />

BOLANDI A.R., BRANCHARD M., ALIBERT G., SERIEYS H.<br />

& SARRAFI A. 1999 Cytoplasmic-nuclear interaction and<br />

medium effect for protoplast culture in sunflower. <strong>Plant</strong> Cell<br />

<strong>Tissue</strong> Organ Cult. 57, 189-193.<br />

BOLL W.G. & STREET H.E. 1951 Studies on the growth <strong>of</strong> excised<br />

roots. 1. <strong>The</strong> stimulatory effect <strong>of</strong> molybdenum and copper on the<br />

growth <strong>of</strong> excised tomato roots. New Phytol. 50, 52-75.<br />

BONNER J. & ADDICOTT F. 1937 Cultivation in vitro <strong>of</strong> excised<br />

pea roots. Bot. Gaz. 99, 144-170.<br />

BONNER J. & DEVIRIAN P.S. 1939 Growth factor requirements<br />

<strong>of</strong> four species <strong>of</strong> isolated roots. Am. J. Bot. 26, 661-665.<br />

BONNER J. & DEVIRIAN P.S. 1939 Growth factor requirements<br />

<strong>of</strong> four species <strong>of</strong> isolated roots. Am. J. Bot. 26, 661-665.<br />

BONNER J. 1937 Vitamin B1, a growth factor for higher plants.<br />

Science 85, 183-184.<br />

BONNER J. 1938a Thiamin (Vitamin B1) and the growth <strong>of</strong> roots:<br />

the relation <strong>of</strong> chemical structure to physiological activity. Am. J.<br />

Bot. 25, 543-549.<br />

BONNER J. 1940a Specificity <strong>of</strong> nicotinic acid as a growth factor<br />

for isolated pea roots. <strong>Plant</strong> Physiol. 15, 553-557.


BORGES M., MENESES S., VAZQUEZ J., GARCIA M.,<br />

AGUILERA N., INFANTE Z., RODRIGUEZ A. & FONSECA<br />

M. 2003 In vitro conservation <strong>of</strong> Dioscorea alata L. germplasm<br />

by slow growth. <strong>Plant</strong> Gen. Res. Newsl. 133, 8-12.<br />

BORKIRD C. & SINK K.C. 1983 Medium components for shoot<br />

cultures <strong>of</strong> chlorophyll-deficient mutants <strong>of</strong> Petunia inflata. <strong>Plant</strong><br />

Cell Rep. 2, 1-4.<br />

BÖTTGER M. & LUTHEN H. 1986 Possible linkage between<br />

NADH-oxidation and proton secretion in Zea mays L. roots. J.<br />

Exp. Bot. 37, 666-675.<br />

BOZHKOV P.V. & VON ARNOLD S. 1998 Polyethylene glycol<br />

promotes maturation but inhibits further development <strong>of</strong> Picea<br />

abies somatic embryos. Physiol. <strong>Plant</strong>. 104, 211-224.<br />

BRETZLOFF C.W. Jr. 1954 <strong>The</strong> growth and fruiting <strong>of</strong> Sordaria<br />

fimicola. Am. J. Bot. 41, 58-67.<br />

BRIDSON E.Y. 1978 Diets, culture media and food supplements.<br />

pp. 91-281 in Rechcigl M. Jr. (ed.) CRC Handbook Series in<br />

Nutrition and Food. Section G. Vol.3.<br />

BRINK R.A., COOPER D.C. & AUSHERMAN L.E. 1944 A<br />

hybrid between Hordeum jubatum and Secale cereale. J. Hered.<br />

35, 67-75.<br />

BROWN D.C.W. & THORPE T.A. 1982 Mitochondrial activity<br />

during shoot formation and growth in tobacco callus. Physiol.<br />

<strong>Plant</strong>. 54, 125-130.<br />

BROWN D.C.W. & THORPE T.A. 1980 Changes in water<br />

potential and its components during shoot formation in tobacco<br />

callus. Physiol. <strong>Plant</strong>. 49, 83-87.<br />

BROWN D.C.W., LEUNG D.W.M. & THORPE T.A. 1979<br />

Osmotic requirement for shoot formation in tobacco callus.<br />

Physiol. <strong>Plant</strong>. 46, 36-41.<br />

BROWN C., BROOKS E.J., PEARSON D. & MATHIAS R.J.<br />

1989 Control <strong>of</strong> embryogenesis and organogenesis in immature<br />

wheat embryo callus using increased medium osmolarity and<br />

abscisic acid. J. <strong>Plant</strong> Physiol. 133, 727-733.<br />

BURNET G. & IBRAHIM R.K. 1973 <strong>Tissue</strong> culture <strong>of</strong> Citrus peel<br />

and its potential for flavonoid synthesis. Z. Pflanzenphysiol. 69,<br />

152-162.<br />

BURSTROM H. 1957 Root surface development, sucrose<br />

inversion and free space. Physiol. <strong>Plant</strong>. 10, 741-751.<br />

BUTCHER D.N. & STREET H.E. 1964 Excised root culture. Bot.<br />

Rev. 30, 513-586.<br />

BUTENKO R.G., LIPSKY A.K.H., CHERNYAK N.D. & ARYA<br />

H.C. 1984 Changes in culture medium pH by cell suspension<br />

cultures <strong>of</strong> Dioscorea deltoidea. <strong>Plant</strong> Sci. Lett. 35, 207-212.<br />

CALLEBAUT A. & MOTTE J.-C. 1988 Growth <strong>of</strong> cucumber cells<br />

in media with lactose or milk whey as carbon source. <strong>Plant</strong> Cell<br />

Rep. 7, 162- 165.<br />

CAMPOSTRINI R., CARTURAN G., CANIATO R., PIOVAN<br />

A., FILIPPINI R., INNOCENTI G. & CAPPELLETTI E.M.<br />

1996 Immobilization <strong>of</strong> plant cells in hybrid sol-gel materials. J.<br />

Sol-Gel Sci. Techn. 7, 87-97.<br />

CANHOTO J.M. & CRUZ G.S. 1994 Improvement <strong>of</strong> somatic<br />

embryogenesis in Feijoa sellowiana Berg (Myrtaceae) by<br />

manipulation <strong>of</strong> culture media composition. In Vitro Cell. Dev. –<br />

Pl. 30, 21-25.<br />

CAPLIN S.M. & STEWARD F.C. 1948 Effects <strong>of</strong> coconut milk on<br />

the growth <strong>of</strong> explants from carrot root. Science 108, 655-657.<br />

CARCELLER M., DAVEY M.R., FOWLER M.W. & STREET<br />

H.E. 1971 <strong>The</strong> influence <strong>of</strong> sucrose, 2,4-D and kinetin on the<br />

growth, fine structure and lignin content <strong>of</strong> cultured sycamore<br />

cells. Protoplasma 73, 367-385.<br />

CARIMI F., DE PASQUALE F. & CRESCIMANNO F.G. 1999<br />

Somatic embryogenesis and plant regeneration from pistil thin<br />

cell layers <strong>of</strong> Citrus. <strong>Plant</strong> Cell Rep. 18, 935-940.<br />

CARIMI F., DE PASQUALE F. & PUGLIA A.M. 1998 In vitro<br />

rescue <strong>of</strong> zygotic embryos <strong>of</strong> sour orange, Citrus aurantium L.,<br />

Chapter 4 159<br />

and their detection based on RFLP analysis. <strong>Plant</strong> Breeding 117,<br />

261-266.<br />

CARPITA N., SABULARSE D., MONTEZINOS D. & DELMER<br />

D.P. (1979) Determination <strong>of</strong> the pore size <strong>of</strong> cell walls <strong>of</strong> living<br />

plant cells. Science 205,1144-1147.<br />

CHANDLER M.T., TANDEAU DE MARSAC N. &<br />

KOUCHKOVSKY Y. 1972 Photosynthetic growth <strong>of</strong> tobacco<br />

cells in liquid suspensions. Can. J. Bot. 50, 2265-2270.<br />

CHANDLER S.F., RAGOLSKY E., PUA E.-C. & THORPE T.A.<br />

1987 Some morphogenic effects <strong>of</strong> sodium sulphate on tobacco<br />

callus. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 11, 141-150.<br />

CHAUVIN J.E. & SALESSES G. 1988 Advances in chestnut<br />

micropropagation (Castanea sp.). Acta Hortic. 227, 340-345.<br />

CHAUVIN J.E., MARHADOUR S., COHAT J. & LE NARD M.<br />

1999 Effects <strong>of</strong> gelling agents on in vitro regeneration and<br />

kanamycin efficiency as a selective agent in plant transformation<br />

procedures. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 58, 213-217.<br />

CHAZEN O., HARTUNG W. & NEUMANN P.M. 1995 <strong>The</strong><br />

different effects <strong>of</strong> PEG 6000 and NaCl on leaf development are<br />

associated with differential inhibition <strong>of</strong> root water transport.<br />

<strong>Plant</strong> Cell Environ. 18, 727-735.<br />

CHEE P.P. 1995 Stimulation <strong>of</strong> adventitious rooting <strong>of</strong> Taxus<br />

species by thiamine. <strong>Plant</strong> Cell Rep. 14, 753-757.<br />

CHEE R.P., SCHULTHEIS J.R. & CANTCLIFFE D.J. 1990 <strong>Plant</strong><br />

recovery from sweet potato somatic embryos. HortScience 25,<br />

795-797.<br />

CHENG T.-Y. & VOQUI T.H. 1977 Regeneration <strong>of</strong> Douglas fir<br />

plantlets through tissue culture. Science 198, 306-307.<br />

CHENG T.-Y. 1978 Clonal propagation <strong>of</strong> woody species through<br />

tissue culture techniques. Comb. Proc. Int. <strong>Plant</strong> Prop. Soc. 28,<br />

139-155.<br />

CHEVRE A.-M., GILL S.S., MOURAS A. & SALESSES G. 1983<br />

In vitro multiplication <strong>of</strong> chestnut. J. Hortic. Sci. 58, 23-29.<br />

CHIEN Y.C. & KAO K.N. 1983 Effects <strong>of</strong> osmolality, cytokinin<br />

and organic acids on pollen callus formation in Triticale anthers.<br />

Can. J. Bot. 61, 639-641.<br />

CHIN C. & MILLER D. 1982 Some characteristics <strong>of</strong> the<br />

phosphate uptake by Petunia cells. HortScience 17, 488.<br />

CHIN C.-K., KONG Y. & PEDERSEN H. 1988 <strong>Culture</strong> <strong>of</strong><br />

droplets containing asparagus cells and protoplasts on polypropylene<br />

membrane. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 15, 59-65.<br />

CHOI H.J., TAO B.Y. & OKOS M.R. 1995 Enhancement <strong>of</strong><br />

secondary metabolite production by immobilized Gossypium<br />

arboreum cells. Biotechnol. Progr. 11, 306-311.<br />

CHONG C. & PUA E.-C. 1985 Carbon nutrition <strong>of</strong> Ottawa 3 apple<br />

rootstock during stages <strong>of</strong> in vitro propagation. J. Hortic. Sci. 60,<br />

285-290.<br />

CHONG C. & TAPER C.D. 1972 Malus tissue culture. I. Sorbitol<br />

(D-glucitol) as carbon source for callus initiation and growth.<br />

Can. J. Bot. 50, 1399-1404.<br />

CHONG C. & TAPER C.D. 1974a Malus tissue cultures. <strong>II</strong>. Sorbitol<br />

metabolism and carbon nutrition. Can. J. Bot. 52, 2361-2364.<br />

CHONG C. & TAPER C.D. 1974b Influence <strong>of</strong> light intensity on<br />

sorbitol metabolism, growth and chlorophyll content <strong>of</strong> Malus<br />

tissue cultures. Ann. Bot. 38, 359-362.<br />

CHU C.-C., WANG C.-C., SUN C.-S., HSU C., YIN K.-C., CHU<br />

C.-Y. & BI F.-Y. 1975 Establishment <strong>of</strong> an efficient medium for<br />

anther culture <strong>of</strong> rice, through comparative experiments on the<br />

nitrogen sources. Sci.Sinica 18, 659-668.<br />

CHUA S.E. 1966 Studies on tissue culture <strong>of</strong> Hevea brasiliensis. I.<br />

Role <strong>of</strong> osmotic concentration, carbohydrate and pH value in<br />

induction <strong>of</strong> callus growth in plumule tissue from Hevea<br />

seedling. J. Rubber Res. Inst. Malaya 19, 272-276.<br />

CHUANG C.-C., OUYANG T.-W., CHIA H., CHOU S.-M. &<br />

CHING C.-K. 1978 A set <strong>of</strong> potato media for wheat anther


160 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

culture. pp. 51-56 in <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong>. Proceedings <strong>of</strong> the<br />

Peking Symposium 1978. Pitman, Boston, London, Melbourne.<br />

CHUONG P.V. & BEVERSDORF W.D. 1985 High frequency<br />

embryogenesis through isolated microspore culture in Brassica<br />

napus L and B. carinata Braun. <strong>Plant</strong> Sci. 39, 219-226.<br />

CLELAND R. 1977 <strong>The</strong> control <strong>of</strong> cell enlargement. pp. 101-115<br />

in Jennings D.H. (ed.) S.E.B. Symp. 31. University Press,<br />

Cambridge.<br />

COFFIN R., TAPER C.D. & CHONG C. 1976 Sorbitol and<br />

sucrose as carbon source for callus culture <strong>of</strong> some species <strong>of</strong> the<br />

Rosaceae. Can. J. Bot. 54, 547-551.<br />

CONNER A.J. & FALLOON P.G. 1993 Osmotic versus<br />

nutritional effects when rooting in vitro asparagus minicrowns on<br />

high sucrose. <strong>Plant</strong> Sci. 89, 101-106.<br />

CONNER A.J. & MEREDITH C.P. 1984 An improved<br />

polyurethane support system for monitoring growth in plant cell<br />

cultures. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 3, 59-68.<br />

CONRAD P.L., CONRAD J.M., DURBIN R.D. & HELGESON J.P.<br />

1981 Effects <strong>of</strong> some organic buffers on division <strong>of</strong> protoplastderived<br />

cells and plant regeneration. <strong>Plant</strong> Physiol. 67, Suppl., 116.<br />

COOPER A. 1979 <strong>The</strong> ABC <strong>of</strong> NFT. Grower Books, London.<br />

ISBN 0 901 361224.<br />

COPPING L.G. & STREET H.E. 1972 Properties <strong>of</strong> the<br />

invertases <strong>of</strong> cultured sycamore cells and changes in their activity<br />

during culture growth. Physiol. <strong>Plant</strong>. 26, 346-354.<br />

CULAFIC L., BUDIMIR S., VUJICIC R. & NESKOVIC M. 1987<br />

Induction <strong>of</strong> somatic embryogenesis and embryo development in<br />

Rumex acetosella L. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 11, 133-139.<br />

DABIN P. & BEGUIN F. 1987 Somatic embryogenesis in<br />

Fuchsia. Acta Hortic. 212, 725-726.<br />

DAIGNY G., PAUL H. & SANGWAN-NORRELL B.S. 1996<br />

Factors influencing secondaru somatic embryogenesis in Malus x<br />

domestica Borkh. (cv ‘Gloster 69’). <strong>Plant</strong> Cell Rep. 16, 153-157.<br />

DALTON C.C., IQBAL K. & TURNER D.A. 1983 Iron phosphate<br />

precipitation in Murashige and Skoog media. Physiol. <strong>Plant</strong>. 57,<br />

472-476.<br />

DAMIANO C., CURIR P. & COSMI T. 1987 Short note on the<br />

effects <strong>of</strong> sugar on the growth <strong>of</strong> Eucalyptus gunnii in vitro. Acta<br />

Hortic. 212, 553-556.<br />

DANTU P.K. & BHOJWANI S.S. 1995 In vitro corm formation<br />

and field evaluation <strong>of</strong> corm-derived plants <strong>of</strong> Gladiolus. Sci.<br />

Hortic. 61, 115-129.<br />

DAS R., BAGGA S. & SOPORY S.K. 1987 Involvement <strong>of</strong><br />

phosphoinositides, calmodulin and glyoxidase-1 in cell<br />

proliferation in callus cultures <strong>of</strong> Amaranthus paniculatus. <strong>Plant</strong><br />

Sci. 53, 45-51.<br />

DAS T., MITRA G.C. & CHATTERJEE A. 1995 Micropropagation<br />

<strong>of</strong> Citrus sinensis var. mosambi: an important scion.<br />

Phytomorph. 45, 57-64.<br />

DAVIS M.J., BAKER R. & HANAN J.J. 1977 Clonal multiplication<br />

<strong>of</strong> carnation by micropropagation. J. Am. Soc. Hortic.<br />

Sci. 102, 48-53.<br />

DAY D. 1942 Thiamin content <strong>of</strong> agar. B. Torrey Bot. Club 69,<br />

11-20.<br />

DE BRUYN M.H. AND FERREIRA D.I. 1992 In vitro corm<br />

production <strong>of</strong> Gladiolus dalenii and G. tristis. <strong>Plant</strong> Cell <strong>Tissue</strong><br />

Organ Cult. 31, 123-128.<br />

DE CAPITE L. 1948 pH <strong>of</strong> nutritive solutions and its influence on<br />

the development <strong>of</strong> plants. Ann. Fac. Agrar. Univ. Perugia 5,<br />

135-145.<br />

DE CAPITE L. 1952a Combined action <strong>of</strong> p-aminobenzoic acid and<br />

indoleacetic acid on the vascular parenchyma <strong>of</strong> Jerusalem<br />

artichoke cultured in vitro. Compt. Rend. Soc. Biol. 146, 863-865.<br />

DE CAPITE L. 1952b Action <strong>of</strong> p-aminophenylsulphonamide and<br />

p-amino-benzoic acid on the vascular parenchyma <strong>of</strong> Jerusalem<br />

artichoke and <strong>of</strong> crown-gall tissue <strong>of</strong> salsify. Compt. Rend. Acad.<br />

Sci. Paris 234, 2478-2480.<br />

DE FOSSARD R.A., MYINT A. & LEE E.C.M. 1974 A broad<br />

spectrum tissue culture experiment with tobacco (Nicotiana<br />

tabacum) pith tissue callus. Physiol. <strong>Plant</strong>. 31, 125-130.<br />

DE JONG A.W. & BRUINSMA J. 1974 Pistil development in<br />

Cleome flowers <strong>II</strong>I. Effects <strong>of</strong> growth-regulating substances on<br />

flower buds <strong>of</strong> Cleome iberidella Welv. ex Oliv. grown in vitro.<br />

Z. Pflanzenphysiol. 73, 142-151.<br />

DE JONG A.W., SMIT A.L. & BRUINSMA J. 1974 Pistil<br />

development in Cleome flowers <strong>II</strong>. Effects <strong>of</strong> nutrients on flower<br />

buds <strong>of</strong> Cleome iberidella Welv. ex Oliv. grown in vitro. Z.<br />

Pflanzenphysiol. 72, 227-236.<br />

DE KLERK G.J., HANECAKOVA J. & JASIK J. 2007 Effect<br />

<strong>of</strong> medium-pH on adventitious root formation from thin stem<br />

disks cut from apple microshoots. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ<br />

Cult. (in press)<br />

DE KLERK-KIEBERT Y.M. & VAN DER PLAS L.H.W. 1985<br />

Relationship <strong>of</strong> respiratory pathways in soybean cell suspensions<br />

to growth <strong>of</strong> the cells at various glucose concentrations. <strong>Plant</strong><br />

Cell <strong>Tissue</strong> Organ Cult. 4, 225-233.<br />

DE KRUIJFF E. 1906 Composition <strong>of</strong> coconut water and presence<br />

<strong>of</strong> diastase in coconuts. B. Dep. Agric. Indes Neerland. 4, 1-8.<br />

DE PASQUALE F., CARIMI F. & CRESCIMANNO F.G. 1994<br />

Somatic embryogenesis from styles <strong>of</strong> different cultivars <strong>of</strong><br />

Citrus limon (L.) Burm. Aust. J. Bot, 42, 587-594.<br />

DE PINTO M.C., FRANCIS D. & DE GARA D. 1999 <strong>The</strong> redox<br />

state <strong>of</strong> the ascorbate-dehydroascrobate pair as a specific sensor<br />

