Вы находитесь на странице: 1из 24

ANNUAL

REVIEWS

Further

Quick links to online content

Copyright

1968. All

rights reserved

ECOLOGY, PHYSIOLOGY, AND BIOCHEMISTRY


OF BLUE-GREEN ALGAE1,2
OSMUND HOLM-HANSEN

By

Institute of Marine Resources, University of California, San Diego,

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

La Jolta, California

CONTENTS
INTRODUCTION
PHOTOSYNTHESIS
NITROGEN ASSIMILATION
RESPIRATION ............
. . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .
. . . . . .

..

. . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.

OBLIGATE PHOTOAUTOTROPHY.......................................... .
NUCLEIC ACIDS ...................................................... .

CHEMICAL COMPOSITION ...............................................


MINERAL NUTRITION
.
.
VITAMINS AND GROWTH SUBSTANCES ....................................
HYDROGEN ION REQUIREMENTS ...........................................
CELLULAR DIFFERENTIATION ..........................................
CELL MOVEMENTS .....................................................
TOXIC METABOLIC FACTORS ...........................................
TOLERANCE TO HIGH AND Low TEMPERATURES AND TO DESICCATION .......
SUSCEPTIBILITY TO VIRUSES
BLUE-GREEN ALGAE IN SYMBIOTIC RELATIONSHIPS .......................
.

. . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

.
.

. . . . . . .

. .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

47
48
50
52
54
ss
57
58
59
59
60
61
61
62
63
64

INTRODUCTION
In the continuum of life from the virus particles through the unicellular
microorganisms and, finally, to higher plants and animals, the Cyanophyta
(blue-green algae) occupy a unique position. As was described in Lang
(1968 ) , the blue-green algae have a procaryotic cellular organization which
is similar to that found in bacteria. They differ from bacteria, however, in
that they are generally obligate photoautotrophs, obtaining their carbon and
energy by photosynthetic mechanisms which are similar to those found in
higher plants. In eucaryotic organisms, various series of metabolic reactions
may be localized within the cell in specialized organelles. Studies with mito
chondria ( 1, 2) and chloroplasts (3, 4) have demonstrated an ordered array
of enzymes in specific subunits on the constituent membranes of these orga
nelles. Although the membranous structures and granules in cells of blue
green algae can often be related to specific metabolic functions, they are not
separated from the surrounding cytoplasm by a continuous membrane. In
1 The survey of the literature pertaining to this review was concluded in January
1968.
Supported by U. S. Atomic Ene rgy 'Commission Contract No. AT(ll-l)Gen
10, P.A. 20.

47

1502

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

HOLM-HANSEN

spite of this lack of compartmentalization, the blue-green algal cell is capa


ble of all metabolic reactions demanded of a photoautotrophic organism,
and grows and respires at rates comparable to those observed in eucaryotic
algae. One of the basic research problems associated with procaryotic cells
is therefore the localization of enzyme packets within the cell, and the cel
lular mechanisms that ,ensure efficient functioning of metabolic pathways
which may be in close proximity. Thus, the enzymes concerned with photo
synthesis, respiration, and nitrogen fixation in blue-green algae are all lo
cated on the lamellae (thylakoids), but as yet have not been separated by
any physical or chemical means. Recent evidence from freeze-etch micros
copy does indicate the existence of subunits in the lamellae of blue-green
cells (5).
Blue-green algae are widely distributed in practically all habitats which
will support life, including hot springs, deserts, saline lakes, the Antarctic,
and oil field sump ponds (6--9) . They are of considerable ecological impor
tance in regard to increasing the fertility of terrestrial and aquatic environ
ments, and to their activity as colonizers of newly exposed land surfaces.
Some of the fresh-water species, however, grow in such profusion that they
result in serious losses of economic and aesthetic resources. Many of the
ecologically important aspects of blue-green algae are discussed in the fol
lowing sections, together with physiological and biochemical considerations.
Information on other aspects which have been omitted from this review
may be found elsewhere (10- 13 ) . Although the blue-green algae comprise a
wide assemblage of free-living and symbiotic forms, the bulk of laboratory
investigations have been done with only a few dozen species. As the species
chosen for laboratory studies are generally those which are easy to culti
vate, it is possible that many deviations from the generalizations presented
in this review may emerge from future studies on some of the lesser-known
genera.
There is a large assemblage of nonphotosynthetic organisms which may
be considered "colorless blue-green algae" and included in the Cyanophyta.
Such apoehlorotic Cyanophyta might include the Flexibacteria, chemosyn
thetic forms such as Beggiatoa and Thiothriz, and heterotrophic forms such
as Oscillospira. The phylogenetic affinities of these varied groups have re
cently been discussed by Soriano & Lewin ( 14 ) . Most of these apochlorotic
forms are grouped with the photosynthetic blue-green algae predominately
on morphological similarities, but are quite distinct on physiological criteria.
This review is concerned only with the photosynthetic Cyanophyta.
PHOTOSYNTHESIS

The photosynthetic mechanism in blue-green algae is basically similar to


that in higher plant photosynthesis, in that water is the ultimate hydrogen
donor and molecular oxygen is liberated. A few strains have been shown to
be able to reduce CO2 with H2 under anaerobic conditions in the light (15).
The basic photochemical and chemical reactions as determined for green

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

BLUE-GREEN ALGAE

49

plant photosynthesis ( 16, 17 ) also occur in blue-green algae ( 18 ) . Thus,


there is convincing evidence for the occurrence of both photosystem I
( long wavelength ) and II ( short wavelength ) , and for both cyclic and non
cyclic photophosphorylation ( 19 ) . The bacterial photosynthetic apparatus is
quite different, having different photosynthetic pigments and having typi
cally only one photor.eaction (20).
Analysis of in vivo absorption spectra indicate that several molecular
forms of chlorophyIl-a exist in the blue-green algal ceIl (21). These forms
are believed to be related to their function in photo systems I and II. Ac
cording to current theory, the main function of the accessory pigments ( the
phycobiliproteins ) is to absorb light quanta and transfer the energy to the
active chlorophyll molecules involved in photo system II, resulting in libera
tion of O2 and in raising the potential of an electron so that it can be trans
ferred via the electron transport chain to photosystem I with concomitant
formation of ATP. There is evidence for the participation of plastoquinone
(22,23), plastocyanin (24), and cytochrome C554 (25) as electron mediators
in this series of reactions between photosystems I and II. The accessory
pigments apparently are not essential for HilI reaction activity (26, 27) or
for photophosphorylation ( 28, 29 ) .
The pigments involved in photo system I are chlorophyIl-a and a minor
chlorophyll-a component designated P-700 ( 30 ) . The photoreactions of sys
tem I involve cylic photophosphorylation, which results in formation of
A TP, and noncyclic photophosphorylation, which results in formation of
both ATP and a strong reductant. This reductant, which is believed to be
ferredoxin, can then reduce NADP in the presence of the enzyme ferredox
in-NADP-reductase (27, 31 ) . Reduction of NADP by hydrogenase has
been reported by Fujita & Myers in Anabaena cylindrica ( 32 ) . The ease of
solubilization of essential components in cell-free preparations of blue-green
cells results in uncoupling of both noncyc1ic and cyclic photophos
phorylation. N oncyclic photophosphorylation can be restored by the addition
of known compounds to the washed lamellae, but cyclic photophos
phorylation seems to require some solubilized factor which does not appear
to be a known compound in the electron transport system of chloroplasts
( 33 ) . Biggins has shown that the bulk of cytochrome C554 is easily washed
off the lamellae, but that a small amount of it is apparently firmly bound to
the membranes and is not easily removed. Such tightly bound forms of elec
tron transport mediators are believed to be of great functional significance
in the photoreactions of the lamellae ( 33 ) . 'Other electron carriers that
have been found in blue-green algae and may play a role in the photosyn
thetic mechanism are a-tocopherolquinone and vitamin K1 ( 34 ) , a specific
cytochrome c-reducing substance (22, 35, 36 ) , and various pteridines
( 37-39 ) .
As the absorption maxima of the accessory photosynthetic pigments are
relatively far from those of chlorophyll-a, the blue-green algae are good ex
perimental organisms for the study of chromatic adaptation, whereby the

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

50

HOLM-HANSEN

concentration of pigments is partially controlled by the wavelength and in


tensity of the absorbed light (21, 40-42), and also by temperature (8).
Ghosh & Govindjee, in studies on the efficiency of energy transfer from
phycocyanin to chlorophyll in cells with widely varying ratios of
phycocyanin/chlorophyll, have postulated that the chlorophyll-a in photosys
tem II is highly fluorescent, while that in photosystem I is weakly fluores
cent (43). Jones & Myers (21) have shown that growth of Anacystis nidu
lans in red light results in a dramatic decrease in the chlorophyll content
while not affecting the concentration of phycocyanin. They interpreted this
pigment alteration as a control mechanism to balance the two light reac
tions of photosynthesis. The concentration of phycocyanin was not, how
ever, affected by growing the cells in monochromatic light of 620 mIJ.; this,
they speculated might be due to energy transfer from phycocyanin directly
to photosystem I in addition to its usual functioning in system II.
Photosynthetic incorporation of CO2 in blue-green algae has been shown
to occur predominately via the Calvin cycle involving carboxylation of ribu
lose diphosphate (44,45). Some CO2 fixation may also occur via such reac
tions as the carbamyl phosphate synthetase reaction (46), the reductive car
boxylic acid cycle as described for some bacteria (47), and by carboxyla
tion of phosphoenolpyruvate (48). There is no evidence, however, that the
carboxylation reactions in this latter group are involved in the light-induced
assimilation of CO2 by cells of blue-green algae. Some of these reactions
are involved in dark-fixation reactions, however, as the compounds which
incorporate radiocarbon from CO2 during dark-fixation are predominantly
the carboxylic acids and associated amino acids, including citrulline (49,
50). There are reports in the literature that blue-green algae can live anaer
obically via a chemosynthetic mechanism under natural conditions (51).
Neither the ability to grow anaerobically nor the ability to live chemo
synthetically has, however, been experimentally demonstrated with any
blue-green alga.
NITROGEN ASSIMILATION

Blue-green algae can assimilate nitrogen from nitrate, nitrite (52), am


monia (53), hydroxylamine (52), urea (54, 55), casein (56), amino acids
(57), uric acid (58), and N2 (59). There is considerable variation among
species as to which forms of nitrogen can be assimilated, but all blue-green
algae apparently can grow with nitrate as the sole nitrogen source. The re
duction of nitrate proceeds through nitrite to ammonia. Nitrate reductase,
which is a Mo-containing protein, is found in the particulate fraction of
cell-free preparations of Anabaena cylindrica, whereas nitrite reductase is
recovered in the soluble fraction (60). Ferredoxin and NADH are required
for nitrate reduction, while ferredoxin and NADPH are required for ni
trite reduction (61). As these reductants are available through the reactions
of photosystem I of photosynthesis, the reduction of nitrate in intact cells is
light-stimulated. Some blue-greens also possess hydrogenase which catalyzes

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

BLUE-GREEN ALGAE

51

the reduction of nitrate to ammonia in light with H2 as the hydrogen donor


(62). The hydrogenase enzyme is located on the lamellae, but is easily solu
bilized by acetone treatment.
The ability to reduce molecular nitrogen is found in many filamentous
forms (63), but has not been reportcd for any unicellular species of the
order Chroococcales ( 59 ) . The importance of nitrogen-fixing algae for
maintenance of soil and water fertility has recently been summarized by
Stewart (59 ) . The fixation of N2 is particularly important in the coloniza
tion of recently exposed land surfaces and in accumulating a reservoir of
reduced carbon and nitrogen compounds which is essential for the establish
ment of metazoans, eucaryotic algae, and higher plants (64, 65). Such colon
ization is important in areas denuded of life by volcanic activity or exces
sive radiation, as well as in desert areas, sand dunes, and on land exposed
by receding glaciers (66). The significance of N2 fixation by marine blue
green algae has not been adequately assessed, although there are many spe
cies in the benthic flora and in the open ocean which can fix nitrogen (67,
68). The open-ocean forms associated with N2 fixation are generally species
of Trichodesmium or Nostoc (63). For most critical studies on N2 fixation,
the stable isotope N15 has been employed, but recently it was found that
nitrogen fixation can be followed in a gas chromatograph by determining
the amount of reduction of acetylene to ethylene (69). As the methodology
and equipment required for the latter method are much simpler than those
required for N15 analysis, this technique should aid considerably in evaluat
ing the contribution of blue-green algae to the fertility of terrestrial and
aquatic environments.
Studies with intact cells revealed that the mechanism of N2 fixation in
blue-green algae had many similarities to that found in some free-living
bacteria and in nodules of leguminous crops (70 ) . The effects of various
inhibitors and of the composition of the gaseous phase in ,experiments with
cell-free preparations (71) have also demonstrated the basic similarity in
the mechanism of N2 fixation in all microorganisms. The enzymes involved
in fixation of molecular nitrogen are located on the lamellae (72, 73) ; they
as yet have not been separated from the enzymes involved in respiration
and photosynthesis. The first stable intermediate which has been demon
strated in these experiments is ammonia, but the intermediates between N 2
and NHa have not been detected (59 ) . Fay & Fogg (74) have suggested
that photochemically produced hydrogen donors are utilized in N2 fixation,
but this has recently been questioned by Cox (75). Although there seems to
be a close functional relationship between photosynthesis and nitrogen fixa
tion in cells, as witnessed by a much higher rate of N2 fixation in the light
as compared to the dark, there does not seem to be any obligate photochemi
cal reaction for the fixation of molecular nitrogen. The evidence for this is
that many blue-green algae can fix N2 in the dark if supplied with certain
organic substrates (76, 77), although the rate of fixation is very low. Cox
(78) has also shown that the rate of fixation in the dark is similar to that

HOLM-HANSEN

52
in the light if

CO2

is excluded. This low rate of fixation of N 2 in the light

in the absence of CO2 indicates the possible dependence of nitrogen fixation


for carbon compounds supplied by the photosynthetic incorporation of

CO2,

Since N2 fixation in cell-free preparations is accelerated by pyruvate and is


accompanied by liberation of CO2 (78), and since decarboxylation enzymes
have previously been shown to be associated with the lamellae (79), the en

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

ergy for N2 reduction may be derived from the phosphoroclastic cleavage


of pyruvate, as has been shown for the bacterial system

(80).

Although the

fixation of N2 can thus be driven by light-independent electron transport


systems, one should not exclude the possibility that in cultures exposed to
light much or all of the required reducing power and ATP is furnished di
rectly by the photosynthetic apparatus. Various mineral elements such as
Mo, Mn, Co, Fe, and Na have been shown to have marked effects on the
nitrogen fixation process (59, 71, 81-83).
RESPIRATION
Very little is known of the respiratory pathways in blue-green algae.
The Embden-Myerhof glycolytic pathway apparently functions in these
algae, as does the pentose phosphate cycle (84, 8S). On the basis of incor
poration patterns from C14-1abeled organic substrates or on enzymatic anal
yses, several investigators have suggested that the tricarboxylic acid cycle

(TCA)
evidence

is either very sluggish or is not fully operative (86-88). The main


from the labeling experiments is that the radiocarbon from

2-C14-acetate is found in lipids and in four amino acids (glutamic acid,


proline, arginine, and leucine), but not in aspartic acid. This indicates that
the TCA cycle may be operative up to the formation of a-ketoglutaric acid,
which is then converted to glutamic acid. The other labeled amino acids can
be obtained from glutamic acid by the usual metabolic pathways. Support
for this hypothesis is afforded by the failure to detect a-ketoglutarate dehy
drogenase and succinyl-CoA in blue-green cells (86, 88). Acetyl-CoA syn
thetase, citrate synthase, aconitase, iso-citrate NADP dehydrogenase, suc
cinic dehydrogenase, and malic dehydrogenase have all been demonstrated
in blue-green algae (86,

87).

The argument that the TCA cycle is not func

tioning in these algae must be considered relative to the fact that the prop
erties of some enzymes in the blue-green algae are sufficiently different
from the corresponding enzymes in higher plants that the assay methods
may not be valid. Thus, although certain investigators were unable to dem
onstrate aldolase in any of the blue-green algae tested

(89), it is now

known that aldolase (type II) is present in these algae, but was not de
tected earlier because of unsatisfactory assay methods

(90-93).

The enzymes, isocitrate lyase and malate synthase, which are necessary
for the operation of the glyoxylic cycle, have been demonstrated in cells of
blue-green algae, though they do not show the adaptive behavior which the
corresponding enzymes show in preparations from higher plants (94, 95).
The net effect of the glyoxylic cycle is to synthesize one succinate molecule

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

BLUE-GREEN ALGAE

53

fram twa acetate malecules. The blue-green algae therefare apparently have
the enzymes required to canvert acetate to succinate and thence to' axalaac
etic acid, which can be canverted to' aspartic acid by transaminatian. This
suggests that the inability to' demanstrate labeled aspartic acid in blue-green
algae c clls when incubated with C14-acetate might be indicative af clase
cellular cantral aver the pathways by which acetate is assimilated, rathcr
than the absence af any particular enzyme system.
The electron transport chain involved in respiration also appears to be
associated with the lamellae. We know little about the electron mediators,
the constituent enzymes, and the terminal oxidase in these organisms. Sev
eral investigators have suggested that blue-green algae are dependent upon
photophosphorylation to supply most if not all the ATP required by the cell.
These speculations were based on the failure to find NADH oxidase (86)
and on the assumption that cells in the dark could not synthesize acetyl
CoA from acetate (96). Such an extreme view concerning the inability of
these algae to synthesize ATP in the dark seems unwarranted in view of
the following points. First, a few blue-green cells have been shown to be
capable of slow heterotrophic growth in complete darkness (76, 97). It is
obvious that such cells must have an ATP-generating mechanism other than
that of phatophaspharylatian Secand, even those blue-green algae which
are nat capable af heteratraphic growth are capable of surviving in com
plete darkness far many manths, during which time they are presumably
generating ATP fram cellular campanents. Third, the ATP cantent of Ana
cy stis sp., bath in light and in darkncss, is abaut 0.2 per cent of the dry
weight (Halm Hansen, unpublished data), which is camparable to' that af
mo s t other algae that have been investigated (98, 99).
Although Smith, London & Stanier (86) were unable to find NADH
oxidase in Anaeystis nidulans, Coeeoehloris penioeystis, or Gloeoeapsa alpi
cola, Webster & Hackett (100 ) demonstrated the presence of this enzyme
in three species of apochlorotic blue-green algae (the authors referred to
these as blue-green algae, but they are Flexibacterales). The NADH oxi
dase activity associated with the particulate fraction of these organisms
showed a response to inhibitors different from that of green algae and
higher plants.
Holton & Myers have described three different e-type cytochromes (e554,
e549, and e552) in Anaeystis nidulans (25, 101). The terminal oxidase in the
electron transport system has not been unequivocally demonstrated, but
studies with inhibitors have indicated the functioning of cytochrome oxidase
(102). Polyphenol-oxidase activity has not been detected (102).
Various studies have pointed to the interaction of the photosynthetic and
respiratory mechanisms in blue-green algae. Carr & Hallway ( 1 03) showed
that cell-free preparatians af Anabaena variabilis reduce phenalinda-2
6-dichlaraphenal in the light, using hydragen fram the phatalysis af water,
whereas, in the dark, endagenaus arganic substrates served as the hydrogen
donor. On the basis af inhibitar studies, they cancluded that either the same
.