<strong>of</strong> cell division in tobacco BY-2 cells. Protoplasma 209, 90-97.<br />

DEBERGH P.C. 1983 Effects <strong>of</strong> agar brand and concentration on<br />

the tissue culture medium. Physiol. <strong>Plant</strong>. 59, 270-276.<br />

DEBERGH P., AITKEN-CHRISTIE J., COHEN D., GROUT B.,<br />

VON ARNOLD S., ZIMMERMAN R. & ZIV M. 1992 Reconsideration<br />

<strong>of</strong> the term ‘vitrification’ as used in micropropagation.<br />

<strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 30, 135-140.<br />

DEBERGH P., HARBAOUI Y. & LEMEUR R. 1981. Mass<br />

propagation <strong>of</strong> globe artichoke (Cyanara scolymus): evaluation<br />

<strong>of</strong> different hypotheses to overcome vitrification with special<br />

reference to water potential. Physiol. <strong>Plant</strong>. 53, 181-187<br />

DELBARRE A., MULLER P., IMHOFF V. & GUERN J. 1996<br />

Comparison <strong>of</strong> mechanisms controlling uptake and accumulation<br />

<strong>of</strong> 2,4-dichlorophenoxy acetic acid, naphthalene-1-acetic acid,<br />

and indole-3-acetic acid in suspension-cultured tobacco cells.<br />

<strong>Plant</strong>a 198, 532-541.<br />

DELFEL N.E. & SMITH L.J. 1980 <strong>The</strong> importance <strong>of</strong> culture conditions<br />

and medium component interactions on the growth <strong>of</strong><br />

Cephalotaxus harringtonia tissue cultures. <strong>Plant</strong>a Medica 40,<br />

237-244.<br />

DIGBY J. & SKOOG F. 1966 Cytokinin activation <strong>of</strong> thiamine<br />

biosynthesis in tobacco callus cultures. <strong>Plant</strong> Physiol. 41, 647-652.<br />

DIJKEMA C., BOOIJ H., DE JAGER P.A., DE VRIES S.C.,<br />

SCHAAFSMA T.J. & VAN KAMMEN A. 1990 Aerobic hexose<br />

metabolism in embryogenic cell lines <strong>of</strong> Daucus carota studied<br />

by 13 C NMR using 13 C-enriched substrates. pp. 343-348 in<br />

Nijkamp et al. (eds.) 1990 (q.v.).<br />

DIX L. & VAN STADEN J. 1982 Auxin and gibberellin-like<br />

substances in coconut milk and malt extract. <strong>Plant</strong> Cell <strong>Tissue</strong><br />

Organ Cult. 1, 239-245.<br />

DOLEY D. & LEYTON L. 1970 Effects <strong>of</strong> growth regulating<br />

substances and water potential on the development <strong>of</strong> wound<br />

callus in Fraxinus. New Phytol. 69, 87-102.<br />

DORAN P.M. 1993 Design <strong>of</strong> bioreactors for plant cells and<br />

organs. pp 116-169 in Fiechter A. (ed.) Bioprocess design and<br />

control. Vol. 48 Springer-Verlag, Berlin. pp 116-169.


DOUGALL D.K. 1980 Nutrition and Metabolism. <strong>Plant</strong> <strong>Tissue</strong><br />

<strong>Culture</strong> as a Source <strong>of</strong> Biochemicals. C.R.C. Press, Boca Raton,<br />

Florida.<br />

DOUGALL D.K., WEYRAUCH K.W. & ALTON W. 1979 <strong>The</strong><br />

effects <strong>of</strong> organic acids on growth and anthocyanin production by<br />

wild carrot cells growing on ammonia as a sole nitrogen source.<br />

In Vitro 15, 189-190 (Abst. 103).<br />

DOUGLAS T.J., VILLALOBOS V.M., THOMPSON M.R. &<br />

THORPE T.A. 1982 Lipid and pigment changes during shootinititaion<br />

in cultured explants <strong>of</strong> Pinus radiata. Physiol. <strong>Plant</strong>.<br />

55, 470-477.<br />

DOVZHENKO A., BERGEN U. & KOOP H.U. 1998 Thinalginate-layer<br />

technique for protoplast culture <strong>of</strong> tobacco leaf<br />

protoplasts: shoot formation in less than two weeks. Protoplasma<br />

204, 114-118.<br />

DREW R.A. & SMITH N.G. 1986 Growth <strong>of</strong> apical and lateral<br />

buds <strong>of</strong> pawpaw (Carica papaya L.) as affected by nutritional<br />

and hormonal factors. J. Hortic. Sci. 61, 535-543.<br />

DREW R.A., MCCOMB J.A. & CONSIDINE J.A. 1993<br />

Rhizogenesis and root growth <strong>of</strong> Carica papaya L. in vitro in<br />

relation to auxin sensitive phases and use <strong>of</strong> rib<strong>of</strong>lavin. <strong>Plant</strong><br />

Cell <strong>Tissue</strong> Organ Cult. 33, 1-7.<br />

DRUART PH. 1988 Regulation <strong>of</strong> axillary branching in<br />

micropropagation <strong>of</strong> woody fruit species. Acta Hortic. 227,<br />

369-380.<br />

DUHAMET L. & MENTZER C. 1955 Essais d’isolement des<br />

substances excito-formatrices du lait de coco. Compt. Rend.<br />

Acad. Sci. Paris 241, 86-88.<br />

DUNSTAN D.I. 1982 Transplantation and post-transplantation <strong>of</strong><br />

micropropagated tree-fruit rootstocks. Comb. Proc. Int. <strong>Plant</strong><br />

Prop. Soc., 1981 31, 39-44.<br />

DUNSTAN W.R. 1906 Report on a sample <strong>of</strong> coconut “water”<br />

from Ceylon. Trop. Agric. (Ceylon) 26, 377-378.<br />

DUNWELL J.M. 1981 Influence <strong>of</strong> genotype and environment on<br />

growth <strong>of</strong> barley Hordeum vulgare embryos in vitro. Ann. Bot.<br />

48, 535-542.<br />

EMONS A.M.C., SAMALLO-DROPPERS A. & VAN DER<br />

TOORN C. 1993 <strong>The</strong> influence <strong>of</strong> sucrose, mannitol, L-proline,<br />

abscisic acid and gibberellic acid on the maturation <strong>of</strong> somatic<br />

embryos <strong>of</strong> Zea mays L. from suspension cultures. J. <strong>Plant</strong><br />

Physiol. 142, 597-604.<br />

EARLE E.D. & TORREY J.G. 1965 Colony formation by isolated<br />

Convolvulus cells plated on defined media. <strong>Plant</strong> Physiol. 40,<br />

520-528.<br />

EDELMAN J. & HANSON A.D. 1972 Sucrose suppression <strong>of</strong><br />

chlorophyll synthesis in carrot-tissue cultures. J. Exp. Bot. 23,<br />

469-478.<br />

EDWARDS K.J. & GOLDSMITH M.H.M. 1980 pH-dependent<br />

accumulation <strong>of</strong> indoleacetic acid by corn coleoptile sections.<br />

<strong>Plant</strong>a 147, 457-466.<br />

EINSET J.W. 1978 Citrus tissue culture - stimulation <strong>of</strong> fruit<br />

explant cultures with orange juice. <strong>Plant</strong> Physiol. 62, 885-888.<br />

EL HINNAWIY E. 1974 Effect <strong>of</strong> some growth regulating<br />

substances and carbohydrates on chlorophyll production in<br />

Melilotus alba (Desr.) callus tissue cultures. Z. Pflanzenphysiol.<br />

74, 95-105.<br />

ERNER Y. & REUVENI O. 1981 Promotion <strong>of</strong> citrus tissue<br />

culture by citric acid. <strong>Plant</strong> Physiol. 67, Suppl., 27.<br />

ERNER Y. & REUVENI O. 1981 Promotion <strong>of</strong> citrus tissue<br />

culture by citric acid. <strong>Plant</strong> Physiol. 67, Suppl., 27.<br />

ERNST R. 1967 Effects <strong>of</strong> carbohydrate selection on the growth<br />

rate <strong>of</strong> freshly germinated Phalaenopsis and Dendrobium seed.<br />

Am. Orchid Soc. Bull. 36, 1068-1073.<br />

ERNST R. 1974 <strong>The</strong> use <strong>of</strong> activated charcoal in asymbiotic<br />

seedling culture <strong>of</strong> Paphiopedilum. Am. Orchid Soc. Bull. 43,<br />

35-38.<br />

Chapter 4 161<br />

ERNST R., ARDITTI J. & HEALEY P.L. 1971 Carbohydrate<br />

physiology <strong>of</strong> orchid seedlings. <strong>II</strong>. Hydrolysis and effects <strong>of</strong><br />

oligosaccharides. Am. J. Bot. 58, 827-835.<br />

EVANS D.A., SHARP W.R. & PADDOCK E.F. 1976 Variation in<br />

callus proliferation and root morphogenesis in leaf tissue cultures<br />

<strong>of</strong> Glycine max strain T 219. Phytomorph. 26, 379-384.<br />

FELLE H.H. 1998. <strong>The</strong> apoplastic pH <strong>of</strong> the Zea mays root cortex<br />

as measured with pH-sensitive microelectrodes: aspects <strong>of</strong><br />

regulation. J. Exp. Bot. 49, 987-995.<br />

FELLE H.H. 2001 pH: Signal and messenger in plant cells. <strong>Plant</strong><br />

Biol. 3, 577-591.<br />

FELLE H.H. & HANSTEIN S. 2002 <strong>The</strong> apoplastic pH <strong>of</strong> the<br />

substomatal cavity <strong>of</strong> Vicia faba leaves and its regulation<br />

responding to different stress factors. J. Exp. Bot. 53, 73-82.<br />

FERGUSON J.D., STREET H.E. & DAVID S.B. 1958 <strong>The</strong> carbohydrate<br />

nutrition <strong>of</strong> tomato roots. V. <strong>The</strong> promotion and<br />

inhibition <strong>of</strong> excised root growth by various sugars and sugar<br />

alcohols. Ann. Bot. 22, 513-524.<br />

FINDENEGG G.R., VAN BEUSICHEM M.L. & KELTJENS<br />

W.G. 1986 Proton balance <strong>of</strong> plants: physiological, agronomical<br />

and economical implications. Neth. J. Agric. Sci. 34, 371-379.<br />

FINNIE S.J., POWELL W. & DYER A.F. 1989 <strong>The</strong> effect <strong>of</strong><br />

carbohydrate composition and concentration on anther culture<br />

response in barley. <strong>Plant</strong> Breeding 103, 110-118.<br />

FONNESBECH M. 1972 Organic nutrients in the media for the<br />

propagation <strong>of</strong> Cymbidium in vitro. Physiol. <strong>Plant</strong>. 27, 360-364.<br />

FOX J.E. & MILLER C. 1959 Factors in corn steep water<br />

promoting growth <strong>of</strong> plant tissues. <strong>Plant</strong> Physiol. 34, 577-579.<br />

FUGGI A., RUGANO V.M., VONA V. & RIGANO C. 1981<br />

Nitrate and ammonium assimilation in algae cell-suspensions and<br />

related pH variations in the external medium, monitored by<br />

electrodes. <strong>Plant</strong> Sci. Lett. 23, 129-138.<br />

FUJIWARA A. (ed.) 1982 <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> 1982. Proc. 5th.<br />

Int. Cong. <strong>Plant</strong> Tiss. Cell Cult., Japan. Assoc. <strong>Plant</strong> <strong>Tissue</strong><br />

<strong>Culture</strong>, Japan.<br />

FUJIWARA K. & KOZAI T. 1995 Physical microenvironment and<br />

its effects. pp. 319-369 in Aitken-Christie J., Kozai T. and Smith<br />

M.L. (eds.) Automation and Environmental Control in <strong>Plant</strong><br />

<strong>Tissue</strong> <strong>Culture</strong>. Kluwer Academic Publishers, <strong>The</strong> Netherlands.<br />

FUKAMI T. & HILDEBRANDT A.C. 1967 Growth and chlorophyll<br />

formation in edible green plant callus tissues in vitro on<br />

media with limited sugar supplements. Bot. Mag. Tokyo 80,<br />

199-212.<br />

FURGUSON J.D. 1967 <strong>The</strong> nutrition <strong>of</strong> excised wheat roots.<br />

Physiol. <strong>Plant</strong>. 20, 276-284.<br />

GALIBA G. & ERDEI L. 1986 Dependence <strong>of</strong> wheat callus<br />

growth, differentiation and mineral content on carbohydrate<br />

supply. <strong>Plant</strong> Sci. 45, 65-70.<br />

GALIBA G. & YAMADA Y. 1988 A novel method for increasing<br />

the frequency <strong>of</strong> somatic embryogenesis in wheat tissue culture<br />

by NaCl and KCl supplementation. <strong>Plant</strong> Cell Rep. 7, 55-58.<br />

GAMBORG O.L. & SHYLUK J.P. 1970 <strong>The</strong> culture <strong>of</strong> plant cells<br />

with ammonium salts as a sole nitrogen source. <strong>Plant</strong> Physiol. 45,<br />

598-600.<br />

GAMBORG O.L., CONSTABEL F. & SHYLUK J.P. 1974<br />

Organogenesis in callus from shoot apices <strong>of</strong> Pisum sativum.<br />

Physiol. <strong>Plant</strong>. 30, 125-128.<br />

GAMBORG O.L., MILLER R.A. & OJIMA K. 1968 Nutrient<br />

requirements <strong>of</strong> suspension cultures <strong>of</strong> soybean root cells. Exp.<br />

Cell Res. 50, 151-158.<br />

GAMBORG O.L., MILLER R.A. & OJIMA K. 1968 Nutrient<br />

requirements <strong>of</strong> suspension cultures <strong>of</strong> soybean root cells. Exp.<br />

Cell Res. 50, 151-158.<br />

GARIN E., BERNIER-CARDOU M. , ISABEL N.,<br />

KLIMASZEWSKA K. & PLOURDE A. 2000 Effect <strong>of</strong> sugars,<br />

amino acids and culture technique on maturation <strong>of</strong> somatic


162 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

embryos <strong>of</strong> Pinus strobus on medium with two gellan gum<br />

concentrations. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 62, 27-37.<br />

GASPAR T., KEVERS C., DEBERGH P., MAENE L., PAQUES<br />

M. & BOXUS P. 1987 Vitrification: morphological, physiological<br />

and ecological aspects. in Bonga J.M. and Durzan D.J.<br />

(eds.) Cell and <strong>Tissue</strong> <strong>Culture</strong> in Forestry Vol. 1 Kluwer<br />

Academic Press Publ., Dordrecht.<br />

GAUTHERET R.J. 1942 Manuel Technique de <strong>Culture</strong> des <strong>Tissue</strong><br />

Végétaux. Masson et Cie, Paris.<br />

GAUTHERET R.J. 1945 Une voie nouvelle en biologie végétale:<br />

la culture des tissus. Gallimard, Paris.<br />

GAUTHERET R.J. 1948 Sur la culture indéfinie des tissus de Salix<br />

caprea. Compt. Rend. Soc. Biol. 142, 807.<br />

GAWEL N.J., RAO A.P. & ROBACKER C.D. 1986 Somatic<br />

embryogenesis from leaf and petiole callus cultures <strong>of</strong><br />

Gossypium hirsutum L. <strong>Plant</strong> Cell Rep. 5, 457-459.<br />

GAWEL N.J. & ROBACKER C.D. 1990 Somatic embryogenesis<br />

in two Gossypium hirsutum genotypes on semi-solid versus<br />

liquid proliferation media. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult.23,<br />

201-204.<br />

GEORGE E.F., PUTTOCK D.J.M. & GEORGE H.J. 1987 <strong>Plant</strong><br />

<strong>Culture</strong> <strong>Media</strong> Vol.1. Exegetics Ltd., Westbury, England.<br />

GERRITS M. & DE KLERK G.J. 1992 Dry-matter partitioning<br />

between bulbs and leaves in plantlets <strong>of</strong> Lilium speciosum<br />

regenerated in vitro. Acta Bot. Neerl. 41, 461-468.<br />

GLASSTONE S. 1946 Textbook <strong>of</strong> Physical Chemistry. Second<br />

Edition. MacMillan, London.<br />

GOLDSCHMIDT E.E. 1976 Edogenous growth substances <strong>of</strong><br />

citrus tissues. HortScience 11, 95-99.<br />

GOOD N.E., WINGET G.D., WINTER W., CONNOLLY T.N.,<br />

IZAWA S. & SINGH R.M. 1966 Hydrogen ion buffers for<br />

biological research. Biochemistry 5, 467-477.<br />

GOODWIN P.B. 1966 An improved medium for the rapid growth<br />

<strong>of</strong> isolated potato buds. J. Exp. Bot. 17, 590-595.<br />

GOPAL J., CHAMAIL A. & SARKAR D. 2002 Slow-growth in<br />

vitro conservation <strong>of</strong> potato germplasm at normal propagation<br />

temperature. Pot. Res. 45, 203-213.<br />

GRANATEK C.H. & COCKERLINE A.W. 1978 Callus formation<br />

versus differentiation <strong>of</strong> cultured barley embryos: hormonal<br />

osmotic interactions. In Vitro 14, 212-217.<br />

GRAY D.J., CONGER B.V. & SONGSTAD D.D. 1987<br />

Desiccated quiescent somatic embryos <strong>of</strong> orchard grass for use as<br />

synthetic seeds. In Vitro Cell. Dev. 23, 29-32.<br />

GRIBBLE K. 1999 <strong>The</strong> influence <strong>of</strong> relative humidity on<br />

vitrification, growth and morphology <strong>of</strong> Gypsophila paniculata<br />

L. <strong>Plant</strong> Growth Regul. 27, 179-188.<br />

GROOTAARTS H., SCHEL J.H.N. & PIERIK R.L.M. 1981 <strong>The</strong><br />

origin <strong>of</strong> bulblets formed on excised twin scales <strong>of</strong> Nerine<br />

bowdenii W. Watts. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 1, 39-46.<br />

GROSS K.C., PHARR D.M. & LOCY R.D. 1981 Growth <strong>of</strong> callus<br />

initiated from cucumber hypocotyls on galactose and galactosecontaining<br />

oligosaccharides. <strong>Plant</strong> Sci. Lett. 20, 333-341.<br />

GUHA S. & JOHRI B.M. 1966 In vitro development <strong>of</strong> overy and<br />

ovule <strong>of</strong> Allium cepa L. Phytomorph. 16, 353-364.<br />

GUHA S. & MAHESHWARI S.C. 1964 In vitro production <strong>of</strong><br />

embryos from anthers <strong>of</strong> Datura. Nature 204, 497.<br />

GUHA S. & MAHESHWARI S.C. 1967 Development <strong>of</strong><br />

embryoids from pollen grains <strong>of</strong> Datura in vitro. Phytomorph.<br />

17, 454-461.<br />

GUNNING B.E.S. & STEER M.W. 1975 Ultrastructure and the<br />

Biology <strong>of</strong> <strong>Plant</strong> Cells. Arnold, London.<br />

GUPTA P.K. & DURZAN D.J. 1985 Shoot multiplication from<br />

mature trees <strong>of</strong> Douglas-fir (Pseudotsuga menziesii) and sugar<br />

pine (Pinus lambertiana). <strong>Plant</strong> Cell Rep. 4, 177-179.<br />

GUPTA P.K., DANDEKAR A.M. & DURZAN D.J. 1988 Somatic<br />

proembryo formation and transient expression <strong>of</strong> a luciferase<br />

gene in Douglas fir and loblolly pine protoplasts. <strong>Plant</strong> Sci. 58,<br />

85-92.<br />

GUTMAN T.S. & SHIRYAEVA G.A. 1980 New method for<br />

growing cultures <strong>of</strong> isolated tissues <strong>of</strong> higher plants. Rastit.<br />

Resur. 16, 601-606.<br />

HACKETT D.P. 1952 <strong>The</strong> osmotic change during auxin-induced<br />

water uptake by potato tissue. <strong>Plant</strong> Physiol. 27, 279-284.<br />