S4

HOLM-HANSEN

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

electron transport chains function in both photosynthesis and respiration, or

that photosynthetic electron transport suppresses the activity of the respira


tory electron transport system. Other reports show that the respiratory up
take of oxygen is depressed at low light levels but may be stimulated by
bright light (104, lOS). Stimulation at high light levels, unaccompanie d by a
corresponding increase in CO2 production, might result from photoreduc
tion of NADP, which is then oxidized via the cytochrome system to yield
ATP (105). This uptake of O2 without concomitant libcration of CO2
might also result from formation of peroxides via pseudocyclic photophos
phorylation.
Hood & Carr (84) have shown that the glyceraldehyde-3-phosphate de
hydrogenas e in blu e-g r een a lg a e can use either NADP or NAD as c o en zyme
They postulate one active site on the molecule, so that it might function
either in the pentose phosphate cycle or in the glycolytic pathway, depend
ing on whi ch coenzyme is occupying the act iv e site.
.

OBLIGATE PHOTOAUTOTROPHY
Most blue-green algae which have been investigated are obligate pho
toautotrophs, in that they require light for continued growth and apparently
can not grow heterotrophically. A few early claims (106, 107) of hetero
trophic growth in certain of these algae, notably species of Nostoe, have

cited so often that facultative heterotrophy is commonly attributed to


all the Cyanophyta to explain the predominance of blue-green algae in wa
ters rich in dissolved organic materials. The early reports of heterotrophy
in cells of blue-green algae, however, show a growth rate in the dark con
siderably below that under autotrophic conditions and indicate only limited
heterotrophic potential, if any.
been

Recently there have been several new reports of the ability of a few spe

cies of

bl ue - gre ens

to

grow very slowly

in complete

darkness if supplied

with the proper carbon sources. A strain of Chlorogloea fritsehii has been
shown to grow continuously in the dark (76), though only after a long ad
aptation period. In the experiments of Fay there was no growth of this alga

during

the first month in the dark, and only slow growth during the second

and third months. Tolypothrix tenuis has also been shown to grow in the

supplied with glucose and casamino acids (97), but not if nitrate is
the sole nitrogen source. The authors report that the heterotrophic growth
was very slow compared to that in autotrophic conditions, and also that phy
coerythrin was not synthesized in the dark. It is not clear if this alga is
capable of co nti nuous growth in the dark, a s light gro wn cultures were in
oculated into flasks which were placed in the dark for just 10 days before
growth was measured. Other investigators have noted apparent growth of
dark if

blue-green algae when first placed in darkness, but that subsequent transfers

failed to show any heterotrophic growth (108).


The lack of heterotrophic growth in most species of blue-green algae and
the lack of stimulation of their respiration upon addition of organic sub-

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

BLUE-GREEN ALGAE

55

strates (95, 109) have often been postulated to be due to failure of the or
ganic substrate to penetrate the plasma membrane of the cell (110). This
hypothesis does not appear to be valid, however, as cell-free preparations of
many blue-green cells show little or no assimilation of added organic sub
strates; also, studies with C14-labeled compounds demonstrate that some spe
cies can assimilate added substrates when grown in the light. Thus, Carr &
Pearce (111) have shown that when cells of A. variabilis or Anacystis ni
dulans are grown in light with U-C14-glucose or 2-CH-acetate, 18 to 32 per
cent of the cellular dry weight is derived from the labeled organic sub
strate. The photoassimilation of labeled acetate results in the radiocarbon
being distributed into lipids, into four protein amino acids ( glutamic acid,
arginine, proline, and leucine), and into carboxylic acids (96, 112). T. ten
His also shows photoassimilation of glucose (113 ) ; in the dark, glucose was
respired, whereas in the light it was polymerized into polysaccharides.
The biochemical basis for obligate photoautotrophy is not known, though
suggestions have recently been made that it involves incomplete functioning
of the TCA cycle or an inability to generate ATP in the dark. The main
evidence for an incomplete TCA cycle is the absence of demonstrable IX-ke
toglutarate dehydrogenase activity and the lack of radiocarbon in aspartic
acid after incubation with C'4-acetate ( 86-88). There is no direct evidence
in support of the speculation ( 86, 96) that blue-green algae cannot generate
ATP in the dark. In view of the demonstrated ( 86) conversion of acetate
to acetyl-CoA ( and thence to IX-ketoglutaric acid and related compounds)
and the concentration of ATP in these cells in the light and in the dark
( see section on Respiration), it is evident that blue-green algae are capable
of at least some A TP synthesis in the dark.
It is possible that blue-green algae may depend upon photophos
phorylation for the bulk of their ATP, but that they can synthesize some in
the dark by substrate phosphorylation reactions. This suggestion (86), which
is based on the inability to demonstrate NADH oxidase activity in these
algae, might account for the slow assimilation of organic compounds and
the failure to couple ATP synthesis with their oxidation. This is an inter
esting speculation, but its validity must await more definitive investigations
of the respiratory pathways in these algae.
NUCLEIC ACIDS

The deoxyribonucleic acid of blue-green algae is not localized in a mem


brane-limited nucleus nor does it ever condense into cytologically de
monstrable chromosomes as during mitosis in eucaryotic cells. In the area
of the nucleoplasm, Ris & Singh (114) have shown the presence of microfi
brils which they refer to as DNA fibrils. Leak ( 115) has recently described
the occurrence of three types of DNA fibrils in A. variabilis, ranging in
size from 20 to 30 A in diameter to aggregations from 100 to 350 A in diam
eter. The manner in which this DNA is duplicated and divided between the
daughter cells has not been studied with blue-green algae. It is likely, how-

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

56

HOLM-HANSEN

ever, that the organization of the DNA may resemble that found in some
bacteria. In Escherichia coli the bacterial chromosome normally consists of
a single loop of two-stranded DNA (116, 117). As it is difficult to obtain
any degree of synchronous division in blue-green algae, it is not known
whether the synthesis of DNA occurs during a relatively short period dur
ing the growth of a cell, or during most of the generation time as in some
bacteria (117). The DNA in eucaryotic cells is complexed with histones,
which ar.e believed to be involved in directing the activity of the nucleotide
units of the DNA molecule. No histones have been demonstrated, however,
in bacteria or in blue-green algae ( 1 18, 1 19) .
In a comprehensive study of 29 strains of blue-green algae, Edelman and
co-workers reported that buoyant density patterns of the DNA in a CsCl
gradient showed a unimodal distribution (120). There werc no DNA satel
lite bands, although there was a small satellite band possibly attributable to
contaminant polysaccharides. This is in contrast to eucaryotic cells, in
which the cellular DNA often shows a bimodal or trimodal pattern of DNA
density profiles after CsCI density gradient centrifugation, reflecting vary
ing chemical composition of the DNA found in the nucleus, chloroplasts,
and mitochondria. The filamentous blue-green algae examined had a re
markably similar base composition of their DNA ( about 45 per cent guan
ine-cytosine) calculated from a CsCl buoyant density of 1.70 g/cm3. The
base composition of the DNA in unicellular species ( members of the Chroo
coccales) was much more variable, ranging from 35 per cent GC to 71 per
cent GC. Other investigato rs have reported from Plectonema boryanum a
satellite DNA band which had a lower GC content than the main band of
DNA ( 121 ) , and which they speculated might represent DNA localized in
the photosynthetic lamellae.
The organization of the genetic material in blue-green algae apparently
confers great resistance to alterations by radiation or certain chemical sub
stances which, in eucaryotic organisms, interfere with cell division ( 122,
123 ) . Kraus has reported that many species of blue-green algae suffered no
apparent damage after exposure to 260,000 rads over a four-hour period
from a C060 source (124 ) . Most blue-green cells are also able to survive
large doses of ultraviolet radiation; this tolerance to ultraviolet has been
utilized to obtain bacteria-free cultures for laboratory investigations ( 125,
126 ) . Kumar has shown that colchicine has no effect on the growth of these
algae (127) ; this is undoubtedly related to the fact that colchicine, which in
terferes with proper functioning of the spindle in eucaryotic cells, has no
corresponding site of action in procaryotic cells. Kumar has also claimed
mutational changes in blue-green algae after exposure to streptomycin and
penicillin, but it is not clear whether these are true mutations or physiologi
cal adaptations ( 128) . Van Baalen has reported obtaining clonal mutants of
Anacystis nidulans by treatment with antibiotics, but gave no details on the
characteristics and stability of the mutant strains ( 129) .
There have been vcry fcw studics of RNA composition and function in

S7

BLUE-GREEN ALGAE

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

blue-green algae. Capesius & Richter have reported that in A. nidulans most
of the polynucleotide phosphorylase activity is associated with the ribosomal
fraction ( 130), and that RNA polymerase is associated with the DNA frac
tion (131). Norton & Roth have described an RNase enzyme from A. nidu
lans which is active against 2'-O-methyl RNA (132). They have also re
ported a second RNase enzyme which seems to be located either on or near
the cell wall of Anacystis (133).
CHEMICAL COMPOSITION
The gross chemical composition of blue-green cells is basically similar to
that of eucaryotic algae. Though environmental factors can greatly influ
ence the composition of cells, typical analyses (134, 135) of marine and
fresh-water species show the following composition, based on dry weights:
carbohydrate, 30 to 55 per cent; protein, 20 to 45 per cent; lipid, 1 5 per
cent; RNA, 2 per cent; DNA, 0.4 to 0.8 per cent; total pigment, 1 .5 per
cent.
The amino acids of the proteins from blue-green algae do not differ
appreciably from those of most other microorganisms ( 136, 137). Orni
thine, a constituent of certain cyclic polypeptides in bacteria (138), has
been reported also in the water-insoluble fraction of certain blue-green
algae (139). The nature of the compound in which ornithine occurs was not
further investigated.
The bulk of the constituent fatty acids in blue-green cells contains be
tween 10 and 18 carbon atoms, and are saturated or have only one double
bond. Many of the filamentous forms also contain di- and tri-polyunsatu
rated fatty acids (140-143). In cells containing a-linolenic acid, most of this
component seems to be localized in the lamellae (142). Nichols & Wood
(144) have reported that Spirulina platensis contains a large quantity of
",(-linolenic acid ( 6,9,12-octadecatrienoic acid). There seems to be no signifi
cant amounts of branched or unusual fatty acids, such as are found in many
bacteria ( 140, 141). The other component lipids of blue-green algae resem
ble those of bacteria in lacking sterols, lecithin, phosphatidyl-ethanolamine,
and phosphatidyl-inositol, and resemble those of higher plants in comprising
two galactosyl diglycerides, sulphoquinovosyl diglyceride, and phosphati
dyl-glycerol ( 143). Carr ( 1 45) has reported that when Chlorogloea fritschii
is grown with acetate in the light, about 10 per cent of the cellular dry
weight can be recovered in poly--hydroxybutyrate.
There is considerable variation in the polysaccharides synthesized by
blue-green alga and in the component monosaccharide units. The so-called
(X- and - granules, which may be storage products, are believed to consist
of a polysaccharide related to glycogen (146). About 25 per cent of the
carbohydrates of Tolypothrix tenuis consists of a nonreducing glucofructan
which was tentatively identified as fructofuranosyl (),,-fructofurano
syl- ( 2-;.I)a-glucopyranoside ( 147). The intracellular polysaccharides of
-

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

58

HOLM-HANSEN

Anabaena fios-aquae consist mostly of glucose and xylose residues, with


lesser amounts of glucuronic acid and ribose (148 ) .
The cell wall ( the three o r four layers lying between the plasma mem
brane and the sheath), and the sheaths of blue-greens when present, have
distinctive structures and compositions which resemble those in various
bacteria ( 10, 134). The cell walls of A. nidulans contain 24 per cent sugar,
28 per cent protein, and 36 per cent lipid, according to Drews & Gollwitzer
( 149). The main components of the carbohydrate fraction were mannose,
glucose, galactose, and fucose. Purified cell walls of Phormidium uncinatum
contain eight constituent amino acids in the peptide units, as well as
2,6-diaminopimelic acid and muramic acid (150). These latter two acids, to
gether with glucosamine, seem to be characteristic of blue-green cell walls
( 10, 15 1 ). Fairly similar cell wall composition has been reported for Lyng
bya and Nostoc species (152). The inner membrane of the eel! wall consists
of globular protein and mucopolymer molecules. This mucopeptide layer,
which is believed to be largely responsible for much of the strength of the
cell wall, can be disrupted by the action of lysozyme and penicillin. The ge
latinous sheath has been described as consisting of celluiose fibrils in a pro
teinaceous matrix (10), though there is some question concerning the valicl
ity of the cellulose identification. When the sheath material is hydrolyzed,
the major monosaccharide units are generally glucose and xylose, with
lesser amounts of arabinose, fucose, and glucuronic acid (148, 1 53).
In addition to the five usual nitrogenous bases in nucleic acids ( gua
nine, cytosine, adenine, thymine, and uracil), the methylated bases
6-methylaminopurine and 5-methylcytosine have also been reported in Plec
tonema boryanum (121).
The main photosynthetic pigments in blue-green algae are chlorophyll-a
and the phycobiliproteins phycocyanin, allophycocyanin, and phycoerythrin
( 154). The three phycobilin pigments have been crystallized by Hattori &
Fujita (ISS); the chromophores of phycocyanin and allophycocyanin are
identical, indicating that the in vivo spectral differences are due to differ
ences in the protein moiety ( 156). A pigment with an absorption band at
750 m[L has also been reported in several blue-green algae, but it has not yet
been identified ( 1 57, 1 5 8 ) . The carotenoids most commonly found are
carotene, zeaxanthin, echinenone, myxoxanthophyII, oscillaxanthin, flavacin,
aphanicin, aphanizophyll, cryptoxanthin, and one hydroxylated ketocaroten
aid (159-162),
MINERAL NUTRITION

In addition to the major mineral elements N, P, Mg, S, and K ( 54, 163),


the following elements are also considered to be essential for at least some,
and perhaps all blue-green algae; calcium (53, 164, 165), iron (53, 163),
boron ( 166), molybdenum (164, 167), manganese ( 168), sodium ( 169), and
cobalt (82, 1 65, 1 70). Requirements for zinc and copper have not been dem
onstrated in these algae, but because of their many known functions as

59

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

BLUE-GREEN ALGAE

enzyme co-factors, it may be assumed that they are also essential microe1e
ments in these algae. The functions of these microelements are basically
similar to those described from higher plant and animal studies (171, 172)
with the implication of Mo, Mn, Fe, and Co also in the process of nitrogen
fixation. Vanadium has been cited to play a role in nitrogen fixation (173),
but the beneficial effects which Bortels obtained by the addition of vana
dium might have been due to molybdenum contaminating his vanadium
salts. Holm-Hansen (168) could not find any evidence for essentiality of
vanadium for either Nastae musear1t11t or Calathrix parietina. There has
been no recent demonstration of V essentiality for blue-green algae.
VITAMINS AND GROWTH SUBSTANCES
There ,ire only two reports in the literature, to the author's knowledge,
in which a vitamin requirement was demonstrated for blue-green algae.
Van Baalen (108) reported that eight of 15 marine blue-green species he
isolated required the addition of vitamin B12 for growth; Pintner & Provasoli
(174) also reported a vitamin B12 requirement by Phormidium persicinum.
None of the commonly cultured strains of blue-greens requires any organic
additions. The cobalt requirement can, however, be partially or wholly
substituted for by the addition of vitamin B12 (82). The vitamin composition
of these algae (other than that of B12) has not been studied, though there
is no reason to believe that it is significantly different from that of other
algae and higher plants.
The author is unaware of any reports on the occurrence of natural
growth substances (auxins, kinetics, etc.) in pure cultures of blue-green
algae, or on the effects on cellular growth elicited by the addition of such
substances to the nutrient medium.3
HYDROGEN ION REQUIREMENTS
Most blue-green algae grow in environments which are neutral to alka
line, with a few species occurring in habitats with a pH between 5.0 to 6.0
(175). Alkaline hot springs generally have abundant growth of these algae,
whereas acid hot springs (pH 5.0 and lower) do not (8). In laboratory cul
tqre, the 'pH optimum for growth seems to be between 7.5 to 10.0, but
growth still occurs at pH values over 11.0 (53,54,163, 176). The lower pH
limit for laboratory cultures is about 6.5 to 7.0 (54,57). There are no data
on the internal pH of blue-green cells,nor is the author aware of any expla
nation as to why these cells generally seem to require such alkaline condi
tions. The pH optima of extracted enzymes from them apparently do not
3 Some auxin activity has been reported
in samp les from a fresh-water bloom
of O.scillatoria species: see Mowat, J. A., Botan. Marina, 8, 149-55 (1965). A

recent paper reports a slight effect of indole-3-acetic acid on the growth of six

species
(1968).

of blue-green algae: see

Ahmad, M. R.,

Winter,

A., Planta, 78, 277-86

60

HOLM-HANSEN

differ significantly from those of eucaryotic cells. It is possible that the high
pH is related to permeability characteristics of the cell wall and membrane
and thus reflects peculiarities either in the ion uptake by the cells or in loss
to the nutrient solution of soluble essential metabolites.

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

CELLULAR DIFFERENTIATION

Various specialized structures occur in the blue-green algae such as het


erocysts, akinetes, endospores, and hormogonia. Little is known concerning
the control of differentiation of structures originating from vegetative cells.
Akinetes and endospores apparently are reproductive structures to enable
survival of the organism during periods of unfavorable growing conditions,
but the function of the heterocyst is much more obscure. It has been sug
gested that heterocysts are active as storage vesicles, as reproductive cells,
in nitrogen fixation, in controlling differentiation of other cells, or are
merely vestigial structures (177, 178 ) . Wolk (179 ) has reported germina
tion of heterocysts, but this has yet to be confirmed by other investigators.
Fay & Walsby (180 ) have reported that heterocyst preparations do not
show any N2- or CO2-fixing ability, but that they do show a greater respira
tory rate than the other cells of the filament. It has recently been shown in
A. cylindrica (181) that heterocysts inhibit the development of nearby cells
into heterocysts. As spores are commonly found adjacent to heterocysts, it
seems that the differentiation of cells is controlled to some extent by the
activity of other cells in the filament.
Although some authors have reported some degree of synchronized
growth by manipulation of light and temperature regimes (182), synchro
nous growth of blue-green algae has not been achieved to the degree ob
tained with eucaryotic algae. Lazaroff and co-workers have utilized a light
triggered growth response of Nostoc muscorum to obtain cultures which
differentiate from a sporogenous (aseriate) stage to a heterocystic stage
(183, 184 ) . In cells grown heterotrophically in the dark, cellular develop
ment is blocked at the aseriate stage; further development into the filamen
tous stage is induced by light in the 600 to 700 mlJ. section of the spectrum.
Such photo-induced morphological differentiation can be reversed by light
having a wavelength between 500 to 600 m[J.. Lazaroff has suggested that
allophycocyanin is the receptor pigment for the photochemical activation
process, and that one or more forms of phycoerythrin are involved in the
reversal of photoactivation (185 ) . Although Lazaroff & Vishniac inter
preted their results as showing photochemical production of some growth
substance necessary for morphological development (186 ) , this has recently
been disproved by the observations of Lazaroff (185 ) , indicating that there
is an endogenous induction of morphological development in the absence of
any photochemical stimulus. Nostoc muscorum can complete its life cycle in
the absence of light but this is a very slow process and was noted only after
a few months in the dark.
Fay, Kumar & Fogg (182 ) have also described a simple life cycle for

BLUE-GREEN ALGAE

61

a short filamentous
stage. These mor phol ogical stages are also reflected by changes in the
chemical compo si tion of the cells (134), and in appearance and distribution
of the lamellae and various cytoplasmic granules (187).
Blue-green algae are commonly described as having no sexual stages.
Lazaroff & Vishniac, however, claimed to have seen fusion of two cells
f rom different filaments of N. muscorum to form one large cell (184).
These observations, which would constitute the first evidence of sexuality in
blue-green algae, have not been verified by other investigators. Further evi
dence for sexuality was claimed by Kumar on the basis of studies on ac
quired tolerance of blue-green algae to antibi otic s (188). He obtained
strains of A. nidulans which were resistant to either penicillin or to strepto
m yci n, but not to both, antibiotics ; upon mixing the two strains and letting
them grow for two weeks, he obtained strains which were resistant to both
antibiotics. These apparent recombinants occurred with a frequency of
about 1 in 108 cells. Kumar's work could not be repeated, however, by
Pikilek (189), who could find n o evidence of recombination in A. nidulans.