HAKKAART F.A. & VERSLUIJS J.M.A. 1983 Some factors<br />

affecting glassiness in carnation meristem tip cultures. Neth. J.<br />

<strong>Plant</strong> Path. 89, 47-53.<br />

HALPERIN W. & WETHERELL D.F. 1964 Adventive embryony<br />

in tissue cultures <strong>of</strong> the wild carrot, Daucus carota. Am. J. Bot.<br />

51, 274-283.<br />

HAMAOKA Y., FUJITA Y. & IWAI S. 1991 Effects <strong>of</strong><br />

temperature on the mode <strong>of</strong> pollen development in anther culture<br />

<strong>of</strong> Brassica campestris. Physiol. <strong>Plant</strong>. 82, 67-72.<br />

HAMMERSLEY-STRAW D.R.H. & THORPE T.A. 1988 Use <strong>of</strong><br />

osmotic inhibition in studies <strong>of</strong> shoot formation in tobacco callus<br />

cultures. Bot. Gaz. 149, 303-310.<br />

HARBAGE J.F., STIMART D.P. & AUER C. 1998 pH affects<br />

1H-indole-3-butyric acid uptake but not metabolism during the<br />

initiation phase <strong>of</strong> adventitious root induction in apple<br />

microcuttings. J. Am. Soc. Hortic. Sci. 123, 6-10.<br />

HARDING K., BENSON E.E. & CLACHER K. 1997 <strong>Plant</strong><br />

conservation biotechnology: an overview. Agro Food Ind. Hi<br />

Tech. 8, 24-29.<br />

HARRAN S. & DICKINSON D.B. 1978 Metabolism <strong>of</strong> myoinositol<br />

and growth in various sugars <strong>of</strong> suspension cultured<br />

cells. <strong>Plant</strong>a 141, 77-82.<br />

HARRINGTON H.M. & HENKE R.R. 1981 Amino acid transport<br />

into cultured tobacco cells. I. Lysine transport. <strong>Plant</strong> Physiol. 67,<br />

373-378.<br />

HARRIS R.E. & STEVENSON J.H. 1979 Virus elimination and<br />

rapid propagation <strong>of</strong> grapes in vitro. Comb. Proc. Int. <strong>Plant</strong> Prop.<br />

Soc. 29, 95-108.<br />

HARVAIS G. 1982 An improved medium for growing the orchid<br />

Cypripedium reginae axenically. Can. J. Bot. 60, 2547-2555.<br />

HELGESON J.P., UPPER C.D. & HABERLACH G.T. 1972<br />

Medium and tissue sugar concentrations during cytokinincontrolled<br />

growth <strong>of</strong> tobacco callus tissues. pp. 484-492 in Carr<br />

D.J. (ed.) <strong>Plant</strong> Growth Substances 1970. Springer Verlag,<br />

Berlin, Heidelberg, New York.<br />

HELLER R. & GAUTHERET R.J. 1949 Sur l’emploi de papier<br />

filtre sans cendres commes support pour les cultures de tissues<br />

végétaux. Compt. Rend. Seance Soc. Biol., Paris 143, 335-337.<br />

HENDERSON W.E. & KINNERSLEY A.M. 1988 Corn starch as<br />

an alternative gelling agent for plant tissue culture. <strong>Plant</strong> Cell<br />

<strong>Tissue</strong> Organ Cult. 15, 17-22.<br />

HESS D., LEIPOLDT G. & ILLG R.D. 1979 Investigations on the<br />

lactose induction <strong>of</strong> β-galactosides activity in callus tissue<br />

cultures <strong>of</strong> Nemesia strumosa and Petunia hybrida. Z.<br />

Pflanzenphysiol. 94, 45-53.<br />

HEW C.S., TING S.K. & CHIA T.F. 1988 Substrate utilization by<br />

Dendrobium tissues. Bot. Gaz. 149, 153-157.<br />

HILDEBRANDT A.C., RIKER A.J. & DUGGAR B.M. 1946 <strong>The</strong><br />

influence <strong>of</strong> the composition <strong>of</strong> the medium on growth in vitro <strong>of</strong><br />

excised tobacco and sunflower tissue cultures. Am. J. Bot. 33,<br />

591-597.<br />

HILDEBRANDT A.C., WILMAR J.C., JOHNS H. & RIKER A.J.<br />

1963 Growth <strong>of</strong> edible chlorophyllous plant tissues in vitro. Am.<br />

J. Bot. 50, 248-254.<br />

HISAJIMA S. & THORPE T.A. 1981 Lactose-adapted cultured<br />

cells <strong>of</strong> Japanese morning-glory. Acta Physiol. <strong>Plant</strong>. 3, 187-191.<br />

HISAJIMA S. & THORPE T.A. 1985 Lactose metabolism in<br />

lactose-adapted cells <strong>of</strong> Japanese morning-glory. J. <strong>Plant</strong><br />

Physiol. 118, 145-151.


HISAJIMA S., ARAI Y. & ITO T. 1978 Changes in sugar<br />

contents and some enzyme activities during growth <strong>of</strong> morningglory<br />

callus (In Japanese). J. Jpn. Soc. Starch Sci. 25, 223-228.<br />

HO W.-J. & VASIL I.K. 1983 Somatic embryogenesis in<br />

sugarcane (Saccharum <strong>of</strong>ficinarum L.): Growth and plant<br />

regeneration from embryogenic cell suspension cultures. Ann.<br />

Bot. 51, 719-726.<br />

HOMÈS J. & VANSVEREN-VAN ESPEN N. 1973 Effets du<br />

saccharose et de la lumière sur le développement et la<br />

morphologie de protocormes d’Orchides cultivés in vitro. Bull.<br />

Soc. Roy. Bot. Belg. 106, 89-106.<br />

HOOKER T.S. & THORPE T.A. 1997 Effects <strong>of</strong> water deficit<br />

stress on the developmental growth <strong>of</strong> excised tomato roots<br />

cultured in vitro. In Vitro Cell. Dev.-Pl. 33, 245-251.<br />

HORAN I., WALKER S., ROBERTS A.V., MOTTLEY J. &<br />

SIMPKINS I. 1995 Micropropagation <strong>of</strong> roses: the benefits <strong>of</strong><br />

pruned mother-plantlets at Stage <strong>II</strong> and a greenhouse<br />

environment at stage <strong>II</strong>I. J. Hortic. Sci., 70, 799-806.<br />

HOWARD B.H. & MARKS T.R. 1987 <strong>The</strong> in vitro - in vivo<br />

interface. pp. 101-111 in Jackson M.B., Mantell S.H. & Blake J.<br />

(eds.) Advances in the Chemical Manipulation <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong><br />

<strong>Culture</strong>s. Monograph 16, British <strong>Plant</strong> Growth Regulator Group,<br />

Bristol.<br />

HU T.C., ZIAUDDIN A., SIMION E. & KASHA K.J. 1995 Isolated<br />

microscopore culture <strong>of</strong> wheat (Triticum aestivum L.) in a<br />

defined media. I: Effects <strong>of</strong> pretreatment, isolation methods, and<br />

hormones. In Vitro Cell. Dev. Biol., <strong>Plant</strong>. 31, 79-83.<br />

HUANG B., BIRD S., KEMBLE R., MIKI B. & KELLER W.<br />

1991 <strong>Plant</strong> regeneration from microspore-derived embryos <strong>of</strong><br />

Brassica napus: effect <strong>of</strong> embryo age, culture temperature,<br />

osmotic pressure and abscisic acid. In Vitro Cell. Dev. <strong>Plant</strong> 27,<br />

28-31.<br />

HUSEMANN W., AMINO S., FISCHER K., HERZBECK H. &<br />

CALLIS R. 1990 Light dependent growth and differentiation<br />

processes in photoautotrophic cell cultures. pp. 373-378 in<br />

Nijkamp et al. (eds.) 1990 (q.v.).<br />

HYNDMAN S.E., HASEGAWA P.M. & BRESSAN R.A. 1982 <strong>The</strong><br />

role <strong>of</strong> sucrose and nitrogen in adventitious root formation on<br />

cultured rose shoots. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 1, 229-238.<br />

ICHI T., KODA T., ASAI I., HATANAKA A. & SEKIYA J. 1986<br />

Effects <strong>of</strong> gelling agents on in vitro culture <strong>of</strong> plant tissues.<br />

Agric. Biol. Chem. 50, 2397-2399.<br />

ICHIMURA K. & ODA M. 1995 Stimulation <strong>of</strong> root growth in<br />

several vegetables by wood pulp extract. J. Jpn Soc. Hortic. Sci.<br />

63 (Suppl. 4), 797-803.<br />

IKEDA-IWAI, M., UMEHARA M., SATOH S. & KAMADA H.<br />

2003 Stress-induced somatic embryogenesis in vegetative tissues<br />

<strong>of</strong> Arabidopsis thaliana. <strong>Plant</strong> J. 34, 107-114.<br />

ILIĆ-GRUBOR K., ATTREE S. M. & FOWKE L. C. 1998<br />

Induction <strong>of</strong> microspore-derived embryos <strong>of</strong> Brassica napus L.<br />

with polyethylene glycol (PEG) as osmoticum in a low sucrose<br />

medium <strong>Plant</strong> Cell Rep. 17, 329-333.<br />

IMAMURA J. &. HARADA H. 1980. Effects <strong>of</strong> abscisic acid and<br />

water stress on the embryo and plantlet formation in anther<br />

culture <strong>of</strong> Nicotiana tabacum cv. Samsun. Z. Pflanzenphysiol.<br />

100, 285-289.<br />

IMAMURA J., OKABE E., KYO M. & HARADA H. 1982<br />

Embryogenesis and plantlet formation through direct culture <strong>of</strong><br />

isolated pollen <strong>of</strong> Nicotiana tabacum cv. Samsum and Nicotiana<br />

rustica cv. rustica. <strong>Plant</strong> Cell Physiol. 23, 713-716.<br />

INOUE M. & MAEDA E. 1982 Control <strong>of</strong> organ formation in rice<br />

callus using two-step culture method. pp. 183-184 in Fujiwara A.<br />

(ed.) 1982 (q.v.).<br />

ISHIHARA A. & KATANO M. 1982 Propagation <strong>of</strong> apple<br />

cultivars and rootstocks by shoot-tip culture. pp. 733-734 in<br />

Fujiwara A. (ed.) 1982 (q.v.).<br />

Chapter 4 163<br />

JACOBSON L., COOPER B.R. & VOLZ M.G. 1971 <strong>The</strong><br />

interaction <strong>of</strong> pH and aeration in Cl uptake by barley roots.<br />

Physiol. <strong>Plant</strong>. 25, 432-435.<br />

JACQUIOT C. 1951 Action <strong>of</strong> meso-inositol and <strong>of</strong> adenine on<br />

bud formation in the cambium tissue <strong>of</strong> Ulmus campestris<br />

cultivated in vitro. Compt. Rend. Acad. Sci. Paris 233, 815-817.<br />

JAIN R.K., DAVEY M.R. COCKING E.C. & WU R. 1997<br />

Carbohydrate and osmotic requirements for high-frequency plant<br />

regeneration from protoplast-derived colonies <strong>of</strong> indica and<br />

japonica rice varieties. J. Exp. Bot. 48, 751-758.<br />

JAY-ALLEMAND C., CAPELLI P. & CORNU D. 1992 Root<br />

development <strong>of</strong> in vitro hybrid walnut microcuttings in a<br />

vermiculite-containing gelrite medium. Sci. Hortic. 51, 335-342.<br />

JEANNIN G., BRONNER R. & HAHNE G. 1995 Somatic<br />

embryogenesis and organogenesis induced on the immature<br />

zygotic embryo <strong>of</strong> sunflower (Helianthus annus L.) cultivated in<br />

vitro: Role <strong>of</strong> Sugar. <strong>Plant</strong> Cell Rep. 15, 200-204.<br />

JEFFS R.A. & NORTHCOTE D.H. 1967 <strong>The</strong> influence <strong>of</strong> IAA<br />

and sugar on the pattern <strong>of</strong> induced differentiation in plant tissue<br />

culture. J. Cell Sci. 2, 77-88.<br />

JEONG B.R., FUJIWARA K. & KOZAI T. 1995 Environmental<br />

control and photoautotrophic micropropagation. Hortic. Rev. 17,<br />

123-170.<br />

JOHANSSON L. & ERIKSSON T. 1984 Effects <strong>of</strong> carbon dioxide<br />

in anther cultures. Physiol. <strong>Plant</strong> 60, 26-30.<br />

JOHNSON J.L. & EMINO E.R. 1979 In vitro propagation <strong>of</strong><br />

Mammillaria elongata. HortScience 14, 605-606.<br />

JOHRI B.M. & GUHA S. 1963 In vitro development <strong>of</strong> onion<br />

plants from flowers. pp. 215-223 in Maheshwari and Ranga<br />

Swamy (eds.) 1963 (q.v.).<br />

JOY IV R.W., PATEL K.R. & THORPE T.A. 1988 Ascorbic acid<br />

enhancement <strong>of</strong> organogenesis in tobacco. <strong>Plant</strong> Cell <strong>Tissue</strong><br />

Organ Cult. 13, 219-228.<br />

JOY IV R.W., PATEL K.R. & THORPE T.A. 1988 Ascorbic acid<br />

enhancement <strong>of</strong> organogenesis in tobacco callus. <strong>Plant</strong> Cell<br />

<strong>Tissue</strong> Organ Cult. 13, 219-228.<br />

JUMIN H.B. 1995 <strong>Plant</strong> regeneration via somatic embryogenesis<br />

in Citrus and its relatives. Phytomorphology 45, 1-8.<br />

JUNG P., TANNER W. & WOLTER K. 1972 <strong>The</strong> fate <strong>of</strong> myoinositol<br />

in Fraximus tissue cultures. Phytochemistry 11, 1655-1659.<br />

KAHANE R. & RANCILLAC M. 1996 Carbohydrates in onion<br />

cultured in vitro. Acta Bot. Gall. 143, 117-123<br />

KANABUS J., BRESSAN R.A. & CARPITA N.C. 1986 Carbon<br />

assimilation in carrot cells in liquid culture. <strong>Plant</strong> Physiol. 82,<br />

363-368.<br />

KANG K.S., VEEDER G.T., MIRRASOUL P.J., KANEKO T. &<br />

COTTRELL W. 1982 Agar-like polysaccharide produced by a<br />

Pseudomonas species: Production and basic properties. Appl.<br />

Environ. Microbiol. 43, 1086-1091.<br />

KAO K.N. & MICHAYLUK M.R. 1975 Nutritional requirements<br />

for growth <strong>of</strong> Vicia hajastana cells and protoplasts at a very low<br />

population density in liquid media. <strong>Plant</strong>a 126, 105-110.<br />

KARSAI I., BEDO Z. & HAYES P.M. 1994 Effect <strong>of</strong> induction<br />

medium pH and maltose concentration on in vitro androgenesis<br />

<strong>of</strong> hexaploid winter triticale and wheat. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ<br />

Cult. 39, 49-53.<br />

KARTHA K.K. 1981 Meristem culture and cryopreservation –<br />

Methods and applications. pp. 181-211 in Thorpe T.A. (ed.)<br />

<strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong>: Methods and Applications in Agriculture.<br />

Academic Press, New York, London, Toronto, Sydney.<br />

KAUL B. & STABA E.J. 1968 Dioscorea tissue cultures. 1.<br />

Biosynthesis and isolation <strong>of</strong> diesgenin from Dioscorea deltoidea<br />

callus and suspension cells. Lloydia 31, 171-179.<br />

KAUL K. & KOCHHAR T.S. 1985 Growth and differentiation <strong>of</strong><br />

callus cultures <strong>of</strong> Pinus. <strong>Plant</strong> Cell Rep. 4, 180-183.


164 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

KAUL K. & SABHARWAL P.S. 1972 Morphogenetic studies on<br />

Haworthia: Establishment <strong>of</strong> tissue culture and control <strong>of</strong><br />

differentiation. Am. J. Bot. 59, 377-385.<br />

KAUL K. & SABHARWAL P.S. 1975 Morphogenetic studies on<br />

Haworthia: Effects <strong>of</strong> inositol on growth and differentiation. Am.<br />

J. Bot. 62, 655-659.<br />

KAVI KISHOR P.B. & MEHTA A.R. 1982 Some aspects <strong>of</strong><br />

carbohydrate metabolism in organ-differentiating tobacco and<br />

non-differentiating cotton callus cultures. pp. 143-144 in<br />

Fujiwara A. (ed.) 1982 (q.v.).<br />

KAVI KISHOR P.B. & REDDY G.M. 1986 Regeneration <strong>of</strong><br />

plants from long-term cultures <strong>of</strong> Oryza sativa L. <strong>Plant</strong> Cell Rep.<br />

5, 391-393.<br />

KETELLAPPER H.J. 1953 <strong>The</strong> mechanism <strong>of</strong> the action <strong>of</strong> indole-<br />

3-acetic acid on the water absorption by Avena coleoptile<br />

sections. Acta Bot. Neerl. 2, 387-444.<br />

KIM K.S., DAVELAAR E. & DE KLERK G.J. (1994) Abscisic<br />

acid controls dormancy development and bulb formation in lily<br />

plantlets regenerated in vitro. Physiol. <strong>Plant</strong>. 90, 59-64.<br />

KHAN A., CHAUHAN Y.S. & ROBERTS L.W. 1986 In vitro<br />

studies on xylogenesis in citrus-fruit vesicles. 2. Effect <strong>of</strong> pH <strong>of</strong><br />

the nutrient medium on the induction <strong>of</strong> cytodifferentiation. <strong>Plant</strong><br />

Sci. 46, 213-216.<br />

KIM Y.-H. & JANICK J. 1989b Somatic embryogenesis and<br />

organogenesis in cucumber. HortScience 24, 702.<br />

KIMBALL S.L., BEVERSDORF W.D. & BINGHAM E.T. 1975<br />

Influence <strong>of</strong> osmotic potential on the growth and development <strong>of</strong><br />

soybean tissue cultures. Crop Sci. 15, 750-752.<br />

KING P.J. & STREET H.E. 1977 Growth patterns in cell cultures.<br />

pp. 307-387 in Street H.E. (ed.) <strong>Plant</strong> <strong>Tissue</strong> and Cell <strong>Culture</strong>.<br />

Bot. Monographs Vol.11, Blackwell Scientific Public-ations.<br />

Oxford, London.<br />

KINNERSLEY A.M. & HENDERSON W.E. 1988 Alternative<br />

carbohydrates promote differentiation <strong>of</strong> plant cells. <strong>Plant</strong> Cell<br />

<strong>Tissue</strong> Organ Cult. 15, 17-22.<br />

KINNERSLEY A.M. & HENDERSON W.E. 1988 Alternative<br />

carbohydrates promote differentiation <strong>of</strong> plant cells. <strong>Plant</strong> Cell<br />

<strong>Tissue</strong> Organ Cult. 15, 3-16.<br />

KIRDMANEE C., KITAYA Y. & KOZAI T. 1995. Effects <strong>of</strong> CO2<br />

enrichment and supporting material in-vitro on photoautotrophic<br />

growth <strong>of</strong> eucalyptus plantlets in-vitro and ex-vitro In Vitro Cell.<br />

Dev. –Pl. 31, 144-149<br />

KIRKHAM M.B. & HOLDER P.L. 1981 Water osmotic and turgor<br />

potentials <strong>of</strong> kinetin-treated callus. HortScience 16, 306-307.<br />

KITAMURA Y., IKENAGA T., OOE Y., HIRAOKA N. &<br />

MIZUKAMI H. 1998 Induction <strong>of</strong> furanocoumarin biosynthesis<br />

in Glehnia littoralis cell suspension cultures by elicitor treatment.<br />

Phytochemistry 48, 113-117.<br />

KLEIN R.M. & MANOS G.E. 1960 Use <strong>of</strong> metal chelates for plant<br />

tissue cultures. Ann. NY. Acad. Sci. 88, 416-425.<br />

KNUDSON L. 1946 A new nutrient solution for the germination <strong>of</strong><br />

orchid seed. Am. Orchid Soc. Bull. 15, 214-217.<br />

KOCH K.E. 1996 Carbohydrate-modulated gene expression in<br />

plants. Annu. Rev. <strong>Plant</strong> Physiol. <strong>Plant</strong> Mol. Biol. 47, 509-540.<br />