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

Chlorogloea fritschii involving an endospore stage and

CELL

flexional, or rotatory movement,


c ilia at any stage. The mechanism
causing this movement is not understood, though various hypotheses have
been suggested which involve excretion of mucilaginous material, propaga
tion of rhythmic waves of contraction in the cells, or mechanisms associated
with osmosis or surface tension (51). Many of the filamentous forms,
which show the most rapid movements, also show the presence of pores
through the cell wall (190) through which strands of mucilaginous sub
stances might be secreted. In some cultures of filamentous forms, the move
ments result in aggregation of free filaments to form tangled tufts or layers
(191). Such aggregati on and intertwining has been suggested to have sur
vival value for thermophilic forms in fast-flowing streams, as it prevents
the filaments from being washed away (191 ) . The gliding movements may
be directed either toward or away from a source of light, the direction de
pending on the light wavelength and intensity (192). Action spectra for the
various types of cell movements have been reported by several investigators
(193, 194) ; the pigments responsible for these light-influenced movements
absorb light in the same spectral regions as the photosynthetic pigments.
Most

bl ue

gre e n

algae

MOVEMENTS

show gliding,

though they do not possess flagella or

TOXIC METABOLIC FACTOES


Many of the bloom-producing blue-green algae produce toxic factors
which can re sult in mortality of

fish and domestic animal s (195). The toxin


from Microcystis aeruginosa i s a cycl ic polypeptide consisting of ten amino
acid residues (196 ) . In tests with various animals, this algal product was
far more toxic than the gramicidins or bacitracin, and slightly less toxic
than the poisons extracted from the mushroom Amanita phalloides (196).

62

HOLM-HANSEN

Shilo (197 ) has recently reviewed the algal toxins, including those from
blue-green algae, and stressed their potential application in physiological re
search and as therapeutic agents.
TOLERANCE TO HIGH AND

Low

TEMPERATURES AND TO

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

DESICCATION

Although there are some reports of blue-green algae growing in hot


springs at temperatures over 90 C, recent data indicate that 75 C may be
close to the upper temperature limit for growth of even the most thermo
philic species (8, 198 ) . The biochemical basis for the ability to grow at such
elevated temperatures remains an enigma. That the thermal strains are
highly specialized to survive under conditions of high temperatures is indi
cated by the lower temperature limit for growth in the laboratory. Most
thermal strains will not grow in the laboratory below about 30 C, while
their temperature optima for best growth lies between 40 to 55 C ( 1 99 ) .
The ability o f these thermal strains to grow at such high temperatures indi
cates that the structure and configuration of their protein molecules confer
stability of these macromolecules to elevated temperatures, in contrast to
the proteins in most nonthermal organisms, which undergo denaturation be
tween 30 to 45 C. Apparently there have been no studies on the chemical
and structural properties of the proteins from thermophilic blue-green algae
other than a few investigations on the heat stability of isolated enzymes
(200 ) .
The blue-green algae are also extremely hardy in regard to surviving
the generally deleterious effects of low temperatures ( 20 1 , 202 ) . BIue
green cells flourish on exposed land surfaces and in fresh water in the Ant
arctic, which presents severe climatic conditions (203 ) . Not only do temper
atures drop as low as -88 C, but cells are exposed to severe desiccating
conditions imposed by low atmospheric humidity. They are also subjected to
high light intensities in the summer months, followed by many months of
darkness. The light intensity in the summer, coupled with below-freezing
air temperatures, result in cells undergoing repeated cycles of thawing and
freezing, even during their optimum growing period. Laboratory studies
with blue-green cells isolated from the Antarctic confirm their ability to sur
vive quick and slow freezing, as well as repeated cycles of freezing and
thawing, a process which is very destructive to most eucaryotic cells (202) .
The basis for this marked ability to survive effects of low temperatures
most likely resides in the basic structure and organization of the cellulal'
proteins. Levitt (204) has postulated that frost hardiness in plants is re
lated to increased stability in their proteins brought about by resistance to
ward formation of intermolecular disulphide bonds. He has not, however,
used procaryotic cells in his investigations; it would be interesting to study
the distribution of sulfhydryl and disulphide groups in the proteins of blue
green algae to see if Levitt's theory might correlate with the natural toler
ance of these algae to both low and high temperatures. It should be pointed

63

BLUE-GREEN ALGAE

out that not all blue-green algae show the ability to survive such extreme
temperature condition s Many isolated from the United States are ex
trem ely sensitive to freezing, being unable to survive even one freeze-thaw
cycle (202 ) .
The ability of blue green algae to survive long periods of desiccation is
well known. Some herbarium specimens have been found to be still viable
after 87 years of storage ( 205 ) . Most of these algae which can survive
freezing and thawing can also survive lyophilization (206) . Electron micro
graphs of lyophilized blue-green cells do not reveal any morphological
changes as compared to control specimens which were not lyophilized
(207 ) . This may partially' be due to the lack of any aqueous vacuoles in
these algae. Some strains of Nastae have shown no decline in viability dur
ing five years of storage in the lyophilized state or in their ability to survive
temperatures of 100 C for periods from 10 minutes to one hour. This is in
sharp contrast to the situation with green algae, in which a marked decline
in viability with time was noted in lyophilized cultures, and whose cells
could not survive heating at 100 C for even 10 minutes ( 208 )
One serious limitation of studies dealing with the viability of blue green
algae, as related to such environmental factors as freezing and drying, has
been the difficulty of quantitating the number of viable cells in any suspen
sion. The filamentous forms present special problems in this regard, but
even the unicellular forms can not be counted in agar plates by conventional
methods. Van Baalen has recently reported that he can obtain close to 100
per cent efficiency of plating from single cells of various coccoid blue-green
species by incorporating catalase or F e E DT A in the nutrient agar to coun
teract the deleterious effect of peroxides ( 209, 210 ) . The results vary a
great deal from species to species, however, and the growth of a colony
from a single cell is dependent also on temperature and other factors (21 1 ) .
.

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

SUSCEPTIBILITY TO VIRUSES

Since Safferman & Morris (212) first described a virus causing lysis of
a blue-green alga, other viruses (cyanophages ) have been isolated which
cause lysis in filamentous blue-green algae ( 1 97, 2 1 3, 21 4) . The lytic ability
of such cyanophages is interesting in regard to cell specificity ; those iso
lated by Singh & Singh (215 ) caused lysis of various filamentous forms,
but did not cause lysis of the spores or heterocysts. There has been rela
tively little work on the interactions between the virus and the algal cell.
Wu, Lewin & Werbin (213 ) studied the effect of ultraviolet and visible ra
diation on the lytic ability of a DNA virus on Plectanema bary anum. They
report that when ultraviolet-inactivated virus was mixed with the algal
suspension, a considerable fraction could be photo reactivated by blue or
white light. They concluded that the photoactivation process occurred after
the entrance of the virus into the algal cell, and that the viral-repair mech
anism was not directly associated with photosynthesis.
The ecological significance of cyanophages is not known, though they

64

HOLM-HANSEN

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

may play a role in the sudden death and decay of blue-green cells constitut
ing a water bloom. Such a rapid disintegration of a large population of cells
may also be caused by p arasitic or s aprophyti c bacteria. Shilo ( 197) has re
cently isolated a number of bacteria which cause decomposition of vegeta
tive blue-green cells, though the heterocysts were not affected. There appar
ently is some specificity involved between the bacterial and algal cells, as
the same bacterial isolates did not cause lysis of Chlorella cells.
BLUE-GREEN ALGAE IN SYMBIOTIC RELATIONSHIPS

Blue-green algae occur in many symbiotic relationships such as in lich


ens (216) , in the roots of cycads (217) , and with certain bryophytes and
pteridophytes (59) . In some of these relationships the blue-green alga is
thought to contribute most or all of the reduced carbon and nitrogen com
pounds required for growth by the host cell. D rew & Smith (218, 2 19 ) have
reported that the Nostoe sp. in a lichen thallus excretes up to 40 per cent of
its newly assimilated carbon in the form of glucose, which is then converted
into mannitol by the fungus. This is in sharp contrast to free-living blue
green algae which excrete only small amounts of amino acids, peptides, or
organic acids into the nutrient medium ( 220 ) . The Nostoe cells in the li
chen thallus lack the thick gelatinous sheath that the free-living forms nor
mally possess. It has been shown experimentally that Nostoe cells, shortly
after isolation from the lichen thallus, continue to excrete much of their
photosynthetic product into the medium ; but within 48 hours they develop a
thick sheath and secretion of organic carbon is suppressed. The mechanisms
whereby the fungus seemingly alters the structure and influences the meta
bolic reactions of the blue-green symbiont is completely unknown. It is of
interest in this regard that the blue-green symbiont ( A nabaena sp. ) in the
water fern AzoUa has not been cultured outside the host tissue ( 177 ) ,
which may reflect some structural or biochemical alterations not apparent
by microscopic examinations. As lichen fungi commonly require a supply of
exogenous vitamin Bl and biotin (221 ) , it is likely that the algal symbionts
in lichens also serve as the source of vitamins and possibly other growth
factors for the fungus.
It has been suggested ( 222) that chloroplasts of higher plants may have
originated as independent cells which invaded and became established as in
tracellular symbionts, eventually becoming modified to function as orga
nelles of the host cell. B ecause of many similarities between chloroplasts
and blue-green algae, authors who make such speculations often consider
the early endosymbionts as being blue-green cells. There is considerable
morphological ( 223, 224 ) , enzymatic (225) , and chemical evidence ( 120,
226) in favor of this possibility, although such an evolutionary development
of chloroplasts from blue-green cells would entail many alterations in struc
ture and chemical composition of the endosymbiont. There have been many
studies of the pigmented plastids ( cyanelles ) in the apochlorotic green alga
Glaueocystis nostoehinearum and in the cryptomonad Cyanophora paradoxa.

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

BLUE-GREEN ALGAE

65

The cyanelles of both these forms appear to be similar to blue-green algae


in electron micrographs (223, 227). The main structural difference between
the cyanelles and free-living blue-green cells is the absence in cyanelles of
an outer sheath and cell wall ; only a unit membrane separates it from its
host. If these cyanelles do represent blue-green algae which have lost their
sheaths and cell walls, they are most likely modified in metabolic functions
as well, for they have so far proved to be incapable of independent growth
and division when removed from the host cell (228).

66

HOLM-HANSEN

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

LITERATURE CITED
1. Parson s, D. F., Science, 140, 985-87
( 1 963)
2. Racker, E., Sci. Am., 218, 3 2-39
( 1 968)
3 . San Pietro, A., in Harvesting the
Sun, 49-68.
(San Pietro,
A.,
Greer, F. A., A rmy, T. J., Eds.,
Academic Press, N.Y., 342 pp.,
1967)
4. Branton, D., Park, R . B., J . Ultra
struct. Res., 19, 283-303 ( 1967)
5 . Lang, N., Ann. Rev. Micro bioi., 22
( I n press)
6. Forrest, H. S., Weston, C. R., J.
Phycol., 2, 1 63-64 ( 1 966)
7 . Shields, L. M., Mitchell, C., Drouet,
F., Am. J. Botany, 44, 489-98
( 19 5 7 )
8. Castenholz, R. W., in Environmental
Requirements of B luegreen Al
gae,
55-79.
(Federal
Water

Pollution

Control

Admin.,

Cor

vallis, Oregon, 1 10 pp., 1967)


9. Vinyard, W. C., in Environmental
Requirements of Blue-green Al
gae,
81-85.
( Federal
Water

Pollution
1 0.
11.

12.

13.
1 4.

15.
16.
17.

Control

Admin.,

F.

24.
25.

26,

27.

28.
29.

30.

Cor

vallis, Oregon, 1 10 pp., 1 9 6 7 )


Echlin, P . , Morris, 1., Bioi. Rev., 40,
143-87 ( 19 6 5 )
Lewin, R. A. (Ed.) , Physiology and
Biochemistry of Algae. (Academic
Press, N.Y., 929 pp., 1962)
Desikachary,
J.
V.,
CyanoPhyta.
( Academic Press, N.Y., 686 pp.,
1959)
Gusev, M. V . , Mikrobiologiya, 30,
1 108-28 ( 1 9 6 1 )
Soriano, S . , Lewin, R . A . , Antonie
van
Leeuwenhoek,
31,
66-80
( 1 965)
Bishop, N . 1., Ann. Rev. Plant
Physioi., 17, 1 85-208 ( 1966)
Arnon, D. I., Experientia, 22, 1-15
( 1966)
Bassham, J. A., Jensen, R. G., in
Harvesting the Sun, 79- 1 1 0. (San

Pietro, A' J Greer,

2 1 . Jones, L. W., Myers, J . , J . Phycol.,


1, 7-14 ( 19 6 5 )
2 2 . Fujita, Y., Myers, J., Arch. Biochem.
Biophys., 1 19, 8-15 ( 1 967)
23. Lightbody, J. J., Krogmann, D. W.,

A., Army,

T. J., Eds., Academic Press, N.Y.,


342 pp., 1 9 6 7 )
18. Amesz, J., Duysens, L. N. M., Bio
chim. Biophys. Acta, 64, 261-78
( 1962)
19. Jones, L. W., Myers, J., Plant
Physiol., 39, 938-46 ( 1964)
20. Vernon, L. P., in Harvesting the
Sun,
1 5-28.
( San Pietro, A.,
Greer, F. A., Army, T. J., Eds.,
Academic Press, N.Y., 342 pp.,
1967)

3 1.

32,
33.
34.
35.

Biochim. Biophys. Acta, 120, 5764 ( 1 966)


Lightbody, J. J" Krogmann, D. W"
Biochim. Biophys. Acta, 131, 50815 ( 19 6 7 )
Holton, R. W . , Myers, J., Biochim.
Biophys. Acta, 131, 362-84 ( 1967)
Susor, W. A" Duane, W. c., Krog
mann,
D.
W.,
Record Chern.
Progr. Kresge-Hooker Sci. Lib.,
25, 1 9 7-208 ( 1964)
Susor, W. A., Krogmann, D. W.,
Biochim. Biophys. Acta, 120, 6572 ( 19 6 6 )
Gerhardt, B., S anto, R . , Z. Natur
forschg., 21B, 673-78 ( 1966)
Duane, W. c., Hohl, M . c., Krog
mann, D. W., Biochim. Biophy.
A cta, 109, 108-16 ( 19 6 5 )
Amesz, J . , Vredenberg, W, J . , in
Biochemistry of Chloroplasts, II,
593-600. ( Goodwin, T. W., Ed.,
Academic Press, N,Y., 7 7 6 pp"
1967)
A rnon, D. I., Tsuj imoto, H. Y . ,
McSwain, B . D., Nature, 214,
562-66 ( 1967)
Fujita, Y., Myers, J., Arch. Biochem.
Biophys., lIl, 6 1 9-25 ( 19 6 5 )
Biggins, J . , Plant Physioi., 42, 144756 ( 19 6 7 )
Carr, N. G . , Hallaway, M . , Biochem.
J., 97, 9c-10c ( 1 965)
Fujita, Y., Murano, F., Plant Cell
Physioi,
(Tokyo),
8,
269-82
( 19 6 7 )

3 6 . Fujita, Y.,
Physiot.
( 19 6 6 )

Myers, J., Plant Cell


( Tokyo),
4,
599-606

3 7 . Maclean, F. 1., Fujita, Y . , Forrest,


H. 5., Myers, J., Plant Physiot.,
41, 7 74-79 ( 19 6 6 )
38. Hatfield, D. L., Van Baalen, C.,
Forrest, H. S., Plant Physiol., 36,
240-43 ( 1 9 6 1 )
39.

Maclean, F . 1., Fujita, Y . , Forrest,


H. S., Myers, J., Science, 149,
636-38 ( 1 9 6 5 )

40. HaIld al, p" Physiol. Plantarum, 11,


401-20 ( 1 958)
4 1 . Fuj ita, Y., Hattori, A., Plant Cell
Physiol.
(Tokyo),
3,
209-20
( 1962)
42. Zhevner, V. D., Gusev, M. V"

Shes-

BLUE-GREEN ALGAE
takov , S. V., Mikrobiologiya, 34,
209-15 ( 1 965)
43. Ghosh, A. K., Govindj ee , Biophys. J.,
6, 61 1-19 ( 1966 )
44. Kindel, P., Gibbs, M., Nature, 200,
260-61 ( 1 9 6 3 )
45. Norris, L., Norris, R. E . , Calvin, M . ,
J. Exptl. Botany, 6, 64-74 ( 19 5 5 )
4 6 . Holm-Hansen, 0 . , B row n G . \\T., Jr.,
Plant Cell Physiol. ( Tokyo), 4,

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

299-3 06 ( 1 963)
47. Evans, M. C. W., Buchanan, B. B.,
Arnon, D. 1., Proc. Natl. A cad.
Sci. U.s., 55, 928-34 ( 1 966)
48. Walker, D. A., En deavour 25, 2126 ( 1 966)
49. Linko, P., H o lm Hansen 0., Bass
ham, J. A., Calvin, M., J. Exptl.
Botany, 8, 147-56 ( 1957)
SO. Moses, V., Holm-Hansen, 0., Cal vi n,
M., 1. Bacteriol., 77, 70-78 ( 19 59 )
5 1 . Echl in, P., Sci. Am., 214, 74-81
( 1966)
52. Hattori, A., Plant Cell Physiol.
(Tokyo), 3, 35 5-69 ( 1 962)
53. Gerloff, G. c., Fitzgerald, G. P.,
Skoog, F., Am. !. Botany, 39, 2 6
32 ( 1 9 52)
54. K ratz , W. A., Myers, J., Am. J.
Botany, 42, 282-87 ( 1 9 5 5 )
55. Cobb, H. D . , Jr., Myers, J., A m . !.
Bo tan y, 51, 753-62 ( 1 9 64)
56. A l l en , M . B., Arch. Mikrobiol., ] 7,
34-53 ( 1 95 2 )
57 . McLachlan, J . , Go rham, P. R . , Can.
1. Micro bial. , 8, 1-1 1 ( 1962)
58. Van Baalen. c., Marler, J. E., J.
Gen.
Microbial.,
32,
457-63
( 1 9 63)
59. S tewart, W. D. P., Nitrogen Fixa
,

tion

in

Plants.

(The

Press, Univ. of London

Athlone

168

pp.,

1966)
60. Hattori, A., Myers, J., Plant Cell
Physiol.
(Tokyo),
8,
327-37

( 1967)
61. Hattori,
A.,
M yers,
J.,
Plant
Physiol., 41, 103 1-36 ( 1 966)
62. Fnjita, Y., Ohama, R., Hattori, A.,
Plant Cell Physiol. (Tokyo), 5,
305-14 ( 1 964)
63. S tewart, W. D. P., Science, 158,

1426-32 ( 1 967)
64. S hield s , L. M., Durrell, L. W.,
Botan. Rev., 30, 9 2-12 8 ( 1 964)
6 5 . Stewart, W. D. P., Nature, 211,

603-4 ( 1 967)

66. Holm-Hansen,
0.,
Science,
139,
1 059-60 ( 1 963)
67. Dugdale, R. c., Goering, J. J..
Ryther, J. H., Limllol. Oceanog.,
9, 507-10 ( 1 964)

67

W. D. P., Ann. Botany,


(London), 31, 385-408 ( 1 967)
69. Dilworth, M. J . , Biochim. BioPhys.
Acta, 127, 285-94 ( 1966)
70. B nrris, R. H., Wil son, P. W., Dotan.