KOCHBA J. & BUTTON J. 1974 <strong>The</strong> stimulation <strong>of</strong><br />

embryogenesis and embryoid development in habituated ovular<br />

callus from the ‘Shamouti’ orange (Citrus sinensis) as affected by<br />

tissue age and sucrose concentration. Z. Pflanzenphysiol. 73,<br />

415-421.<br />

KOCHBA J., SPIEGEL-ROY P., SAAD S. & NEUMANN H.<br />

1978 Stimulation <strong>of</strong> embryogenesis in Citrus tissue culture by<br />

galactose. Naturwissenschaften 65, 261-262.<br />

KOETJE D.S., GRIMES H.D., WANG Y.-C. & HODGES T.K.<br />

1989 Regeneration <strong>of</strong> indica rice (Oryza sativa L.) from primary<br />

callus from immature embryos. J. <strong>Plant</strong> Physiol. 13, 184-190.<br />

KOMOR E., ROTTER M. & TANNER W. 1977 A protoncotransport<br />

system in a higher plant: sucrose transport in Ricinus<br />

communis. <strong>Plant</strong> Sci. Lett. 9, 153-162.<br />

KOMOR E., THOM M. & MARETZKI A. 1981 <strong>The</strong> mechanism <strong>of</strong><br />

sugar uptake by sugarcane suspension cells. <strong>Plant</strong>a 153, 181-192.<br />

KOOP H.U., WEBER G. & SCHWEIGER H.G. 1983 Individual<br />

culture <strong>of</strong> selected single cells and protoplasts <strong>of</strong> higher plants in<br />

microdroplets <strong>of</strong> defined media. Z. Pflanzenphysiol. 112, 21-34.<br />

KORDAN H.A. 1959 Proliferation <strong>of</strong> excised juice vesicles <strong>of</strong><br />

lemon in vitro. Science 129, 779-780.<br />

KOTT L.S. & BEVERSDORF W.D. 1990 Enhanced plant<br />

regeneration from microspore-derived embryos <strong>of</strong> Brassica<br />

napus by chilling, partial desiccation and age selection. <strong>Plant</strong><br />

Cell <strong>Tissue</strong> Organ Cult. 23, 187-194.<br />

KOZAI T. & SEIKIMOTO K. 1988. Effects <strong>of</strong> the number <strong>of</strong> air<br />

changes per hour <strong>of</strong> the closed vessel and the photosynthetic<br />

photon flux on the carbon dioxide concentration inside the vessel<br />

and the growth <strong>of</strong> strawberry plantlets in vitro. Environ. Contr.<br />

Biol. 26, 21-29.<br />

KOZAI T., FUJIWARA K. & KITAYA Y. 1995 Modelling,<br />

measurement and control in plant tissue culture. Acta Hortic.<br />

393, 63-73.<br />

KOZAI T. 1991a Photoautotrophic micropropagation. In Vitro<br />

Cell. Dev. – Pl. 27, 47-51.<br />

KOZAI T. 1991b Micropropagation under photoautotrophic<br />

conditions. pp. 447-469 in Debergh P.C. and Zimmerman R.H.<br />

(eds.) Micropropagation. Technology and Application. Kluwer<br />

Academic Publishers, Dordrecht, Boston, London. ISBN 0-7923-<br />

0818-2.<br />

KRIEG N.R. & GERHARDT P. 1981 pp. 143-156 in Gerhardt P.<br />

Murray R.G.E., Costilow R.N., Nester E.W., Wood W.A., Krieg<br />

N.R. and Briggs Phillips G. Manual <strong>of</strong> Methods for General<br />

Bacteriology. Am. Soc. Microbiol. Washington D.C.<br />

KROMER K. & KUKULCZANKA K. 1985 In vitro cultures <strong>of</strong><br />

meristem tips <strong>of</strong> Canna indica L. Acta Hortic. 167, 279-285.<br />

KUMAR A.S., GAMBORG O.L. & NABORS M.W. 1988 <strong>Plant</strong><br />

regeneration from suspension cultures <strong>of</strong> Vigna aconitifolia.<br />

<strong>Plant</strong> Cell Rep. 7, 138-141.<br />

KUNITAKE H., IMAMIZO H. & M<strong>II</strong> M. 1993 Somatic<br />

embryogenesis and plant regeneration from immature seedderived<br />

calli <strong>of</strong> rugosa rose (Rosa rugosa Thumb). <strong>Plant</strong> Sci. 90,<br />

187-194.<br />

KUNITAKE H., NAKASHIMA T., MORI, K. & TANAKA, M.<br />

1997 Normalization <strong>of</strong> asparagus somatic embryogenesis using a<br />

maltose-containing medium. J. <strong>Plant</strong> Physiol. 150, 458-461.<br />

KURAISHI S. & OKUMURA F.S. 1961 A new green-leaf growth<br />

stimulating factor, phyllococosine, from coconut milk. Nature<br />

189, 148-149.<br />

KURATA K., IBARAKI Y. & GOTO E. 1991 System for micropropagation<br />

by nutrient mist supply. Trans. ASAE 34, 621-624.<br />

KURITAKE H., NAKASHIMA T., MORI K. & TANAKA M.<br />

1997 Normalization <strong>of</strong> asparagus somatic embryogenesis using a<br />

maltose-containing medium. J. <strong>Plant</strong> Physiol. 150, 458-461.<br />

KURKDJIAN A., MATHIEU Y. & GUERN J. 1982 Evidence for<br />

an action <strong>of</strong> 2,4-dichlorophenoxyacetic acid on the vacuolar pH<br />

<strong>of</strong> Acer pseudoplatanus cells in suspension culture. <strong>Plant</strong> Sci.<br />

Lett. 27, 77-86.<br />

LA RUE C.D. 1949 <strong>Culture</strong>s <strong>of</strong> the endosperm <strong>of</strong> maize. Am. J.<br />

Bot. 36, 798.<br />

LAMOTTE C.E. 1960 <strong>The</strong> effects <strong>of</strong> tyrosine and other amino<br />

acids on the formation <strong>of</strong> buds in tobacco callus. Dissertation<br />

Abstracts 21, 31-32.<br />

LANDRY L.G. & SMYTH D.A. 1988 Characterization <strong>of</strong> starch<br />

produced by suspension cultures <strong>of</strong> Indica rice (Oryza sativa L.).<br />

<strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 15, 23-32.


LANG A.R.G. 1967 Osmotic coefficients and water potentials <strong>of</strong><br />

sodium chloride solutions from 0 to 40ºC. Austr. J. Chem. 20,<br />

2017-2023.<br />

LANGFORD P.J. & WAINWRIGHT H. 1987 Effects <strong>of</strong> sucrose<br />

concentration on the photosynthetic ability <strong>of</strong> rose shoots in vitro.<br />

Ann. Bot. 60, 633-640.<br />

LANGHANSOVÁ L., KONRÁDOVÁ H. & VANĔK T. 2004<br />

Polyethylene glycol and abscisic acid improve maturation and<br />

regeneration <strong>of</strong> Panax ginseng somatic embryos. <strong>Plant</strong> Cell Rep.<br />

22, 725-730.<br />

LAPEÑA L., PÉREZ-BERMÚDEZ P. & SEGURA J. 1988<br />

Morphogenesis in hypocotyl cultures <strong>of</strong> Digitalis obscura:<br />

influence <strong>of</strong> carbohydrate levels and sources. <strong>Plant</strong> Science 57,<br />

247-252.<br />

LARKIN P.J., DAVIES P.A. & TANNER G.J. 1988 Nurse culture<br />

<strong>of</strong> low numbers <strong>of</strong> Medicago and Nicotiana protoplasts using<br />

calcium alginate beads. <strong>Plant</strong> Sci. 58, 203-210.<br />

LARKIN P.J., DAVIES P.A. & TANNER G.J. 1988. Nurse<br />

cultures <strong>of</strong> low numbers <strong>of</strong> Medicago and Nicotiana protoplasts<br />

using calcium alginate beads. <strong>Plant</strong> Sci. 58, 203-210.<br />

LAROSA P.C., HASEGAWA P.M. & BRESSAN R.A. 1981<br />

Initiation <strong>of</strong> photoautotrophic potato cells. HortScience 16, 433.<br />

LASSOCINSKI W. 1985 Chlorophyll-deficient cacti in tissue<br />

cultures. Acta Hortic. 167, 287-293.<br />

LAZZERI P.A., HILDEBRAND D.F., SUNEGA J., WILLIAMS<br />

E.G. & COLLINS G.B. 1988 Soybean somatic embryogenesis:<br />

interactions between sucrose and auxin. <strong>Plant</strong> Cell Rep. 7, 517-520.<br />

LEMOS E.E.P. & BLAKE J. 1996 Micropropagation <strong>of</strong> juvenile<br />

and adult Annona squamosa. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 46,<br />

77-79.<br />

LETHAM D.S. 1966 Regulation <strong>of</strong> cell division in plant tissues. <strong>II</strong>.<br />

A cytokinin in plant extracts: isolation and interaction with other<br />

growth regulators. Phytochemistry 5, 269-286.<br />

LETHAM D.S. 1968 A new cytokinin bioassay and the naturally<br />

occurring cytokinin complex. pp. 19-31 in Wightman and<br />

Setterfield (eds.) Biochemistry and Physiology <strong>of</strong> <strong>Plant</strong> Growth<br />

Substances. Runge Press, Ottawa.<br />

LETHAM D.S. 1974 Regulators <strong>of</strong> cell division in plant tissues.<br />

XX. <strong>The</strong> cytokinins <strong>of</strong> coconut milk. Physiol. <strong>Plant</strong>. 32, 66-70.<br />

LETHAM D.S. 1982 A compound in coconut milk which actively<br />

promoted radish cotyledon expansion, but exhibited negligible<br />

activity in tissue culture bioassays for cytokinins, was identified<br />

as the 6 oxypurine, 2-(3-methylbut-2-enylamino)-purin-6-one.<br />

<strong>Plant</strong> Sci. Lett. 26, 241-249.<br />

LEVA A.R., BARROSO M. & MURILLO J.M. 1984 La<br />

moltiplicazione del melo con la tecnica della micropropagazione.<br />

Vriazione del pH in substrati diversi durante la fase di<br />

multiplicazione. Riv. Ort<strong>of</strong>lor<strong>of</strong>rutti. Ital. 68, 483-492.<br />

LI X.Y. & HUANG F.H.. 1996 Induction <strong>of</strong> somatic<br />

embryogenesis in Loblolly pine (Pinus taeda L.). In Vitro Cell.<br />

Dev. – Pl. 32, 129-135.<br />

LI X.Y., HUANG F.H., MURPHY J.B. & GBUR E.R. JR. 1998<br />

Polyethylene glycol and maltose enhance somatic embryo<br />

maturation in Loblolly pine (Pinus taeda L.). In Vitro Cell. Dev.<br />

-Pl. 34, 22-26.<br />

LICHTER R. 1981 Anther culture <strong>of</strong> Brassica napus in a liquid<br />

culture medium. Z. Pflanzenphysiol. 103, 229-237.<br />

LIMBERG M., CRESS D. & LARK K.G. 1979 Variants <strong>of</strong><br />

soybean cells which can grow in suspension with maltose as a<br />

carbon-energy source. <strong>Plant</strong> Physiol. 63, 718-721.<br />

LIN M.-L. & STABA E.J. 1961 Peppermint and spearmint tissue<br />

cultures. I. Callus formation and submerged culture. Lloydia 24,<br />

139-145.<br />

LINOSSIER L., VEISSEIRE P., CAILLOUX F. & COUDRET A.<br />

1997 Effects <strong>of</strong> abscisic acid and high concentrations <strong>of</strong> PEG on<br />

Chapter 4 165<br />

Hevea brasiliensis somatic embryos development. <strong>Plant</strong> Sci. 124,<br />

183-191.<br />

LINSMAIER E.M. & SKOOG F. 1965 Organic growth factor<br />

requirements <strong>of</strong> tobacco tissue cultures. Physiol. <strong>Plant</strong>. 18, 100-<br />

127.<br />

LIPAVSKA H. & VREUGDENHIL D. 1996 Uptake <strong>of</strong> mannitol<br />

from the media by in vitro grown plants. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ<br />

Cult.45, 103-107.<br />

LIPPMANN B. & LIPPMANN G. 1984 Induction <strong>of</strong> somatic<br />

embryos in cotyledonary tissue <strong>of</strong> soybean, Glycine max L. Merr.<br />

<strong>Plant</strong> Cell Rep. 3, 215-218.<br />

LITZ R.E. 1988 Somatic embryogenesis from cultured leaf<br />

explants <strong>of</strong> the tropical tree Euphoria longan Stend. J. <strong>Plant</strong><br />

Physiol. 132, 190-193.<br />

LIU Y.K., SEKI M. & FURUSAKI S. 1999 <strong>Plant</strong> cell<br />

immobilization in lo<strong>of</strong>a sponge using two-way bubble circular<br />

system. J. Chem. Eng. Jpn. 32, 8-14.<br />

LLOYD D., ROBERTS A.V. & SHORT K.C. 1988 <strong>The</strong> induction<br />

in vitro <strong>of</strong> adventitious shoots in Rosa. Euphytica 37, 31-36.<br />

LLOYD G. & McCOWN B. 1981 Commercially-feasible<br />

micropropagation <strong>of</strong> Mountain laurel, Kalmia latifolia, by use <strong>of</strong><br />

shoot tip culture. Int. <strong>Plant</strong> Prop. Soc. Proc. 30, 421-427.<br />

LLOYD G. & McCOWN B. 1981 Commercially-feasible<br />

micropropagation <strong>of</strong> Mountain laurel, Kalmia latifolia, by use <strong>of</strong><br />

shoot tip culture. Int. <strong>Plant</strong> Prop. Soc. Proc. 30, 421-427.<br />

LOEWUS F.A. & LOEWUS M.W. 1980 Myo-inositol:<br />

Biosynthesis and metabolism. pp. 43-76 in Stumpf and Conn<br />

(eds.): <strong>The</strong> Biochemistry <strong>of</strong> <strong>Plant</strong>s 3. Academic Press N.Y. 43-76.<br />

LOEWUS F.A. 1974 <strong>The</strong> biochemistry <strong>of</strong> myo-inositol in plants.<br />

Recent Adv. Phytochem. 8, 179-207.<br />

LOEWUS F.A., KELLY S. & NEUFELD E.F. 1962 Metabolism<br />

<strong>of</strong> myo-inositol in plants: Conversion to pectin hemicellulose, Dxylose<br />

and sugar acids. Proc. Natl. Acad. Sci. USA 48, 421-425.<br />

LOU H. & KAKO S. 1995 Role <strong>of</strong> high sugar concentrations in<br />

inducing somatic embryogenesis from cucumber cotyledons. Sci.<br />

Hortic. 64, 11-20.<br />

LOVELL P.H., ILLSLEY A. & MOORE K.G. 1972 <strong>The</strong> effects <strong>of</strong><br />

light intensity and sucrose on root formation, photosynthetic<br />

ability, and senescence in detached cotyledons <strong>of</strong> Sinapis alba L.<br />

and Raphanus sativus L. Ann. Bot. 36, 123-134.<br />

LU C., VASIL I.K. & OZIAS-AKINS P. 1982 Somatic embryogenesis<br />

in Zea mays. <strong>The</strong>or. Appl. Genet. 62, 109-112.<br />

LUMSDEN P.J., PRYCE S. & LEIFERT C. 1990 Effect <strong>of</strong><br />

mineral nutrition on the growth and multiplication <strong>of</strong> in vitro<br />

cultured plants. pp. 108-113 in Nijkamp et al. (eds.) 1990 (q.v.).<br />

MADHUSUDHAN R., RAS S.R. & RAVISHANKAR G.A. 1995<br />

Osmolarity as a measure <strong>of</strong> growth <strong>of</strong> plant cells in suspension<br />

cultures. Enz. Microb. Tech. 17, 989-991.<br />

MAPES M.O. & ZAERR J.B. 1981 <strong>The</strong> effect <strong>of</strong> the female<br />

gametophyte on the growth <strong>of</strong> cultured Douglas fir embryos.<br />

Ann. Bot. 48, 577-582.<br />

MAHESHAWARI P. AND RANGASWAMY N.S. 1963 (eds.).<br />

<strong>Plant</strong> <strong>Tissue</strong> and Organ <strong>Culture</strong>. Int. Soc. <strong>Plant</strong> Morphologists,<br />

Univ. Delhi, India.<br />

MARETZKI A. & THOM, M. 1978 Characteristics <strong>of</strong> a galactose-adapted<br />

sugarcane cell line grown in suspension culture.<br />

<strong>Plant</strong> Physiol. 61, 544-548.<br />

MARETZKI A., DELA-CRUZ A. & NICKELL L.G. 1971<br />

Extracellular hydrolysis <strong>of</strong> starch in sugarcane cell suspensions.<br />

<strong>Plant</strong> Physiol. 48, 521-525.<br />

MARETZKI A., THOM A. & NICKELL L.G. 1972 Influence <strong>of</strong><br />

osmotic potentials on the growth and chemical composition <strong>of</strong><br />

sugarcane cell cultures. Hawaii <strong>Plant</strong> Res. 58, 183-199.<br />

MARETZKI A., THOM M. & NICKELL L.G. 1974 Utilisation<br />

and metabolism <strong>of</strong> carbohydrates in cell and callus cultures. pp.