68. Stewart,

Gaz., 108, 254-62 ( 1 946)


7 1 . Burris, R. H., Ann. Rev. Plant
Ph:,'sioi., 17, 155-84 ( 1966)
72. Cox, R. M., Fay, 1"., Fogg, G. E.,
Biochim. Biophys. Acta, 88, 2 0 8
10 ( 1964)
73. Fay, P., Cox, R. M., Bio cllim. Bio
Phys. Acta, 143, 5 62-69 ( 1967)
74. Fay, P., Fogg, G . E . , Arch. Mikro
bioi., 42, 3 1 0-2 1 ( 1 9 62)
75. Cox, R. M., Arch. Mikrobiol., 56,
-

193-201 ( 1967)
76. Fay, P., !. Gen. Microbiol., 39, 1 1
20 ( 1 965)
77. Watanabe, A., Yamamoto, Y., Na
ture, 214, 738 ( 1967)
78. Cox, R. M., Arch. Mikrobiol., 53,
263-76 ( 1966)
79. Fay, P., Cox, R. M., Biochim. Bio
phys. Acta, 126, 402-4 ( 1966)
-

80. Carnahan, J. E., Mortenson, L. E.,


Mower, R. F., Cas fie, J. E., Bio
chim. Biophys. Acta, 44, 520-35

( 1960)
81. Brownell, P. F., Nicholas , D. J. D.,
Plant Physiol., 42, 9 1 5-21 ( 1967)
82. Holm-Hansen, 0., Gerloff, G. C.,
Skoog, F., Physiol. Plantarmn, 7,

665-75 ( 1954)
83. Bartels, H., Landwirtsch. !. Schwei::.,
90, 2 7 1 ( 1 9 4 1 )
84. Hood, W., Carr, N. G., Biochim.
Biophys. A cta, 146, 309-1 1 ( 1967)
85. Ch eung, W. Y . , Gibbs, M., Plalll
Physiol., 41, 73 1-37 ( 1 966)
86. Smith, A. J., London, J., Stanier,
R. Y., !. Bacteriol., 94, 972-83

( 1967)
87. Hoare, D. S., Hoare, S. L., Moore,
R. B., !. Gen. Microbiol., 49,
3 5 1-70 ( 1 9 6 7 )
88. Pearce, J., Carr, N. G., Biochem .
I., 105, 45 P ( 1967)
89. Fewson, C. A., AI-Hafidh, M., Gibbs,
M., Plant Physiol., 37, 402-6
( 1962)
90. Antia, N. J., !. Phycol., 3, 81-85
( 1967)
91. Willard, J. M., Schulman, M., Gibbs,
M., Nat ure 206, 195 ( 1965)
92. Van B aalen C., Nature. 206, 193-95
( 1965)
93. Rutter, W. J., Federat-ion Proc., 23,
1248-57 ( 1964)
94. Kornberg, H. L., Krebs, H. A.,
Nature, 179, 988-9 1 ( 19 5 7)
,

HOLM-HANSEN

68

J., Can', N. G., J. Gen.


Microbiol., 49, 3 0 1- 1 4 ( 1 9 6 7 )
96. Hoare, D. S., Moore, R. B., Bio

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

9 5 . Pearce,

chim . Biophys. Acta, 109, 622-25


( 1965)
97. Kiyohara, T . , Fujita, Y., Hattori,
A., Watanabe, A., J. Ge n. Appl.
Microbiol., 6, 1 7 6-82 ( 1960)
0., Booth, C. R.,
98. Holm-Hansen,
Limnol.
Oceal1og.,
II, 5 1 0- 1 9
( 1 966)
99. Coombs, J., Halicki, P. J . , Holm

1 00.

101.
1 02.

103.
1 04.
105.

1 06.
107.

1 08.

Hansen, 0., Volcani, B. E., E.rpt/.


Cel! Res., 47, 302-28 ( 1-9 67)
Webster, D. A., Ha ckett, D. P.,
Plant
Physiol.,
41,
5 9 9-605
( 1 9 66 )
Holton, R W., Myers, J., Science,
142, 234-35 ( 1 9 6 3 )
Webster, G . c . , Frenkel, A. W.,
Plant Physiol., 28, 63-69 ( 1 9 5 3 )
Carr, N. G . , Hallaway, M . . J . Gell.
Microbiol., 39, 3 3 5-44 ( 1 965)
Brown, A. H., Webster, G. C, Am.
J. Botany, 40, 7 5 3-5 8 ( 1 9 5 3 )
Hoch, G . , Owens, 0., K ok , B . , A rch.
Biochem. Biopllys., 1 0 1 , 1 7 1-80
( 1 963)
Harder, R, Z, Botan., 9 , 1 4 5-242
( 1 9 1 7)
Allison, F. E., Hoover, S. R, M o rr i s,
H. J., Botan. Ga::., 98, . 433-62
( 1937)
Van Baalen, c., Botall. Marina, 4,

1 29-39 ( 1 962)
109. Carr, N. G., Exell, G., Flynn, V.,
Hallaway, M., Talukdar, S., Arch.
Biochem. Biophys.,
120, 5 03- 7
( 1967)
1 1 0 . Kratz, W. A.,
Myers, J., Plant
Ph ysio l. , 30, 27 5-80 ( 1 9 5 5 )

1 1 1 . Ca rr, N . G . , Pearce, J., Biochem. J ..


99, 28 P ( 1 966)
1 1 2. Allison. R. K . , Skipper, H . E . , Reid,
M. R, Short, W. A., Hogan, G.
C., J. B ioi. Chem., 204, 1 9 7-205
( 1953)
1 1 3 . Kiyohara, T., F uj i ta, Y., Hattori, A . .
Watanabe,
A.,
J. Gen. Appl.
Microbiol., 8, 1 65-68 ( 1 962)
1 1 4. Ris, H., Singh, R N., J. Biophys.
Bioc/ze", . Cytol" 9, 63-80 ( 19 6 1 )
1 1 5. Leak, L . V., !. Ultrastruct. Res.,
20, 1 90-2 05 ( 1 967)
1 1 6. Jacob, F., Science, 152,
1 470-78
( 1966)
1 1 7 . Cairns, J., J. Mol. BioI., 6, 208-13

(1963)

1 1 8. Zubay, G., \Vatson, M . R, J. Bio


phys. Biochem. Cytol., 5, 5 1 -5 4
( 1 95 9 )
1 1 9. Bonner, J., in The Molecular Biology

of
Development,
1 7.
(Oxfonl
Univ. Press, N.Y., 196 5 )
120. Edelman, M., Swinton, D., S ch iff,
J. A" Epstein, H. T., Zeldin, B.,
Bacterial. Rev., 31, 3 1 5-3 1 ( 1 967)
1 2 1 . Kaye. A . M., Salomon, R, Fr id
lender, B., J. Mol. Bioi., 24, 47983 ( 1 9 6 7 )
1 2 2 . Shields, L. M., Drouet, P., Am. J.
Botany, 49, 547-54 ( 1 962)
1 23. Hortob:igy i, T., Vigassy, J "
Acta
Bioi. Acad. Sci. IIung., 1 8, 1 5 1 60 ( 1967)
1 24. Kraus, M. P., NatHl'e, 2 1 1 , 3 1 0
( 19 6 6 )
1 25. Gerloff, G . C . , Fitzgerald, G. P.,
Skoog, F., Am. J. Botany, 37,
2 1 6- 1 8 ( 1 9 5 0 )
1 :26. Kumar, H. D., A n n . Botany, 27,
723-33 ( 1 9 6 3 )
1 2 7 . Kumar, H, D., Call. J. Botany , 43,
1 5 2 3 -32 ( 1965 )
128. Kumar, H. D., J. Exptl. Botany, 15,
232-50 ( 1 964)
129. Van B aale n, c., Science, 1 49, 70
( 19 6 5 )
1 3 0. Capcsius, I., Ric hte r , G., Z. Natur
forscll., 22B, 204-15 ( 1 967)
1 3 1 . Capesius, 1., Richter, G., Z. Na tur
for-sch., 22B, 876-85 ( 1 96 7 )
1 3 2 . Norton, J., Roth, J. S., J. Bioi.
Chern" 242, 202 9 -3 4 ( 1 967)
1 3 3 . No rton, J., Roth, J. S., Compo Bio
chem. Ph3'siol., 23, 3 6 1-71 ( 1967 )
P.
D.,
Moore,
R. B. ,
1 3 4. Hol oh an ,
Bacterial.
PIOC. (Abstr.) , 1 2 1

( 1967)
1 3 5 . Pa rsons, T . R, Stephens, K, Strick
land, J. D. H., J. Fisheries Res.
Board, Can., 18, 1 00 1 - 1 6 ( 1 9 6 1 )
1 3 6 . Williams, A . E., Burris, R H . , A m .
J. Botany, 39, 340-42 ( 1 9 5 2 )
1 3 7 . R zhan ova ,

G. N., Goryunova, S. V ..

Mikrobiologiya, 34, 268-72 ( 1965)


1 3 8. Miller, M. W.o in Pfizer Handbook
of Microbial Metabolites, 3 3 2-97.
( M cGraw-Hi!1 Book Co., N.Y .,
196 1 )
139 . Holm-Hansen, 0., Lewi n, R A . ,
Physiol.