166 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

329-361 in Street H.E. (ed.) <strong>Tissue</strong> <strong>Culture</strong> and <strong>Plant</strong> Science.<br />

Academic Press, London, New York, San Francisco.<br />

MARGARA J. & RANCILLAC M. 1966 Observations<br />

préliminaires sur le rôle du milieu nutritif dans l’initiation florale<br />

des bourgeons né<strong>of</strong>ormés in vitro chez Cichorium intybus L.<br />

Compt. Rend. Acad. Sci. Paris 263D, 1455-1458.<br />

MARSOLAIS A.A., WILSON D.P.M., TSUJITA M.J. &<br />

SENARATNA T. 1991 Somatic embryogenesis and artificial seed<br />

production in Zonal (Pelargonium horterum) and Regal<br />

(Pelargonium domesticum) geranium. Can. J. Bot. 69, 1188-1193.<br />

MARTIN S.M. & ROSE D. 1976 Growth <strong>of</strong> plant cell (Ipomoea)<br />

suspension cultures at controlled pH levels. Can. J. Bot. 54,<br />

1264-1270.<br />

MARUYAMA E., ISH<strong>II</strong> K. & KINOSHITA I. 1998 Alginate<br />

encapsulation technique and cryogenic procedures for long-term<br />

storage <strong>of</strong> the tropical forest tree Guazuma crinita Mart. in vitro<br />

cultures. Jarq-Jpn. Agr. Res. Q. 32, 301-309.<br />

MATHES M.C., MORSELLI M. & MARVIN J.W. 1973 Use <strong>of</strong><br />

various carbon sources by isolated maple callus cultures. <strong>Plant</strong><br />

Cell Physiol. 14, 797-801.<br />

MAUSETH J.D. 1979 Cytokinin-elicited formation <strong>of</strong> the pith-rib<br />

meristem and other effects <strong>of</strong> growth regulators on the<br />

morphogenesis <strong>of</strong> Echino cereus (Cactaceae) seedling shoot<br />

apical meristems. Am. J. Bot. 66, 446-451.<br />

McCANCE R.A. & WIDDOWSON E.M. 1940 <strong>The</strong> Chemical<br />

Composition <strong>of</strong> Foods. British M.R.C. Spec. Rep. Serv. No. 235.<br />

MCGREGOR L.J. & MCHUGHEN A. 1990 <strong>The</strong> influence <strong>of</strong><br />

various cultural factors on anther culture <strong>of</strong> four cultivars <strong>of</strong> spring<br />

wheat (Triticum aestivum L.). Can. J. <strong>Plant</strong> Sci. 70, 183-192.<br />

MEHTA A.R. 1982 Comparative studies on some physiological<br />

aspects in organ forming (diploid as well as haploid) tobacco and<br />

non-organ forming cotton callus tissues. pp. 125-126 in Fujiwara<br />

A. (ed.) 1982 (q.v.).<br />

MELLOR F.C. & STACE-SMITH R. 1969 Development <strong>of</strong> excised<br />

potato buds in nutrient medium. Can. J. Bot. 47, 1617-1621.<br />

MENON M.K.C. & LAL M. 1972 Influence <strong>of</strong> sucrose on the<br />

differentiation <strong>of</strong> cells with zygote-like potentialities in a moss.<br />

Naturwissenschaften 59, 514.<br />

MERKLE S.A., PARROTT W.A. & FLINN B.S. 1995 Morphogenetic<br />

aspects <strong>of</strong> somatic embryogenesis. pp. 155-205 in Thorpe<br />

T.A. (ed.) In Vitro Embryogenesis in <strong>Plant</strong>s. Kluwer Acad. Publ.,<br />

Dordrecht.<br />

MERYL SMITH M. & STONE B.A. 1973 Studies on Lolium<br />

multiflorum endosperm in tissue culture. 1. Nutrition. Aust. J.<br />

Biol. Sci. 26, 123-133.<br />

MILLER C.O. 1961 A kinetin-like compound in maize. Proc. Natl.<br />

Acad. Sci. USA 47, 170-174.<br />

MILLER J.H. 1968 Fern gametophytes as experimental material.<br />

Bot. Rev. 34, 361-426.<br />

MINOCHA S.C. & HALPERIN W. 1974 Hormones and<br />

metabolites which control tracheid differentiation with or without<br />

concomitant effects on growth in cultured tuber tissue <strong>of</strong><br />

Helianthus tuberosus L. <strong>Plant</strong>a 116, 319-331.<br />

MITCHELL E.D., JOHNSON B.B. & WHITTLE T. 1980 βgalactosidase<br />

activity in cultured cotton cells (Gossypium<br />

hirsutum L.). A comparison between cells growing on sucrose<br />

and lactose. In Vitro 16, 907-912.<br />

MOLNAR S.J. 1988 Nutrient modifications for improved growth<br />

<strong>of</strong> Brassica nigra cell suspension cultures. <strong>Plant</strong> Cell <strong>Tissue</strong><br />

Organ Cult. 15, 257-267.<br />

MONDAL H., MANDAL R.K. & BISWAS B.B. 1972 RNA<br />

polymerase from eukaryotic cells. Isolation and purification <strong>of</strong><br />

enzymes and factors from chromatin <strong>of</strong> coconut nuclei. Eur. J.<br />

Biochem. 25, 463-470.<br />

MONNIER M. 1976 <strong>Culture</strong> in vitro de l’embryon immature de<br />

Capsella bursa-pastoris Moench (L.). Rev. Cytol. Biol. Vég. 39,<br />

1-120.<br />

MONNIER M. 1978 <strong>Culture</strong> <strong>of</strong> zygotic embryos. pp. 277-286 in<br />

Thorpe T. A. (ed.) 1978 Frontiers <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong>. Int.<br />

Assoc. for <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong>. Distrib. by <strong>The</strong> Bookstore,<br />

Univ. Calgary, Alberta, T2N 1N4, Canada.<br />

MOREL G. & MULLER J.-F. 1964 In vitro culture <strong>of</strong> the apical<br />

meristem <strong>of</strong> the potato. Compt. Rend. Acad. Sci., Paris 258,<br />

5250-5252.<br />

MOREL G. & WETMORE R.H. 1951 <strong>Tissue</strong> culture <strong>of</strong><br />

monocotyledons. Am. J. Bot. 38, 138-140.<br />

MOREL G. 1946 Action de l’acid pantothénique sur la croissance<br />

de tissus d’Aubépine cultivés in vitro. Compt. Rend. Acad. Sci.,<br />

Paris 223, 166-168.<br />

MORRIS D.A. 2000 Transmembrane auxin carrier systems –<br />

Dynamic regulators <strong>of</strong> polar auxin transport. <strong>Plant</strong> Growth Reg.<br />

32, 161-172.<br />

MÜLLER J.F., MISSIONIER C. & CABOCHE M. 1983 Low<br />

density growth <strong>of</strong> cells derived from Nicotiana and Petunia<br />

protoplasts: Influence <strong>of</strong> the source <strong>of</strong> protoplasts and<br />

comparison <strong>of</strong> the growth- promoting activity <strong>of</strong> various auxins.<br />

Physiol. <strong>Plant</strong>. 57, 35-41.<br />

MURASHIGE T. & NAKANO R.T. 1968 <strong>The</strong> light requirement<br />

for shoot initiation in tobacco callus culture. Am. J. Bot. 55, 710.<br />

MURASHIGE T. & SKOOG F. 1962 A revised medium for rapid<br />

growth and bio-assays with tobacco tissue cultures. Physiol.<br />

<strong>Plant</strong>. 15, 473-497.<br />

MURASHIGE T. & TUCKER D.P.H. 1969 Growth factor<br />

requirements <strong>of</strong> citrus tissue culture. pp. 1155-1161 in Chapman<br />

H. D. (ed.) Proc. 1st Int. Citrus Symp. Vol. 3, Univ. Calif.,<br />

Riverside Publication.<br />

MURASHIGE T. & TUCKER D.P.H. 1969 Growth factor<br />

requirements <strong>of</strong> citrus tissue culture. pp. 1155-1161 in Chapman<br />

H. D. (ed.) Proc. 1st Int. Citrus Symp. Vol. 3, Univ. Calif.,<br />

Riverside Publication.<br />

MURASHIGE T. 1974 <strong>Plant</strong> propagation through tissue cultures.<br />

Annu. Rev. <strong>Plant</strong> Physiol. 25, 135-166.<br />

MURASHIGE T., SHABDE M.N., HASEGAWA P.N.,<br />

TAKATORI F.H. & JONES J.B. 1972 Propagation <strong>of</strong> asparagus<br />

through shoot apex culture. I. Nutrient media for formation <strong>of</strong><br />

plantlets. J. Am. Soc. Hort. Sci. 97, 158-161.<br />

MURASHIGE T., SERPA M. & JONES J.B. 1974 Clonal<br />

multiplication <strong>of</strong> Gerbera through tissue culture. HortScience 9,<br />

175-180.<br />

MUTAFTSCHIEV S., COUSSON A. & TRAN THANH VAN K.<br />

1987 Modulation <strong>of</strong> cell growth and differentiation by pH and<br />

oligosaccharides. pp. 29-42 in Jackson M.B., Mantell S.H. &<br />

Blake J. (eds.) Advances in the Chemical Manipulation <strong>of</strong> <strong>Plant</strong><br />

<strong>Tissue</strong> <strong>Culture</strong>s. Monograph 16, British <strong>Plant</strong> Growth Regulator<br />

Group, Bristol.<br />

NAIRN B.J. 1988 Significance <strong>of</strong> gelling agents in a production<br />

tissue culture laboratory. Comb. Proc. Int. <strong>Plant</strong> Prop. Soc. 1987<br />

37, 200-205.<br />

NARAYANASWAMI S. & LARUE C.D. 1955 <strong>The</strong> morphogenic<br />

effects <strong>of</strong> various physical factors on the gemmae <strong>of</strong> Lunaria.<br />

Phytomorph. 5, 99-109.<br />

NEAL C.A. & TOPOLESKI L.D. 1983 Effects <strong>of</strong> the basal<br />

medium on growth <strong>of</strong> immature tomato embryos in vitro. J. Am.<br />

Soc. Hortic. Sci. 108, 434-438.<br />

NEGASH A., KRENS F., SCHAART J. & VISSER B. 2001 In<br />

vitro conservation <strong>of</strong> enset under slow-growth conditions. <strong>Plant</strong><br />

Cell <strong>Tissue</strong> Organ Cult. 66, 107-111.<br />

NERNST W. 1904 <strong>The</strong>orie der Reaktiongeschwindigkeit in heterogenen<br />

Systemen. Phys. Chem. 47, 52-55.


NESIUS K.K. & FLETCHER J.S. 1973 Carbon dioxide and pH<br />

requirement <strong>of</strong> non-photosynthetic tissue culture cells. Physiol.<br />

<strong>Plant</strong>. 28, 259-263.<br />

NICHOL J.W., SLADE D., VISS P. & STUARD D.A. 1991 Effect<br />

<strong>of</strong> organic acid pretreatment on the regeneration and<br />

development (conversion) <strong>of</strong> whole plants from callus cultures <strong>of</strong><br />

alfalfa, Medicago sativa L. <strong>Plant</strong> Sci. 79, 181-192.<br />

NICHOL J.W., SLADE D., VISS P. & STUART D.A. 1991<br />

Effect <strong>of</strong> organic acid pretreatment on the regeneration and<br />

development (conversion) <strong>of</strong> whole plants from callus cultures <strong>of</strong><br />

alfalfa, Medicago sativa L. <strong>Plant</strong> Sci. 79, 181-192.<br />

NICKELL L.G. & BURKHOLDER P.R. 1950 Atypical growth <strong>of</strong><br />

plants. <strong>II</strong>. Growth in vitro <strong>of</strong> virus tumors <strong>of</strong> Rumex in relation to<br />

temperature, pH and various sources <strong>of</strong> nitrogen, carbon and<br />

sulfur. Am. J. Bot. 37, 538-547.<br />

NICKELL L.G. & BURKHOLDER P.R. 1950 Atypical growth <strong>of</strong><br />

plants. <strong>II</strong>. Growth in vitro <strong>of</strong> virus tumors <strong>of</strong> Rumex in relation to<br />

temperature, pH and various sources <strong>of</strong> nitrogen, carbon and<br />

sulfur. Am. J. Bot. 37, 538-547.<br />

NICKELL L.G. & MARETZKI A. 1969 Growth <strong>of</strong> suspension<br />

cultures <strong>of</strong> sugarcane cells in chemically defined media. Physiol.<br />

<strong>Plant</strong>. 22, 117-125.<br />

NICKELL L.G. & MARETZKI A. 1970 <strong>The</strong> utilization <strong>of</strong> sugars<br />

and starch as carbon sources by sugarcane cell suspension<br />

cultures. <strong>Plant</strong> Cell Physiol. 11, 183-185.<br />

NIJKAMP H.I.J., VAN DER PLAS I.H.W. & VAN AARTRIJK J.<br />

(eds.) Progress in <strong>Plant</strong> Cellular and Molecular Biology. Proc.<br />

V<strong>II</strong>th Int. Cong. on <strong>Plant</strong> <strong>Tissue</strong> and Cell <strong>Culture</strong>. Amsterdam,<br />

<strong>The</strong> Netherlands. 24-29 June 1990. Kluwer Academic Publishers,<br />

Dortrecht, Netherlands.<br />

NITSCH J.P. & NITSCH C. 1956 Auxin-dependent growth <strong>of</strong><br />

excised Helianthus tissues. Am. J. Bot. 43, 839-851.<br />

NITSCH J.P. & NITSCH C. 1965 Né<strong>of</strong>ormation de fleurs in vitro<br />

chez une espèce de jours courts: Plumbaga indica L. Ann. Phys.<br />

Vég. 7, 251-256.<br />

NITZSCHE W. & WENZEL G. 1977 Haploids in <strong>Plant</strong> Breeding.<br />

Suppl. 8 to Z. Pflanzenzucht. Parey. Berlin and Hamburg.<br />

NOH E.W., MINOCHA S.C. & RIEMENSCHNEIDER D.E. 1988<br />

Adventitious shoot formation from embryonic explants <strong>of</strong> red<br />

pine (Pinus resinosa). Physiol. <strong>Plant</strong>. 74, 119-124.<br />

NORGAARD J.V. 1997 Somatic embryo maturation and plant regeneration<br />

in Abies nordmanniana LK. <strong>Plant</strong> Sci. 124, 211-221.<br />

NORSTOG K. & SMITH J. 1963 <strong>Culture</strong> <strong>of</strong> small barley embryos<br />

on defined media. Science 142, 1655-1656.<br />

NORSTOG K. & SMITH J. 1963 <strong>Culture</strong> <strong>of</strong> small barley embryos<br />

on defined media. Science 142, 1655-1656.<br />

NORSTOG K. 1965b Induction <strong>of</strong> apogamy in megagametophytes<br />

<strong>of</strong> Zamia integrifolia. Am. J. Bot. 52, 993-999.<br />

NORSTOG K. 1973 New synthetic medium for the culture <strong>of</strong><br />

premature barley embryos. In Vitro 8, 307-308.<br />

OBATA-SASAMOTO H. & THORPE T.A. 1983 Cell wall<br />

invertase activity in cultured tobacco tissues. <strong>Plant</strong> Cell <strong>Tissue</strong><br />

Organ Cult. 2, 3-9.<br />

OBATA-SASAMOTO H., BEAUDOIN-EAGAN L. & THORPE<br />

T.A. 1984 C-metabolism during callus induction in tobacco.<br />

<strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 3, 291-299.<br />

OJIMA K. & OHIRA K. 1980 <strong>The</strong> exudation and characterization<br />

<strong>of</strong> iron-solubilizing compounds in a rice cell suspension culture.<br />

<strong>Plant</strong> Cell Physiol. 21, 1151-1161.<br />

OKA S. & OHYAMA K. 1982 Sugar utilization in mulberry<br />

(Morus alba L.) bud culture. pp. 67-68 in Fujiwara A. (ed.) 1982<br />

(q.v.).<br />

OWEN H.R., WENGERD D. & MILLER A.R. 1991 <strong>Culture</strong><br />

medium pH influenced by basal medium, carbohydrate source,<br />

gelling agent, activated charcoal, and medium storage method.<br />

<strong>Plant</strong> Cell Rep. 10, 583-586.<br />

Chapter 4 167<br />

OWENS L.D. & WOZNIAK C.A. 1991 Measurement and effects<br />

<strong>of</strong> gel matric potential and expressibility on production <strong>of</strong><br />

morphogenic callus by cultured sugarbeet leaf discs. <strong>Plant</strong> Cell<br />

<strong>Tissue</strong> Organ Cult. 26, 127-133.<br />

PAMPLIN E.J. & CHAPMAN J.M. 1975 Sucrose suppression <strong>of</strong><br />

chlorophyll synthesis in tissue cultures. J. Exp. Bot. 26, 212-220.<br />

PARFITT D.E., ALMEHDI A.A. & BLOKSBERG L.N. 1988 Use<br />

<strong>of</strong> organic buffers in plant tissue-culture systems. Sci. Hortic. 36,<br />

157-163.<br />

PARIS D. & DUHAMET L. 1953 Action d’un mélange d’acides<br />

aminés et vitamines sur la proliferation des cultures de croun-gall<br />

de Scorsonère; comparison avec l’action du lait de coco. Compt.<br />

Rend. Acad. Sci. Paris 236, 1690-1692.<br />

PARR D.R. & EDELMAN J. 1975 Release <strong>of</strong> hydrolytic<br />

enzymes from the cell walls <strong>of</strong> intact and disrupted carrot callus<br />

tissue. <strong>Plant</strong>a 127, 111-119.<br />

PASQUA G., MANES F., MONACELLI B., NATALE L. &<br />

ANSELMI S. 2002 Effects <strong>of</strong> the culture medium pH and ion<br />

uptake in in vitro vegetative organogenesis in thin cell layers <strong>of</strong><br />

tobacco. <strong>Plant</strong> Sci. 162, 947-955.<br />

PASQUALETTO P.-L., WERGIN W.P. & ZIMMERMAN R.H.<br />

1988 Changes in structure and elemental composition <strong>of</strong> vitrified<br />

leaves <strong>of</strong> `Gala’ apple in vitro. Acta Hortic. 227, 352-357.<br />

PASQUALETTO P.-L., ZIMMERMAN R.H. & FORDHAM I.<br />

1986a Gelling agent and growth regulator effects on shoot<br />

vitrification <strong>of</strong> `Gala’ apple in vitro. J. Am. Soc. Hortic. Sci. 111,<br />

976-980 & 112, 407.<br />

PASQUALETTO P.-L., ZIMMERMAN R.H. & FORDHAM I.<br />

1988b <strong>The</strong> influence <strong>of</strong> cation and gelling agent concentrations<br />

on vitrification <strong>of</strong> apple cultivars in vitro. <strong>Plant</strong> Cell <strong>Tissue</strong><br />

Organ Cult. 14, 31-40.<br />

PATEL K.R. & BERLYN G.P. 1983 Cytochemical investigations<br />

on multiple bud formation in tissue cultures <strong>of</strong> Pinus coulteri.<br />

Can. J. Bot. 61, 575-585.<br />

PECH J.C. & ROMANI R.J. 1979 Senescence <strong>of</strong> pear fruit cells<br />

cultured in a continuously renewed auxin-deprived medium.<br />

<strong>Plant</strong> Physiol. 64, 814-817.<br />

PHILLIPS D.V. & SMITH A.E. 1974 Soluble carbohydrates in<br />

soybean. Can. J. Bot. 52, 2447-2452.<br />

PHILLIPS G.C. & COLLINS G.B. 1979 In vitro tissue culture <strong>of</strong><br />

selected legumes and plant regeneration from callus cultures <strong>of</strong><br />

red clover. Crop Sci. 19, 59-64.<br />

PHILLIPS G.C. & COLLINS G.B. 1984 Chapter 7 – Red Clover<br />

and other forage legumes. pp. 169-210 in Sharp W.R., Evans<br />

D.A., Ammirato P.V. & Yamada Y. (eds.) Handbook <strong>of</strong> <strong>Plant</strong><br />

Cell <strong>Culture</strong>. Vol. 2. Crop Species. Macmillan Publishing Co.,<br />

New York, London.<br />

PHILLIPS G.C. & HUBSTENBERGER J.F. 1985 Organogenesis in<br />

pepper tissue cultures. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 4, 261-269.<br />

PICCIONI E. & STANDARDI A. 1995 Encapsulation <strong>of</strong> micropropagated<br />

buds <strong>of</strong> six woody species. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ<br />

Cult. 42, 221-226.<br />

PICCIONI E. 1997 <strong>Plant</strong>lets from encapsulated micropropagated<br />

buds <strong>of</strong> M.26 apple rootstock. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 47,<br />

255-260.<br />

PIERIK R.L.M. 1988 In vitro culture <strong>of</strong> higher plants as a tool in<br />

the propagation <strong>of</strong> horticultural crops. Acta Hortic. 226, 25-40.<br />

PIERIK R.L.M., SPRENKELS P.A., VAN DER HARST B. &<br />

VAN DER MEYS Q.G. 1988 Seed germination and further<br />

development <strong>of</strong> plantlets <strong>of</strong> Paphiopedilum ciliolare Pfitz. in<br />

vitro. Sci. Hortic. 34, 139-153.<br />

PLIEGO-ALFARO F. 1988 Development <strong>of</strong> an in-vitro rooting<br />

bioassay using juvenile-stage stem cuttings <strong>of</strong> Persea americana<br />

Mill. J. Hort. Sci. 63,295-301.


168 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

POLIKARPOCHKINA R.T., GAMBURG K.Z. & KHAVIN E.E.<br />

1979 Cell-suspension culture <strong>of</strong> maize (Zea mays L.). Z.<br />

Pflanzenphysiol. 95, 57-67.<br />

POLLARD J.K., SHANTZ E.M. & STEWARD F.C. 1961<br />

Hexitols in coconut milk: their role in the nurture <strong>of</strong> dividing<br />

cells. <strong>Plant</strong> Physiol. 36, 492-501.<br />

PREECE J.E. & SUTTER E.G. 1991 Acclimatization <strong>of</strong> micropropagated<br />

plants to the greenhouse and field. pp. 71-93 in<br />

Debergh P.C. and Zimmerman R.H. (eds.) Micropropagation.<br />

Technology and Application. Kluwer Academic Publishers,<br />

Dordrecht, Boston, London. ISBN 0-7923-0818-2.<br />

PREIL W. 1991 Application <strong>of</strong> bioreactors in plant propagation.<br />

pp. 425-445 in Debergh P.C. and Zimmerman R.H. (eds.) Micropropagation.<br />

Technology and Application. Kluwer Academic<br />

Publishers, Dordrecht, Boston, London. ISBN 0-7923-0818-2.<br />

PRETOVA A. & WILLIAMS E.G. 1986 Direct somatic<br />

embryogenesis from immature zygotic embryos <strong>of</strong> flax (Linum<br />

usitatissimum L.). J. <strong>Plant</strong> Physiol. 126, 155-161.<br />

PUA E.C. & CHONG C. 1984 Requirement for sorbitol (Dglucitol)<br />

as carbon source for in vitro propagation <strong>of</strong> Malus<br />

robusta No. 5. Can. J.Bot. 62, 1545-1549.<br />

PUA E.-C., RAGOLSKY E., CHANDLER S.F. & THORPE T.A.<br />

1985a Effect <strong>of</strong> sodium sulfate on in vitro organogenesis <strong>of</strong><br />

tobacco callus. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 5, 55-62.<br />

PUA E.-C., RAGOLSKY E. & THORPE T.A. 1985b Retention <strong>of</strong><br />

shoot regeneration capacity <strong>of</strong> tobacco callus by Na2SO4. <strong>Plant</strong><br />

Cell Rep. 4, 225-228.<br />

RABE E. 1990 Stress physiology: the functional significance <strong>of</strong> the<br />

accumulation <strong>of</strong> nitrogen-containing compounds. J. Hortic. Sci.<br />

65, 231-243<br />

RADLEY M. & DEAR E. 1958 Occurrence <strong>of</strong> gibberellin-like<br />

substances in coconut. Nature 182, 1098.<br />

RAGHAVAN V. 1977 Diets and culture media for plant embryos.<br />

pp. 361-413 in Recheigl M. Jr. (ed.) 1977 CRC Handbook Series<br />

in Nutrition & Food Vol. 4. CRC Press, Baton Rouge, Florida.<br />

RAHMAN M.A. & BLAKE J. 1988 <strong>The</strong> effects <strong>of</strong> medium<br />

composition and culture conditions on in vitro rooting and ex<br />

vitro establishment <strong>of</strong> jackfruit (Artocarpas heterophyllus Lam.).<br />

<strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 13, 189-200.<br />

RAINS D.W. 1989 <strong>Plant</strong> tissue and protoplast culture: application<br />

to stress physiology and biochemistry. In: Jones H.G., Flowers<br />

T.J. & Jones M.B. (eds) <strong>Plant</strong>s under stress. Cambridge<br />

University Press, London, pp 181-196.<br />

RAINS D.W., VALENTINE R.C. & HOLLAENDER A. (eds.)<br />

1980 Genetic Engineering <strong>of</strong> Osmoregulation. Plenum Press,<br />

New York, London.<br />

RAMAGE C.M. & WILLIAMS R.R. 2002 Inorganic nitrogen<br />

requirements during shoot organogenesis in tobacco leaf discs. J.<br />

Exp. Bot 53, 1437-1443.<br />

RAMMING D.W. 1990 <strong>The</strong> use <strong>of</strong> embryo culture in fruit<br />

breeding. HortScience 25, 393-398.<br />

RANGA SWAMY N.S. 1963 Studies on culturing seeds <strong>of</strong><br />

Orobanche aegyptiaca. pp. 345-354 in Maheshwari and Ranga<br />

Swamy (eds.) 1963 (q.v.).<br />

RANGAN T.S. 1984 Clonal propagation. pp. 68-73 in Vasil I.K.<br />

(ed.) Cell <strong>Culture</strong> and Somatic Cell Genetics <strong>of</strong> <strong>Plant</strong>s Vol. 1.<br />

Acad. Press, New York.<br />

RANGAN T.S. 1974 Morphogenic investigations on tissue<br />

cultures <strong>of</strong> Panicum miliaceum. Z. Pflanzenphysiol. 72, 456-459.<br />

RANGAN T.S., MURASHIGE T. & BITTERS W.P. 1968 In<br />

vitro initiation <strong>of</strong> nucellar embryos in monoembryonic Citrus.<br />

HortScience 3, 226-227.<br />

RAQUIN C. 1983 Utilization <strong>of</strong> different sugars as carbon sources<br />

for in vitro anther culture <strong>of</strong> petunia. Z. Pflanzenphysiol. 111,<br />

Suppl., 453-457.<br />

RAVEN J.A. & SMITH F.A. 1976 Cytoplasmic pH regulation and<br />

electrogenic H + extrusion. Curr. Adv. <strong>Plant</strong> Sci. 8, 649-660.<br />

RAVEN J.A. 1986 Biochemical disposal <strong>of</strong> excess H + in growing<br />

plants? New Phytol. 104, 175-206.<br />

RAWAL S.K. & MEHTA A.R. 1982 Changes in enzyme activity<br />

and isoperoxidases in haploid tobacco callus during<br />

organogenesis. <strong>Plant</strong> Sci. Lett. 24, 67-77.<br />

REDENBAUGH K., VISS P., SLADE D. & FUJ<strong>II</strong> J.A. 1987<br />

Scale-up: artificial seeds. pp. 473-493 in Green C.E, Sommers<br />

D.A., Hackett W.P. and Biesboer D.D. (eds.) <strong>Plant</strong> <strong>Tissue</strong> and<br />

Cell <strong>Culture</strong>. Alan R. Liss, Inc., New York<br />

REIDIBOYM-TALLEUX L., DIEMER F., SOURDIOUX M.,<br />

CHAPELAIN K. & DE-MARCH G.G. 1999 Improvement <strong>of</strong><br />

somatic embryogenesis in wild cherry (Prunus avium), effect <strong>of</strong><br />

maltose and ABA supplements. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult.<br />

55, 199-209.<br />

REPUNTE V.P., TAYA M. & TONE S. 1995 Preparation <strong>of</strong><br />

artificial seeds using cell aggregates from horseradish hairy roots<br />

encapsulated in alginate gel with paraffin coat. J. Ferment.<br />

Bioeng. 79, 83-86.<br />

RIER J.P. & BESLOW D.T. 1967 Sucrose concentration and the<br />

differentiation <strong>of</strong> xylem in callus. Bot. Gaz. 128, 73-77.<br />

RIER J.P. & CHEN P.K. 1964 Pigment induction in plant tissue<br />

cultures. <strong>Plant</strong> Physiol. 39, Suppl.<br />

RIERA M., VALON C., FENZI F., GIRAUDAT J. & LEUNG J.<br />

2005 <strong>The</strong> genetics <strong>of</strong> adaptive responses to drought stress:<br />

abscisic acid-dependent and abscisic acid-independent signalling<br />

components. Physiol. <strong>Plant</strong>. 123, 111-119.<br />

ROBBINS W.J. & BARTLEY M.A. 1937 Vitamin B, and the<br />

growth <strong>of</strong> excised tomato roots. Science 85, 246-247.<br />

ROBBINS W.J. & SCHMIDT M.B. 1939a Vitamin B6, a growth<br />

substance for isolated tomato roots. Proc. Nat. Acad. Sci. Wash.<br />

25, 1-3.<br />

ROBBINS W.J. & SCHMIDT M.B. 1939b Further experiments on<br />

excised tomato roots. Am. J. Bot. 26, 149-159.<br />

ROBBINS W.J. 1922 Cultivation <strong>of</strong> excised root tips and stem tips<br />

under sterile conditions. Bot. Gaz. 73, 376-390.<br />

ROBERTS A.V. & SMITH E.F. 1990 <strong>The</strong> preparation in vitro <strong>of</strong><br />

chrysanthemum for transplantation to soil. (1). Protection <strong>of</strong> roots<br />

by cellulose plugs. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 21, 129-32.<br />

ROBERTS D.R., SUTTON B.C.S. & FLINN B.S. 1990<br />

Synchronous and high frequency germination <strong>of</strong> interior spruce<br />

somatic embryos following partial drying at high relative<br />

humidity. Can. J. Bot. 68, 1086-1090.<br />

ROBERTS-OEHLSCHLAGER, S.L. & DUNWELL J.M. 1990<br />

Barley anther culture – pretreatment on mannitol stimulates<br />

production <strong>of</strong> microspore-derived embryos. <strong>Plant</strong> Cell <strong>Tissue</strong><br />

Organ Cult. 20, 235-240.<br />

RODRIGUEZ G. & LORENZO MARTIN J.R. 1987 In vitro<br />

propagation <strong>of</strong> Canary Island banana (Musa acuminata Colla<br />

AAA var. Dwarf Cavendish). Studies <strong>of</strong> factors affecting culture<br />

obtention, preservation and conformity <strong>of</strong> the plants. Acta Hortic.<br />

212, 577-583.<br />

ROEST S. & BOKELMANN G.S. 1975 Vegetative propagation <strong>of</strong><br />

Chrysanthemum morifolium Ram. in vitro. Sci. Hortic. 3, 317-330.<br />

ROMBERGER J.A. & TABOR C.A. 1971 <strong>The</strong> Picea abies shoot<br />

apical meristem in culture. I. Agar and autoclaving effects. Am.<br />

J. Bot. 58, 131-140.<br />

ROSE D. & MARTIN S.M. 1975 Effect <strong>of</strong> ammonium on growth<br />

<strong>of</strong> plant cells (Ipomoea sp.) in suspension cultures. Can. J. Bot.<br />

53, 1942-1949.<br />

ROSNER A., GRESSEL J. & JAKOB K.M. 1977 Discoordination<br />

<strong>of</strong> ribosomal RNA metabolism during metabolic shifts <strong>of</strong><br />

Spirodela plants. Biochim. Biophys. Acta 474, 386-397.<br />

RUBERY P.H. 1980 <strong>The</strong> mechanism <strong>of</strong> transmembrane auxin<br />

transport and its relation to the chemiosmotic hypothesis <strong>of</strong> the


polar transport <strong>of</strong> auxin. pp. 50-60 in Skoog F. (ed.) 1980 <strong>Plant</strong><br />

Growth Substances. Proc. 10 th Int. Cong. <strong>Plant</strong> Growth<br />

Substances, Madison, Wisconsin 1979. Springer Verlag, Berlin,<br />

Heidelberg.<br />

RUGINI E., TARINI P. & ROSSODIVITA M.E. 1987 Control <strong>of</strong><br />

shoot vitrification <strong>of</strong> almond and olive grown in vitro. Acta<br />

Hortic. 212, 177-183.<br />

RUSSELL J.A. & McCOWN B.H. 1986 <strong>Culture</strong> and regeneration<br />

<strong>of</strong> Populus leaf protoplasts isolated from non-seedling tissue.<br />

<strong>Plant</strong> Sci. 46, 133-142.<br />

SAALBACH G. & KOBLITZ H. 1978 Attempts to initiate callus<br />

formation from barley leaves. <strong>Plant</strong> Sci. Lett. 13, 165-169.<br />

SACHAR R.C. & IYER R.D. 1959 Effect <strong>of</strong> auxin, kinetin and<br />

gibberellin on the placental tissue <strong>of</strong> Opuntia dillenii Haw.<br />

cultured in vitro. Phytomorph. 9, 1-3.<br />

SADASIVAN V. 1951 <strong>The</strong> phosphatases in coconut (Cocos<br />

nucifera). Arch. Biochem. 30, 159-164.<br />

SAGARE A.P., LEE Y.L., LIN T.C., CHEN C.C. & TSAY H.S.<br />

2000 Cytokinin-induced somatic embryogenesis and plant<br />

regeneration in Corydalis yanhusuo (Fumariaceae) -a medicinal<br />

plant. <strong>Plant</strong> Sci. 160,139-147.<br />

SAGAWA Y. & KUNISAKI J.T. 1982 Clonal propagation <strong>of</strong><br />

orchids by tissue culture. pp. 683-684 in Fujiwara A. (ed.) 1982<br />

(q.v.).<br />

SAVAGE A.D., KING J. & GAMBORG O.L. 1979 Recovery <strong>of</strong> a<br />

pantothenate auxotroph from a cell suspension culture <strong>of</strong> Datura<br />

innoxia Mill. <strong>Plant</strong> Sci. Lett. 16, 367-376.<br />

SCHENK R.U. & HILDEBRANDT A.C. 1972 Medium and<br />

techniques for induction and growth <strong>of</strong> monocotyledonous and<br />

dicotyledonous plant cell cultures. Can. J. Bot. 50, 199-204.<br />

SCHERER P.A. 1988 Standardization <strong>of</strong> plant micropropagation<br />

by usage <strong>of</strong> a liquid medium with polyurethane foam plugs or a<br />

solidified medium with the gellan gum gelrite instead <strong>of</strong> agar.<br />

Acta Hortic. 226, 107-114.<br />

SCHERER P.A., MULLER E., LIPPERT H. & WOLFF G. 1988<br />

Multielement analysis <strong>of</strong> agar and gelrite impurities investigated<br />

by inductively coupled plasma emission spectrometry as well as<br />

physical properties <strong>of</strong> tissue culture media prepared with agar or<br />

gellan gum gelrite. Acta Hortic. 226, 655-658.<br />

SCHOLTEN H.J. & PIERIK R.L.M. 1998 Agar as a gelling agent:<br />

differential biological effects in vitro. Sci. Hortic. 77, 109-116.<br />

SCHUBERT S. & MATZKE H. 1985 Influence <strong>of</strong> phytohormones<br />

and other effectors on proton extrusion by isolated protoplasts<br />

from rape leaves. Physiol. <strong>Plant</strong>. 64, 285-289.<br />

SCHULLER A. & REUTHER G. 1993 Response <strong>of</strong> Abies alba<br />

embryonal-suspensor mass to various carbohydrate treatments.<br />

<strong>Plant</strong> Cell Rep. 12, 199-202.<br />

SCOTT J. & BREEN P. 1988 Sugar uptake by strawberry fruit<br />

discs and protoplasts. HortScience 23, 756.<br />

SEARS R.G. & DECKARD E.L. 1982 <strong>Tissue</strong> culture variability<br />

in wheat: callus induction and plant regeneration. Crop Sci. 22,<br />

546-550.<br />

SENARATNA T., MCKERSEE B.D. & BOWLEY S.R. 1989a<br />

Desiccation tolerance <strong>of</strong> alfalfa (Medicago sativa L.) somatic<br />

embryos. Influence <strong>of</strong> abscisic acid, stress pretreatments and<br />

drying rates. <strong>Plant</strong> Sci. 65, 253-259.<br />

SENARATNA T., McKERSIE B. & ECCLESTONE S. 1989b<br />

Germination <strong>of</strong> desiccated somatic embryos <strong>of</strong> alfalfa (Medicago<br />

sativa L.). <strong>Plant</strong> Physiol. 89, Suppl., 135.<br />

SERP D., CANTANA E., HEINZEN C., VON STOCKAR U. &<br />

MARISON I.W. 2000 Characterization <strong>of</strong> an encapsulation<br />

device for the production <strong>of</strong> monodisperse alginate beads for cell<br />

immobilization. Biotechnol. Bioeng. 70, 41-53.<br />

SERRES R.A. 1988 Effects <strong>of</strong> sucrose and basal medium<br />

concentration on in vitro rooting <strong>of</strong> American chestnut.<br />

HortScience 23, 739.<br />

Chapter 4 169<br />

SHANTZ E.M. & STEWARD F.C. 1952 Coconut milk factor: the<br />

growth-promoting substances in coconut milk. J. Am. Chem.<br />

Soc. 74, 6133.<br />

SHANTZ E.M. & STEWARD F.C. 1955 <strong>The</strong> identification <strong>of</strong><br />

compound A from coconut milk as 1,3-diphenylurea. J. Am.<br />

Chem. Soc. 77, 6351-6353.<br />

SHANTZ E.M. & STEWARD F.C. 1956 <strong>The</strong> general nature <strong>of</strong><br />

some nitrogen free growth-promoting substances from Aesculus<br />

and Cocos. <strong>Plant</strong> Physiol. 30, Suppl., XXXV.<br />

SHANTZ E.M. & STEWARD F.C. 1964 Growth-promoting<br />

substances from the environment <strong>of</strong> the embryo. <strong>II</strong>. <strong>The</strong> growthstimulating<br />

complexes <strong>of</strong> coconut milk, corn and Aesculus. pp.<br />

59-75 in Actes du Coll. Int. C.N.R.S. No. 123.<br />

SHARMA A.K., PRASAD R.N. & CHATURVEDI H.C. 1981<br />

Clonal propagation <strong>of</strong> Bougainvillea glabra `Magnifica’ through<br />

shoot apex culture. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 1, 33-38.<br />

SHEAT D.E.G., FLETCHER B.H. & STREET H.E. 1959 Studies<br />

on the growth <strong>of</strong> excised roots. V<strong>II</strong>I. <strong>The</strong> growth <strong>of</strong> excised<br />

tomato roots supplied with various inorganic sources <strong>of</strong> nitrogen.<br />

New Phytol. 58, 128-141.<br />

SHEPARD J.F. & TOTTEN R.E. 1977 Mesophyll cell protoplasts<br />

<strong>of</strong> potato: isolation, proliferation and plant regeneration. <strong>Plant</strong><br />

Physiol. 60, 313-316.<br />

SHILLITO R.D., PASZKOWSKI J. & POTRYKUS I. 1983<br />

Agarose plating and a bead type culture technique enable and<br />

stimulate development <strong>of</strong> protoplast-derived colonies in a number<br />

<strong>of</strong> plant species. <strong>Plant</strong> Cell Rep. 2, 244-247.<br />

SHININGER T.L. 1979 <strong>The</strong> control <strong>of</strong> vascular development.<br />

Annu. Rev. <strong>Plant</strong> Physiol. 30, 313-337.<br />

SHOLTO-DOUGLAS J. 1976 Advanced Guide to Hydroponics.<br />

Pelham Books, London. ISBN 0-7207-08303.<br />

SHORT K.C., WARBURTON J. & ROBERT A.V. 1987 In vitro<br />

hardening <strong>of</strong> cultured cauliflower and chrysanthemum plantlets<br />

to humidity. Acta Hortic. 212, 329-334.<br />

SHVETSOV S.G. & GAMBURG K.Z. 1981 Uptake <strong>of</strong> 2,4-D by<br />

corn cells in a suspension culture. Dokl. Akad. Nauk S.S.S.R.<br />

257, 765-768.<br />

SIEVERT R.C. & HILDEBRANDT A.C. 1965 Variations within<br />

single cell clones <strong>of</strong> tobacco tissue cultures. Am. J. Bot. 52,<br />

742-750.<br />

SINGHA S. 1982 Influence <strong>of</strong> agar concentration on in vitro shoot<br />

proliferation <strong>of</strong> Malus sp. `Almey’ and Pyrus communis `Seckel’.<br />

J. Am. Soc. Hortic. Sci. 107, 657-660.<br />

SKINNER J.C. & STREET H.E. 1954 Studies on the growth <strong>of</strong><br />

excised roots. <strong>II</strong>. Observations on the growth <strong>of</strong> excised<br />

groundsel roots. New Phytol. 53, 44-67.<br />

SKIRVIN R.M. 1981 Fruit crops. pp. 51-139 in Conger B.V. (ed.)<br />

Cloning Agricultural <strong>Plant</strong>s via In Vitro Techniques. CRC Press,<br />

Inc., Boca Raton, Florida.<br />

SKIRVIN R.M. & CHU M.C 1979 In vitro propagation <strong>of</strong><br />

‘Forever Yours’ rose (Rosa hybrida) HortScience 14, 608-610.<br />

SKIRVIN R.M., CHU M.C., MANN M.L., YOUNG H.,<br />

SULLIVAN J. & FERMANIAN T. 1986 Stability <strong>of</strong> tissue<br />

culture medium pH as a function <strong>of</strong> autoclaving, time, and<br />

cultured plant material. <strong>Plant</strong> Cell Rep. 5, 292-294.<br />

SMITH C.W. 1967 A study <strong>of</strong> the growth <strong>of</strong> excised embryo shoot<br />

apices <strong>of</strong> wheat in vitro. Ann. Bot. 31, 593-605.<br />

SMITH D.L. & KRIKORIAN A.D. 1989 Release <strong>of</strong> somatic<br />

embryogenic potential from excised zygotic embryos <strong>of</strong> carrot<br />

and maintenance <strong>of</strong> polyembryonic cultures in hormone-free<br />

medium. Am. J. Bot. 76, 1832-1843.<br />

SMITH D.L. & KRIKORIAN A.D. 1990a Somatic proembryo<br />

production from excised wounded zygotic carrot embryos on<br />

hormone-free medium; evaluation <strong>of</strong> the effects <strong>of</strong> pH, ethylene<br />

and activated charcoal. <strong>Plant</strong> Cell Rep. 9, 34-37.


170 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

SMITH D.L. & KRIKORIAN A.D. 1990b pH control <strong>of</strong> carrot<br />

somatic embryogenesis. pp. 449-453 in Nijkamp et al. (eds.)<br />

1990 (q.v.).<br />

SMITH E.F., ROBERTS A.V. & MOTTLEY J. 1990a <strong>The</strong><br />

preparation in vitro <strong>of</strong> chrysanthemum for transplantation to soil.<br />

(2) Improved resistance to desiccation conferred by<br />

paclobutrazol. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult., 21, 133-40.<br />

SMITH E.F., ROBERTS A.V. & MOTTLEY J. 1990b <strong>The</strong><br />

preparation in vitro <strong>of</strong> chrysanthemum for transplantation to soil.<br />

(3) Improved resistance to desiccation conferred by reduced<br />

humidity. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. , 21, 141-5.<br />

SMITH F.A. & RAVEN J.A. 1979 Intercellular pH and its<br />

regulation. Annu. Rev. <strong>Plant</strong> Physiol. 30, 289-311.<br />

SMITH M.M. & STONE B.A. 1973 Studies on Lolium multiflorum<br />

endosperm in tissue culture. I. Nutrition. Aust. J. Biol. 26, 123-133.<br />

SOCZEK U. & HEMPEL M. 1988 <strong>The</strong> influence <strong>of</strong> some organic<br />

medium compounds on multiplication <strong>of</strong> gerbera in vitro. Acta<br />

Hortic. 226, 643-646.<br />

SOPORY S.K., JACOBSEN E. & WENZEL G. 1978 Production<br />

<strong>of</strong> monohaploid embryoids and plantlets in cultured anthers <strong>of</strong><br />

Solanum tuberosum. <strong>Plant</strong> Sci. Lett. 12, 47-54.<br />

SORVARI S. 1986a <strong>The</strong> effect <strong>of</strong> starch gelatinized nutrient media<br />

in barley anther cultures. Ann. Agric. Fenn. 25, 127-133.<br />

SORVARI S. 1986b Differentiation <strong>of</strong> potato discs in barley starch<br />

gelatinized nutrient media. Ann. Agric. Fenn. 25, 135-138.<br />

SORVARI S. 1986c Comparison <strong>of</strong> anther cultures <strong>of</strong> barley<br />

cultivars in barley-starch and agar gelatinized media. Ann. Agric.<br />

Fenn. 25, 249-254.<br />

SPANGENBERG G., KOOP H.-U., LICHTER R. & SCHWEI-<br />

GER H.G. 1986 Microculture <strong>of</strong> single protoplasts <strong>of</strong> Brassica<br />

napus. Physiol. <strong>Plant</strong>. 66, 1-8.<br />

STAFFORD A. & DAVIES D.R. 1979 <strong>The</strong> culture <strong>of</strong> immature<br />

pea embryos. Ann. Bot. 44, 315-321.<br />

STASOLLA C. & YEUNG E.C. 1999 Ascorbic acid improves<br />

conversion <strong>of</strong> white spruce somatic embryos. In Vitro Cell.<br />

Dev.-Pl. 35, 316-319.<br />

STASOLLA C., KONG L., YEUNG E.C. & THORPE T.A. 2003<br />

Maturation <strong>of</strong> somatic embryos in conifers: morphogenesis,<br />

physiology, biochemistry and molecular biology. In Vitro Cell.<br />

Dev. – Pl. 38, 93-105.<br />

STAUDT G. 1984 <strong>The</strong> effect <strong>of</strong> myo-inositol on the growth <strong>of</strong><br />

callus tissue in Vitis. J. <strong>Plant</strong> Physiol. 116, 161-166.<br />

STEFFEN J.D., SACHS R.M. & HACKETT W.P. 1988 Growth<br />

and development <strong>of</strong> reproductive and vegetative tissues <strong>of</strong><br />

Bougainvillea cultured in vitro as a function <strong>of</strong> carbohydrate.<br />

Am. J. Bot. 75, 1219-1224.<br />

STEHSEL M.L. & CAPLIN S.M. 1969 Sugars: autoclaving versus<br />

sterile filtration and the growth <strong>of</strong> carrot root tissue. Life Sci. 8,<br />

1255-1259.<br />

STEINER S. & LESTER R.L. 1972 Studies on the diversity <strong>of</strong><br />

inositol-containing yeast phospholipids: incorporation <strong>of</strong> 2deoxyglucose<br />

into lipid. J. Bacteriol. 109, 81-88.<br />

STEINER S., SMITH S., WAECHTER C.J. & LESTER R.L. 1969<br />

Isolation and partial characterization <strong>of</strong> a major inositolcontaining<br />

yeast phospholipid in baker’s yeast, mannosyldiinositol,<br />

diphosphorylceramide. Proc. Natl. Acad. Sci. USA 64,<br />

1042-1048.<br />

STEINITZ B. (1999) Sugar alcohols display nonosmotic roles in<br />

regulating morphogenesis and metabolism in plants that do not<br />

produce polyols as primary photosynthetic products. J. <strong>Plant</strong><br />

Physiol. 155, 1-8.<br />

STEWARD F.C. & CAPLIN S.M. 1951 A tissue culture from<br />

potato tuber: the synergistic action <strong>of</strong> 2,4-D and <strong>of</strong> coconut milk.<br />

Science 113, 518-520.<br />

STEWARD F.C. & CAPLIN S.M. 1952 Investigations on growth<br />

and metabolism <strong>of</strong> plant cells. IV. Evidence on the role <strong>of</strong> the<br />

coconut milk factor. Ann. Bot. 16, 491-504.<br />

STEWARD F.C. & MOHAN RAM H.Y. 1961 Determining<br />

factors in cell growth: some implications for morphogenesis in<br />

plants. Adv. Morph. 1, 189-265.<br />

STEWARD F.C. & RAO K.V.N. 1970 Investigations on the growth<br />

and metabolism <strong>of</strong> cultured explants <strong>of</strong> Daucus carota. <strong>II</strong>I. <strong>The</strong><br />

range <strong>of</strong> responses induced in carrot explants by exogenous growth<br />

factors and by trace elements. <strong>Plant</strong>a 91, 129-145.<br />

STEWARD F.C. & SHANTZ E.M. 1956 <strong>The</strong> chemical induction<br />

<strong>of</strong> growth in plant tissue cultures. pp. 165-186 in Wain and<br />

Wightman (eds.) 1956 <strong>The</strong> Chemistry and Mode <strong>of</strong> Action <strong>of</strong><br />

<strong>Plant</strong> Growth Substances. Butterworth Scientific Publications<br />

Ltd., London.<br />

STEWARD F.C. & SHANTZ E.M. 1959 <strong>The</strong> chemical regulation<br />

<strong>of</strong> growth: Some substances and extracts which induce growth<br />

and morphogenesis. Annu. Rev. <strong>Plant</strong> Physiol. 10, 379-404.<br />

STEWARD F.C. 1963 Carrots and coconuts: Some investigations<br />

on growth. pp. 178-197 in Maheshwari and Ranga Swamy (eds.)<br />

1963 (q.v.).<br />

STEWARD F.C., CAPLIN S.M. & MILLAR F.K. 1952<br />

Investigations on growth and metabolism <strong>of</strong> plant cells. I. New<br />

techniques for the investigation <strong>of</strong> metabolism, nutrition, and<br />

growth in undifferentiated cells. Ann. Bot. 16, 57-77.<br />

STEWARD F.C., MAPES M.O. & AMMIRATO P.V. 1969<br />

Growth and morphogenesis in tissue and free cell cultures. pp.<br />

329-376 in Steward F.C. (ed.) 1969 <strong>Plant</strong> Physiology - A Treatise<br />

Vol. VB. Academic Press, New York, London.<br />

STEWARD F.C., MAPES M.O., KENT A.E. & HOLSTEN R.D.<br />

1964 Growth and development <strong>of</strong> cultured plant cells. Science<br />

143, 20-27.<br />

STEWARD F.C., SHANTZ E.M., POLLARD J.K., MAPES M.O.<br />

& MITRA J. 1961 Growth induction in explanted cells and<br />

tissues: Metabolic and morphogenetic manifestations. pp. 193-<br />

246 in Rudnick D. (ed.) Synthesis <strong>of</strong> Molecular and Cellular<br />

Structure. Ronald Press, New York.<br />

STONE O.M. 1963 Factors affecting the growth <strong>of</strong> carnation<br />

plants from shoot apices. Ann. Appl. Biol. 52, 199-209.<br />

STRAUS J. & LA RUE C.D. 1954 Maize endosperm tissue grown<br />

in vitro. 1. <strong>Culture</strong> requirements. Am. J. Bot. 41, 687-694.<br />

STREET H.E. & HENSHAW G.G. 1966 Introduction and methods<br />

employed in plant tissue culture. pp. 459-532 in Willmer E.N.<br />

Cells and <strong>Tissue</strong>s in <strong>Culture</strong> – Methods and Physiology. Vol 3.<br />

Academic Press, London, New York].<br />

STREET H.E. & McGREGOR S.M. 1952 <strong>The</strong> carbohydrate<br />

nutrition <strong>of</strong> tomato roots. <strong>II</strong>I. <strong>The</strong> effects <strong>of</strong> external sucrose<br />

concentration on the growth and anatomy <strong>of</strong> excised roots. Ann.<br />

Bot. 16, 185-205.<br />

STREET H.E. 1969 Growth in organised and unorganised systems<br />

- knowledge gained by culture <strong>of</strong> organs and tissue explants. pp.<br />

3-224 in Steward F. C. (ed.) 1969. <strong>Plant</strong> Physiology - a Treatise.<br />

5B. Academic Press, New York.<br />

STREET H.E. 1977 Laboratory organization. pp. 11-30 in Street<br />

H.E. (ed.) <strong>Plant</strong> <strong>Tissue</strong> and Cell <strong>Culture</strong>. Bot. Monographs<br />

Vol.11, Blackwell Scientific Public-ations. Oxford, London.<br />

STREET H.E., McGONAGLE M.P. & LOWE J.S. 1951<br />

Observations on the `staling’ <strong>of</strong> White’s medium by excised<br />

tomato roots. Physiol. <strong>Plant</strong>. 4, 592-616.<br />

STREET H.E., McGONAGLE M.P. & McGREGOR S.M. 1952<br />

Observations on the ‘staling’ <strong>of</strong> White’s medium by excised<br />

tomato roots. <strong>II</strong>. Iron availability. Physiol. <strong>Plant</strong>. 5, 248-276.<br />

STRICKLAND S.G., NICHOL J.W., McCALL C.M. & STUART<br />

D.A. 1987 Effect <strong>of</strong> carbohydrate source on alfalfa<br />

embryogenesis. <strong>Plant</strong> Sci. 48, 113-121.


STUART D.A., STRICKLAND S.G. & NICHOL J.W. 1986<br />

Enhanced somatic embryogenesis using maltose. U.S. Patent No.<br />

4801545 ,<br />

STUART D.A., STRICKLAND S.G. & WALKER K.A. 1987<br />

Bioreactor production <strong>of</strong> alfalfa somatic embryos. HortScience<br />

22, 800-803.<br />

SUNDERLAND N. & DUNWELL J.M. 1977 Anther and pollen<br />

culture. pp. 223-265 in Street H.E. (ed.) <strong>Plant</strong> <strong>Tissue</strong> and Cell<br />

<strong>Culture</strong>. Bot. Monographs Vol.11, Blackwell Scientific Publications.<br />

Oxford, London.<br />

SUTTLE J.C. & HULTSTRAND J.F. 1994 Role <strong>of</strong> endogenous<br />

abscisic acid in potato microtuber dormancy. <strong>Plant</strong> Physiol. 105,<br />

891-896.<br />

SWEDLUND B. & LOCY R.D. 1988 Somatic embryogenesis and<br />

plant regeneration in two-year old cultures <strong>of</strong> Zea diploperennis.<br />

<strong>Plant</strong> Cell Rep. 7, 144-147.<br />

TABER R.P., ZHANG C. & HU W.S. 1998 Kinetics <strong>of</strong> Douglasfir<br />

(Pseudotsuga menziesii) somatic embryo development. Can.<br />

J. Bot. 76, 863-871.<br />

TAKAYAMA S. & MISAWA M. 1979 Differentiation in Lilium<br />

bulb scales grown in vitro - effect <strong>of</strong> various cultural conditions.<br />

Physiol. <strong>Plant</strong>. 46, 184-190.<br />

TAKAYAMA S. & MISAWA M. 1979 Differentiation in Lilium<br />

bulb scales grown in vitro - effect <strong>of</strong> various cultural conditions.<br />

Physiol. <strong>Plant</strong>. 46, 184-190.<br />

TANAKA M., HIRANO T., GOI M., HIGASHIURA T.,<br />

SASAHARA H. & MURASAKI K. 1991 Practical application <strong>of</strong><br />

a novel disposable film culture vessel in micropropagation. Acta<br />

Hortic. 300, 77-84.<br />

TANDEAU DE MARSAC N. & PEAUD-LENOEL C. 1972a<br />

<strong>Culture</strong>s photosynthtique de lignes clonales de cellules de tabac.<br />

Compt. Rend. Acad. Sci. Paris 274D, 1800-1802.<br />

TANDEAU DE MARSAC N. & PEAUD-LENOEL C. 1972b<br />

Exchanges d’oxygène et assimilation de gaz carbonique de tissus<br />

de tabac en culture photosynthétique. Compt. Rend. Acad. Sci.<br />

Paris 274D, 2310-2313.<br />

TAUTORUS T.E. & DUNSTAN D.I. 1995 Scale-up <strong>of</strong> embryonic<br />

plant suspension cultures in bioreactors. In: S Jain, P Gupta, and<br />

R Newton (Eds) Somatic embryogenesis in woody plants Vol. 1.<br />

Kluwer Academic Publishers, <strong>The</strong> Netherlands pp 265-292.<br />

TEISSON C. & ALVARD D. 1999 In vitro production <strong>of</strong> potato<br />

microtubers in liquid medium using temporary immersion. Potato<br />

Res. 42, 499-504.<br />

TELLE J. & GAUTHERET R.J. 1947 Sur la culture indéfinie des<br />

tissus de la racine de jusquiame (Hyascyamus niger). Compt.<br />

Rend. Acad. Sci. Paris 224, 1653-1654.<br />

TENG W.L. 1997 Regeneration <strong>of</strong> Anthurium adventitious shoots<br />

using liquid or raft culture. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 49,<br />

153-156.<br />

THOM M., MARETZKI A., KOMOR E. & SAKAI W.S. 1981<br />

Nutrient uptake and accumulation by sugarcane cell cultures in<br />

relation to the growth cycle. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 1, 3-14.<br />

THOMPSON M.R. & THORPE T.A. 1981 Mannitol metabolism<br />

in cell cultures. <strong>Plant</strong> Physiol. 67, Suppl., 27.<br />

THOMPSON M.R., DOUGLAS T.J., OBATA-SASAMOTO H. &<br />

THORPE T.A. 1986 Mannitol metabolism in cultured plant<br />

cells. Physiol. <strong>Plant</strong>. 67, 365-369.<br />

THORPE T.A & MEIER D.D. 1973 Sucrose metabolism during<br />

tobacco callus growth. Phytochemistry 12, 493-497.<br />

THORPE T.A. 1982 Carbohydrate utilization and metabolism.<br />

pp. 325-368 in Bonga J.M. & Durzan D.J. (eds.) <strong>Tissue</strong> <strong>Culture</strong><br />

in Forestry. Martinus Nijh<strong>of</strong>f/Dr. W. Junk Publishers, <strong>The</strong><br />

Hague.<br />

THORPE T.A. & LAISHLEY E.J. 1974 Carbohydrate oxidation<br />

during Nicotiana tabacum callus growth. Phytochemistry 13,<br />

1323-1328.<br />

Chapter 4 171<br />

THORPE T.A. & LAISHLEY E.J. 1973 Glucose oxidation<br />

during shoot initiation in tobacco callus cultures. J. Exp. Bot.<br />

24, 1082-1089.<br />

THORPE T.A. & MEIER D.D. 1972 Starch metabolism, repiration<br />

and shoot formation in tobacco callus cultures. Physiol. <strong>Plant</strong>. 27,<br />

365-369.<br />

THORPE T.A. & MEIER D.D. 1973 Effects <strong>of</strong> gibberellic acid<br />

and abscisic acid on shoot formation in tobacco callus cultures.<br />

Physiol. <strong>Plant</strong>. 29, 121-124.<br />

THORPE T.A. & MEIER D.D. 1974 Starch metabolism in shootforming<br />

tobacco callus. J. Exp. Bot. 25, 288-294.<br />

THORPE T.A. & MEIER D.D. 1975 Effect <strong>of</strong> gibberellic acid on<br />

starch metabolism in tobacco callus cultured under shoot-forming<br />

conditions. Phytomorph. 25, 238-245.<br />

THORPE T.A. & MURASHIGE T. 1968a Some histochemical<br />

changes underlying shoot initiation in tobacco callus culture. Am.<br />

J. Bot. 55, 710.<br />

THORPE T.A. & MURASHIGE T. 1968b Starch accumulation in<br />

shoot-forming tobacco callus cultures. Science 160, 421-422.<br />

THORPE T.A. & MURASHIGE T. 1970 Some histochemical<br />

changes underlying shoot initiation in tobacco callus cultures.<br />

Can. J. Bot. 48, 277-285.<br />

THORPE T.A. 1982 Physiological and biochemical aspects <strong>of</strong><br />

organogenesis in vitro. pp. 121-124 in Fujiwara A. (ed.) 1982<br />

(q.v.).<br />

THORPE T.A., JOY R.W. IV & LEUNG W.M. 1986 Starch turnover<br />

in shoot-forming tobacco callus. Physiol. <strong>Plant</strong>. 66, 58-62.<br />

TIAN H.Q. & RUSSELL S.D. 1998 <strong>Culture</strong>-induced changes in<br />

osmolality <strong>of</strong> tobacco cell suspensions using four exogenous<br />

sugars. <strong>Plant</strong> Cell Tiss Organ Cult. 55, 9-13<br />

TIBURCIO A.F., GENDY C.A. & TRAN THANH VAN K. 1989<br />

Morphogenesis in tobacco subepidermal cells: putrescine as a<br />

marker <strong>of</strong> root differentiation. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 19,<br />

43-54.<br />

TIMBERT R., BARBOTIN J.-N., KERSULEC A., BAZINET C.<br />

& THOMAS D. 1995 Physico-chemical properties <strong>of</strong> the<br />

encapsulation matrix and germination <strong>of</strong> carrot somatic embryos.<br />

Biotechnol. Bioeng. 46, 573-578.<br />

TOMAR U.K. & GUPTA S.C. 1988 Somatic embryogenesis and<br />

organogenesis in callus cultures <strong>of</strong> a tree legume - Albizia<br />

richardiana King. <strong>Plant</strong> Cell Rep. 7, 70-73.<br />

TOR M., TWYFORD C.T., FUNES I., BOCCON-GIBOD J.,<br />

AINSWORTH C.C. & MANTELL S.H. 1998 Isolation and<br />

culture <strong>of</strong> protoplasts from immature leaves and embryogenic<br />

cell suspensions <strong>of</strong> Dioscorea yams: tools for transient gene<br />

expression studies. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 53, 113-125.<br />

TORRES K. & CARLISI J.A. 1986 Shoot and root organogenesis<br />

<strong>of</strong> Camellia sasanqua. <strong>Plant</strong> Cell Rep. 5, 381-384.<br />

TORREY J.G. 1956 Chemical factors limiting lateral root<br />

formation in isolated pea roots. Physiol. <strong>Plant</strong>. 9, 370-388<br />

TRAN THANH VAN K. & TRINH H. 1978 <strong>Plant</strong> propagation:<br />

non-identical and identical copies. pp. 134-158 in Hughes K.W.,<br />

Henke R. & Constantin M. (eds.) Propagation <strong>of</strong> Higher <strong>Plant</strong>s<br />

through <strong>Tissue</strong> <strong>Culture</strong>. Technical Inf. Center, U.S. Dept.<br />

Energy. Spingfield, Va. CONF-7804111.<br />

TRELEASE S.F. & TRELEASE H.M. 1933 Physiologically<br />

balanced culture solutions with stable hydrogen-ion concentration.<br />

Science 78, 438-439.<br />

TREMBLAY F.M. & LALONDE M. 1984 Requirements for in<br />

vitro propagation <strong>of</strong> seven nitrogen-fixing Alnus species. <strong>Plant</strong><br />

Cell <strong>Tissue</strong> Organ Cult. 3, 189-199.<br />

TREMBLAY F.M., NESME X. & LALONDE M. 1984 Selection<br />

and micropropagation <strong>of</strong> nodulating and non-nodulating clones <strong>of</strong><br />

Alnus crispa (Ait.) Pursh. <strong>Plant</strong> Soil 78, 171-179.<br />

TRINDADE H. & PAIS M.S. 1997 In vitro studies on Eucalyptus<br />

globulus rooting ability. In Vitro Cell. Dev.-Pl. 33, 1-5.


172 <strong>The</strong> <strong>Components</strong> <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong> <strong>Media</strong> <strong>II</strong><br />

TROLINDER N.L. & GOODIN J.R. 1987 Somatic embryogenesis<br />

and plant regeneration in cotton (Gossypium hirsutum L.). <strong>Plant</strong><br />

Cell Rep. 6, 231-234.<br />

TULECKE W., WEINSTEIN L.H., RUTNER A. &<br />

LAURENCOT H.J. 1961 <strong>The</strong> biochemical composition <strong>of</strong><br />

coconut water as related to its use in plant tissue culture. Contrib.<br />

Boyce Thomps. 21, 115-128.<br />

TURNER S.R. & SINGHA S. 1990 Vitrification <strong>of</strong> crabapple, pear<br />

and geum on gellan gum-solidified culture medium. HortScience<br />

25, 1648-1650.<br />

UMBECK P., JOHNSON G., BARTON K. & SWAIN W. 1987<br />

Genetically transformed cotton (Gossypium hirsutum L.) plants.<br />

Nat. Biotechnol. 5, 263-266.<br />

VACIN E.F. & WENT F.W. 1949 Some pH changes in nutrient<br />

solutions. Bot. Gaz. 110, 605-613.<br />

VAIN P., MCMULLEN M.D. & FINER J.J. 1993 Osmotic<br />

treatment enhances particle bombardment-mediated transient and<br />

stable transformation <strong>of</strong> maize. <strong>Plant</strong> Cell Rep. 12, 84-88.<br />

VAN DER KRIEKEN W.M., BRETELER H., VISSER M.H.M. &<br />

JORDI W. 1992 Effect <strong>of</strong> light and rib<strong>of</strong>lavin on indolebutyric<br />

acid-induced root formation on apply in vitro. Physiol. <strong>Plant</strong>. 85,<br />

589-594.<br />

VAN HUYSTEE R.B. 1977 Porphyrin metabolism in peanut cells<br />

cultured in sucrose containing medium. Acta Hortic. 78, 83-87.<br />

VAN OVERBEEK J. 1942 Water uptake by excised root systems <strong>of</strong><br />

the tomato due to non-osmotic forces. Am. J. Bot. 29, 677-683.<br />

VAN OVERBEEK J., CONKLIN M.E. & BLAKESLEE A.F. 1941<br />

Factors in coconut milk essential for growth and development <strong>of</strong><br />

very young Datura embryos. Science 94, 350-351.<br />

VAN OVERBEEK J., CONKLIN M.E. & BLAKESLEE A.F.<br />

1942 Cultivation in vitro <strong>of</strong> small Datura embryos. Am. J. Bot.<br />

29, 472-477.<br />

VAN OVERBEEK J., SIU R. & HAAGEN-SMIT A.J. 1944<br />

Factors affecting the growth <strong>of</strong> Datura embryos in vitro. Am. J.<br />

Bot. 31, 219-224.<br />

VAN STADEN J. & DREWES S.E. 1975 Identification <strong>of</strong> zeatin<br />

and zeatin riboside in coconut milk. Physiol. <strong>Plant</strong>. 34, 106-109.<br />

VAN STADEN J. 1976 <strong>The</strong> identification <strong>of</strong> zeatin glucoside from<br />

coconut milk. Physiol. <strong>Plant</strong>. 36, 123-126.<br />

VANDENBELT J.M. 1945 Nutritive value <strong>of</strong> coconut. Nature 156,<br />

174-175.<br />

VANSVEREN-VAN ESPEN N. 1973 Effets du saccharose sur le<br />

contenu en chlorophylles de protocormes de Cymbidium Sw.<br />

(Orchidaceae) cultivés in vitro. Bull. Soc. Roy. Bot. Belg. 106,<br />

107-115.<br />

VASIL I.K. & HILDEBRANDT A.C. 1966a Variations <strong>of</strong><br />

morphogenetic behaviour in plant tissue cultures. I. Cichorium<br />

endivia. Am. J. Bot. 53, 860-869.<br />

VASIL I.K. & HILDEBRANDT A.C. 1966b Variations <strong>of</strong><br />

morphogenetic behaviour in plant tissue cultures. <strong>II</strong>.<br />

Petroselinum hortense. Am. J. Bot. 53, 869-874.<br />

VASIL I.K. & HILDEBRANDT A.C. 1966c Growth and<br />

chlorophyll production in plant callus tissues grown in vitro.<br />

<strong>Plant</strong>a 68, 69-82.<br />

VASIL V. & VASIL I.K. 1981a Somatic embryogenesis and plant<br />

regeneration from tissue cultures <strong>of</strong> Pennisetum americanum, and<br />

P. americanum x P. purpureum hybrid. Am. J. Bot. 68, 864-872.<br />

VASIL V. & VASIL I.K. 1981b Somatic embryogenesis and plant<br />

regeneration from suspension cultures <strong>of</strong> pearl millet<br />

(Pennisetum americanum). Ann. Bot. 47, 669-678.<br />

VENVERLOO C. 1976 <strong>The</strong> formation <strong>of</strong> adventitious organs. <strong>II</strong>I.<br />

A comparison <strong>of</strong> root and shoot formation on Nautilocalyx<br />

explants. Z. Pflanzenphysiol. 80, 310-322.<br />

VERMA D.C. & DOUGALL D.K. 1977 Influence <strong>of</strong> carbohydrates<br />

on quantitative aspects <strong>of</strong> growth and embryo formation in wild<br />

carrot suspension cultures. <strong>Plant</strong> Physiol. 53, 81-85.<br />

VERMA D.C. & DOUGALL D.K. 1979 Biosynthesis <strong>of</strong> myo-<br />

inositol and its role as a precursor <strong>of</strong> cell-wall polysaccharides in<br />

suspension cultures <strong>of</strong> wild-carrot cells. <strong>Plant</strong>a 146, 55-62.<br />

VERMA D.P.S. & VAN HUYSTEE R.B. 1971 Derivation,<br />

characteristics and large scale culture <strong>of</strong> a cell line from Arachis<br />

hypogaea L. cotyledons. Exp. Cell Res. 69, 402-408.<br />

VISEUR J. 1987 Micropropagation <strong>of</strong> pear, Pyrus communis L., in<br />

a double-phase culture medium. Acta Hortic. 212, 117-124.<br />

VON ARNOLD S. 1987 Effect <strong>of</strong> sucrose on starch accumulation<br />

in, and adventitous bud formation on, embryos <strong>of</strong> Picea abies.<br />

Ann. Bot. 59, 15-22.<br />

VREUGDENHIL D., BOOGAARD Y., VISSER, R.G.F. & DE<br />

BRUIJN S.M. 1998 Comparison <strong>of</strong> tuber and shoot formation<br />

from in vitro cultured potato explants. <strong>Plant</strong> Cell Tiss Organ Cult.<br />

53, 197-204.<br />

VYSKOT B. & BEZDEK M. 1984 Stabilization <strong>of</strong> synthetic media<br />

for plant tissue culture and cell cultures. Biol. <strong>Plant</strong>. 26, 132-143.<br />

VYSKOT B. & JARA Z. 1984 Clonal propagation <strong>of</strong> cacti through<br />

axillary buds in vitro. J. Hortic. Sci. 59, 449-452.<br />

WALKER D.R. & PARROTT W.A. 2001 Effect <strong>of</strong> polyethylene<br />

glycol and sugar alcohols on soybean somatic embryo<br />

germination and conversion. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 64,<br />

55-62.<br />

WATAD A.A., AHRONI A., ZUKER A., SHEJTMAN H.,<br />

NISSIM A. & VAINSTEIN A. 1996 Adventitious shoot formation<br />

from carnation stem segments: A comparison <strong>of</strong> different<br />

culture procedures. Sci. Hortic. 65, 313-320<br />

WANN S.R., VEAZEY R. L., KAPHAMMER J. 1997 Activated<br />

charcoal does not catalyze sucrose hydrolysis in tissue culture<br />

media during autoclaving. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 50,<br />

221-224.<br />

WATAD A.A., KOCHBA M., NISSIM A. & GABA V. 1995<br />

Improvement <strong>of</strong> Aconitum napellus micropropagation by liquid<br />

culture on floating membrane rafts. <strong>Plant</strong> Cell Rep. 14, 345-348.<br />

WATANABE K., TANAKA K., ASADA K. & KASAL Z. 1971<br />

<strong>The</strong> growth promoting effect <strong>of</strong> phytic acid on callus tissues <strong>of</strong><br />

rice seed. <strong>Plant</strong> Cell Physiol. 12, 161-164.<br />

WEI Z.M., KYO M. & HARADA H. 1986 Callus formation and<br />

plant-regeneration through direct culture <strong>of</strong> isolated pollen <strong>of</strong><br />

Hordeum vulgare cv. Sabarlis. <strong>The</strong>or. Appl. Genet.72, 252-255.<br />

WELANDER T. 1977 In vitro organogensis in explants from<br />

different cultivars <strong>of</strong> Begonia hiemalis. Physiol. <strong>Plant</strong>. 41,<br />

142-145.<br />

WESTON G.D. & STREET H.E. 1968 Sugar absorption and sucrose<br />

inversion by excised tomato roots. Ann. Bot. 32, 521-529.<br />

WETHERELL D.F. 1969 Phytochrome in cultured wild carrot<br />

tissue. I. Synthesis. <strong>Plant</strong> Physiol. 44, 1734-1737.<br />

WETHERELL D.F. 1984 Enhanced adventive embryogenesis<br />

resulting from plasmolysis <strong>of</strong> cultured wild carrot cells. <strong>Plant</strong><br />

Cell <strong>Tissue</strong> Organ Cult. 3, 221-227.<br />

WETMORE R.H. & RIER J.P. 1963 Experimental induction <strong>of</strong><br />

vascular tissues in callus <strong>of</strong> angiosperms. Am. J. Bot. 50, 418-430.<br />

WHITE M.C., DECKER A.M. & CHANEY R.L. 1981 Metal<br />

complexation in xylem fluid. I. Chemical composition <strong>of</strong> tomato<br />

and soybean stem exudate. <strong>Plant</strong> Physiol. 67, 292-300.<br />

WHITE P.R. 1932 Influence <strong>of</strong> some environmental conditions on<br />

the growth <strong>of</strong> excised root-tips <strong>of</strong> wheat seedlings in liquid<br />

media. <strong>Plant</strong> Physiol. 7, 613-628.<br />

WHITE P.R. 1934 Potentially unlimited growth <strong>of</strong> excised tomato<br />

root tips in a liquid medium. <strong>Plant</strong> Physiol. 9, 585-600.<br />

WHITE P.R. 1937 Survival <strong>of</strong> isolated tomato roots at sub-optimal<br />

and supra-optimal temperatures. <strong>Plant</strong> Physiol. 12, 771-776.<br />

WHITE P.R. 1943a A Handbook <strong>of</strong> <strong>Plant</strong> <strong>Tissue</strong> <strong>Culture</strong>. <strong>The</strong><br />

Jacques Catlell Press, Lancaster, Pa.


WHITE P.R. 1943b Nutrient deficiency studies and an improved<br />

inorganic nutrient for cultivation <strong>of</strong> excised tomato roots. Growth<br />

7, 53-65.<br />

WHITE P.R. 1954 <strong>The</strong> Cultivation <strong>of</strong> Animal and <strong>Plant</strong> Cells. 1st.<br />

edition. Ronald Press, New York.<br />

WHITE P.R. 1963 <strong>The</strong> Cultivation <strong>of</strong> Animal and <strong>Plant</strong> Cells. 2nd<br />

edition. Ronald Press, New York.<br />

WHITTIER D.P. & STEEVES T.A. 1960 <strong>The</strong> induction <strong>of</strong><br />

apogamy in the bracken fern. Can. J. Bot. 38, 925-930.<br />

WILLIAMS R.R., TAJI A.A. & BOLTON J.A. 1984 In vitro<br />

propagation <strong>of</strong> Dampiera diversifolia and Prostanthera<br />

rotundifolia. <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 3, 273-281.<br />

WILLIAMS R.R., TAJI A.M. & BOLTON J.A. 1985 Specificity<br />

and interaction among auxins, light and pH in rooting <strong>of</strong><br />

Australian woody species in vitro. HortScience 20, 1052-1053.<br />

WILSON K.S. & CUTTER V.M. 1955 Localization <strong>of</strong> acid<br />

phosphatase in the embryo sac and endosperm <strong>of</strong> Cocos nucifera.<br />

Am. J. Bot. 42, 116-119.<br />

WOLF S., KALMAN-ROTEM N., YAKIR D. & ZIV M. 1998<br />

Autotrophic and heterotrophic carbon assimilation <strong>of</strong> in vitro<br />

grown potato (Solanum tuberosum L.) J. <strong>Plant</strong> Physiol. 135,<br />

574-580.<br />

WOLTER K.E. & SKOOG F. 1966 Nutritional requirements <strong>of</strong><br />

Fraxinus callus cultures. Am. J. Bot. 53, 263-269.<br />

WOOD H.N. & BRAUN A.C. 1961 Studies on the regulation <strong>of</strong><br />

certain essential biosynthetic systems in normal and crown-gall<br />

tumor cells. Proc. Natl. Acad. Sci. USA 47, 1907-1913.<br />

WOODWARD S., THOMSON R.J. & NEALE W. 1991<br />

Observations on the propagation <strong>of</strong> carnivorous plants by in vitro<br />

culture. Newsl. Int. <strong>Plant</strong> Prop. Soc. Spring 1991.<br />

WRIGHT K. & NORTHCOTE D.H. 1972 Induced root differentiation<br />

in sycamore callus. J. Cell Sci. 11, 319-337.<br />

WRIGHT M.S., KOEHLER S.M., HINCHEE M.A. & CARNES<br />

M.G. 1986 <strong>Plant</strong> regeneration by organogenesis in Glycine max.<br />

<strong>Plant</strong> Cell Rep. 5, 150-154.<br />

XIE J., GAO M., CAI Q., CHENG X., SHEN Y. & LIANG Z..<br />

1995 Improved isolated microspore culture efficiency in medium<br />

with maltose and optimized growth regulator combination in<br />

japonica rice (Oryza sativa). <strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 42,<br />

245-250.<br />

XU X., VAN LAMMEREN A.A.M., VERMEER E. &<br />

VREUGDENHIL D. 1998 <strong>The</strong> role <strong>of</strong> gibberellin, abscisic acid,<br />

and sucrose in the regulation <strong>of</strong> potato tuber formation in vitro.<br />

<strong>Plant</strong> Physiol. 117, 575-584.<br />

YATAZAWA M., FURUHASHI K. & SHIMIZU M. 1967<br />

Growth <strong>of</strong> callus tissue from rice root in vitro. <strong>Plant</strong> Cell<br />

Physiol. 8, 363-373.<br />

Chapter 4 173<br />

YOSHIDA F., KOBAYASHI T. & YOSHIDA T. 1973 <strong>The</strong><br />

mineral nutrition <strong>of</strong> cultured chlorophyllous cells <strong>of</strong> tobacco. I.<br />

Effects <strong>of</strong> p salts, p sucrose, Ca, Cl and B in the medium on the<br />

yield, friability, chlorophyll contents and mineral absorption <strong>of</strong><br />

cells. <strong>Plant</strong> Cell Physiol. 14, 329-339.<br />

YOUNG R.E., HALE A., CAMPER N.D., KEESE R.J. &<br />

ADELBERG J.W. 1991 Approaching mechanization <strong>of</strong> plant<br />

micropropagation. Trans. ASAE 34, 328-333<br />

YU S.X. & DORAN P.M. 1994 Oxygen requirements and masstransfer<br />

in hairy-root culture. Biotechnol. Bioeng. 44, 880-887.<br />

ZAGHMOUT O. & TORRES K.C. 1985 <strong>The</strong> effects <strong>of</strong> various<br />

carbohydrate sources and concentrations on the growth <strong>of</strong> Lilium<br />

longiflorum ‘Harson’ grown in vitro. HortScience 20, 660.<br />

ZAMSKI E. & WYSE R.E. 1985 Stereospecificity <strong>of</strong> glucose<br />

carrier in sugar beet suspension cells. <strong>Plant</strong> Physiol. 78, 291-295.<br />

ZAPATA F.J, KHUSH G.S., CRILL J.P., NEU M.R., ROMERO<br />

R.O., TORRIZO L.R. & ALEJAR M. 1983 Rice anther culture at<br />

IRRI. pp. 27-86 in Cell and <strong>Tissue</strong> <strong>Culture</strong> Techniques for Cereal<br />

Crop Improvement. Science Press Beijing, China. Gordon and<br />

Breach Science Publications, Inc., New York.<br />

ZATYKO J.M. & MOLNAR I. 1986 Adventitious root formation<br />

<strong>of</strong> different fruit species influenced by the pH <strong>of</strong> medium. p. 29<br />

in Abstracts VI Intl. Cong. <strong>Plant</strong> <strong>Tissue</strong> & Cell <strong>Culture</strong>.<br />

Minneapolis, Minn.<br />

ZHANG B., STOLTZ L.P. & SNYDER J.C. 1986 In vitro<br />

proliferation and in vivo establishment <strong>of</strong> Euphorbia fulgens.<br />

HortScience 21, 859.<br />

ZHOU T.S. 1995 In vitro culture <strong>of</strong> Doritaenopsis: comparison<br />

between formation <strong>of</strong> the hyperhydric protocorm-like-body<br />

(PLB) and the normal PLB. <strong>Plant</strong> Cell Rep. 15, 181-185.<br />

ZHU J.K. (2002) Salt and drought stress signal transduction in<br />

plants. Annu. Rev. <strong>Plant</strong> Biol. 53, 247-273.<br />

ZIMMERMAN R.H. 1979 <strong>The</strong> laboratory <strong>of</strong> micropropagation at<br />

Cesena, Italy. Comb. Proc. Int. <strong>Plant</strong> Prop. Soc. 29, 398-400.<br />

ZIV M, 1991. Quality <strong>of</strong> micropropagated plants - vitrification. In<br />

Vitro Cell. Dev. – Pl. 27, 64-69.<br />

ZIV, M. 1989. Enhanced shoot and cormlet proliferation in liquid<br />

cultured gladiolus buds by growth retardants. <strong>Plant</strong> Cell <strong>Tissue</strong><br />

Organ Cult. 17, 101-110.<br />

ZIV M. 2005 Simple bioreactors for mass propagation <strong>of</strong> plants.<br />

<strong>Plant</strong> Cell <strong>Tissue</strong> Organ Cult. 81, 277-285<br />

ZWAR J.A. & BRUCE M.I. 1970 Cytokinins from apple extract<br />

and coconut milk. Aust. J. Biol. Sci. 23, 289-297.<br />

ZWAR J.A., KEFFORD N.P., BOTTOMLEY W. & BRUCE M.I.<br />

1963 A comparison <strong>of</strong> plant cell division inducers from coconut<br />

milk and apple fruitlets. Nature 200, 679-680.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!