Plalltarum,

18,

4 1 8-23

( 1 965)
1 40. Parker, P. L, Van Baalen, C.,
M aurer, L., Science, 1 55, 707-8
( 1 967)
1 4 1 . Ho l ton, R. W., Bleck e r, H. H.,
Gnore,
1'.1., Phyto chemistry, 3,
5 9 5-602 (1964)
142. Lev in , E., Lennarz, W. J., moch,
K., Biochim. Biophys. A c ta, 84,
47 1-74 ( 1 9 6 4 )

1 43 . Nichols, B . W., H ar ri s , R V., James,

69

BLUE-GREEN ALGAE
A. T., Biochem. Biophys. Res.
COnlnlun., 20, 256-62 ( 1 965)
1 44. Nichols,

B . W., Wood, B . J. B . ,
Lipids, 3 , 46-50 ( 1 9 68)
1 45 . Carr, N. G., Biochim. Biophys. Acta,
1 20, 308-10 ( 1966)
1 46. Giesy, R. M., Ani. J. Botany, 51,

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

388-96 ( 1 964)
1 47. Tsusue, Y., Yamakawa, T., J. Bio
chem. (Tokyo), 58, 5 87-94 ( 1 965 )
1 48. Moore, B. G., Tischer, R. G., Call.
J. Microbiol., 1 1, 877-85 ( 1 9 6 5 )
149. Drews, G . , Gollwitzer, W., Arch.
Mikrobiol., 51, 1 79-85 ( 1 965)
1 50. F rank, H., L efort, M., Martin, H.
R., Biochem. Biophys. Res. Com
mun., 7, 322-25 ( 1962)
1 5 1 . Holm- Han sen,
0.,
Prasad,
R.,
Lewin, R. A., Phyc% gia, 5, 1-14
( 1 965 )
1 52. Punnett, T., Derrenbacker, E. c., J.
Gen.
Microbiol.,
44,
105-14

( 1 966)
1 5 3 . Moore, B. G., Tischer, R. G., Sci
ence, 145, 586-87 ( 1 964)
1 54. 6hEocha, C., in Chemistry and Bio
chemistry of Plant Pigments, 1 7596. ( Goodwin, T. W., Ed., A ca
demic P ress, N.Y., 583 pp., 1965 )
1 5 5 . Hattori, A., Fuj ita , Y., J. Biochem.
( Tokyo), 46, 633-44 ( 1 959)
1 5 6. Chapman, D. J., Cole, W . J., Siegel
man, H. W., Biochem. J., 105,

903-5

( 1 967)

1 5 7. Gassner, E. B . , Plant Physiol., 37,


63 7-39 ( 1 9 6 2 )
1 58. Govindjee, Cederstrand, C . , Rabino
witch, E., Science, 134, 3 9 1-92

( 1961)
1 5 9. Hertzberg, S., Jensen, S. L . , Phyto
chemistry, 5, 5 5 7-63 ( 1 966)

1 60. Hertzberg, S., Jensen, S. L., PhSi'o


chemistry, 5, 565 -70 ( 1 9 66)
1 6 1 . Hertzberg, S . , Jen se n , S . L., Phyto
chemistry, 6, 1 1 1 9-2 6 ( 1967)
162. Goodwin, T. W., i n Chemistry and
Biochemistry of Plant Pigments,
127-42. (Goodwin, T. W., Ed.,
Academic Press, N.Y., 583 pp.,
1 96 5 )
163. Gerloff, G. C., Fitzgerald, G. P.,
Skoog, F., Am. J. Botany, 37,

83 5-40 ( 1950)
1 64. Allen, M . B . , Arnon, D. 1., Pla n t
Physiol., 30, 366-72 ( 19 5 5 )
1 6 5 . Taha, E . E. M . , Elrefai, A. E. M .
H., Arch. Mikrobiol., 43, 67-75
( 1962)
1 66. Eyster, C., Nature, 170, 755 ( 1 9 5 2 )

167. Wolfe,

M., Ann. Botany (London ) ,

1 8 , 309-25

1 68. Holm-Hansen,

( 1954)

0., A Study o f Major

and l1Jillor Ele1ll e nt Requirements


in the Nutri/ion of Blue-Green
A lgae. (Doctoral thesis, Univ. of
Wisconsin, Madison, Wise .. 1 9 5 4 )
1 69 . Allen, M . B . , Arnon, D . 1., Physiol.
Plantarum, 8, 653-60 ( 1 9 5 5 )
1 7 0. Johnson, G . V . , Mayeux, P. A.,
Evans, H. J., Plant Physiol., 4 1 ,
8 5 2-5 5 ( 1966)
1 7 1 . Bo wen, H. J. M., Trace Elements in

Biochemistry. (Acadeluic Press,


N.Y., 2 4 1 pp., 1 9 6 6 )
1 72. Lamb, C. A., B en tley, O. G., Beattie,
J. M., (Eds . ) , Trace Elements.
(Academic Press, N.Y., 4 1 0 pp.,
1 958)
173. Bertels, H., Arch.
1 5 5-86 ( 1 940)

Mikrobiol.,

11,

174. Pintner, 1. J., Provasoli, L., J. Gen.


Microbiol., 18, 190-97 ( 1 9 5 8 )
1 7 5 . Fogg, G . E., Bacterial. Rev., 20,
1 48-65 ( 1 9 5 6 )
1 7 6 . Clendenning, K. A., B ro wn, T. E.,
Eyster, H . C., Can. J. Botany,
34, 943-66 ( 1 956)
1 7 7 . Lang, N. J., J. Phycol., 1, 127-34

( 1965)

1 78. Fr it sch, F . E . , Proc. Linnean Soc.


London, 162, 1 94-2 1 1 ( 1 9 5 1 )
1 79. Wolk, C. P., Nature, 205, 20 1-2
( 1 96 5 )
180. Fay, P . , Walsby, A . E., Nature, 209,
94-95 ( 1 966)
181. Wolk, C. P., Proc. Natl. A cad. Sci.
U.S., 57, 1246-5 1 ( 1967)
1 82. Fay, P., Kumar, H. D., Fogg, G. E.,
J. Gell. Microbiol., 35, 35 1-60

( 1964)
183. Laza roff, N., Schiff, J., Science, 137,
603-4 ( 1 962 )
184. Lazaroff, N., Vishniac, W., J. Gen.
J'vIic,obiol., 28, 203-10 ( 1 962)
185. Lazaroff, N., J, Phycol., 2, 7-17
( 1966 )
1 8 6.

Lazaroff, N., Vish ni ac, W., J. Cen.


Microbiol., 25, 365-74 ( 1 9 6 1 )

A ., Whitton, B . A . , Arch.
Mikrobiol., 57, 1 5 5-80 ( 1967)

1 8 7 . Peat,

1 8 8 . Kumar, H. D . , Nature, 196, 1 5 2 122 ( 1962)

189. Pikalek, P., Nature, 215, 666-67


( 1967)
190. P an kra tz, H. 5., Bowen, C. c., Am.
J. Botany, 50, 3 8 7-9 9 ( 1 963)
191. Castenholz, R. W . , Nature, 215,
1 285-86 ( 1 967)
192. Bendix, S., in Comparative Bio
chemistry of Photoreactive Sys
tems, 107-27. (Allen, M. E., Ed.,
Academic Press, KY., 437 pp.,

1 960)

70

HOLM-HANSEN

Annu. Rev. Microbiol. 1968.22:47-70. Downloaded from www.annualreviews.org


by WIB6242 - Universitaets- und Landesbibliothek Duesseldorf on 11/04/13. For personal use only.

193. Nultsch, W.,


( 1962)

Planta,

58,

647-63

194. Halldal, P., Phytochem. Photobiol.,


6, 445-60 ( 1 967)
195. Flint, E. A., New Zealand Vet. I.,
14, 181-85 ( 1 966)
196. Konst, H., McKercher, P. D., Gor
ham,
P.
R, Robertson , A.,
Howel l, J., Can. 1. Compo Med.
Vet. Sci., 29, 22 1-28 ( 1965)

1 9 7 . Shi1o, M., Bacteriol. Rev., 31, 1809 3 ( 19 6 7 )


198. Brock, T. D., Science, US, 1012-19
( 1 9 67 )
199. Peary, J. A., Castenholz, R. W.,
Nature, 202, 72 0-2 1 ( 1964)
200. Marre, E., in Physiology and. Bio
chemistry
of
Algae,
541-50.
(Lewin, R. A., Ed., Academic
Press, N.Y., 929 pp., 1962)
201. Whitton, B . A., Brit. Phycol. Bull.,
2, 1 7 7-78 ( 1962)
202. Holm-Hanse<!, 0., Physiol. Plan
tamm, 16, 530-40 ( 1963)

203. Holm-Hansen, 0., Phyco logia, 4, 4351 ( 1964)


204. Levitt, J., 1. Theoret. Bioi., 3, 3 5 59 1 ( 1962)

205. Lipman, C. B., Bull. Torrey Bo tan.

Club, 68, 664-66 ( 19 4 1 )


206. Holm-Hansen, 0., Can. I. Botany,
42, 127-37 ( 1 964)

207. Roge rs, T. D., Scholes, V. E.,


Schlichting, H. E., Jr., f. Ph)'col.,
3, 16 1-65 ( 1967)
208. Holm-Hansen, 0., Cryobiology, 4,

1 7-23 ( 1 967)
209. Van Baalen, C., 1. Phycoi., 1, 1922 ( 1 965 )
2 1 0 . Marler, J. E.,
Phycol.,

1,

Van Baalen, c., f.


180-85 ( 1 965)

2 1 1 . Van B aal en , C., 1. Phycol., 3, 15457 ( 1967)


212. Safferman, R. 5., Morris, M. E ,
Science, 140, 679-80 ( 1 963)
2 1 3. Wu, J. H., Lewin. R. A., Werbin,
H., Virology, 31, 657-64 ( 1967)
2 14. Safferman, R 5., M orris , M . E., 1.
Bacteriol., 88, 77 1-75 ( 1964)
2 1 5. Singh, R N., Singh, P. K., Nature,
216, 1 020-2 1 ( 1967)
2 1 6. Ahmadj ian, V., Phyc% gia, 6, 12760 ( 1967)

2 1 7. Watanabe,

A.,

Kiyohara,

Microalgae
and
Bacteria, 189-96.

T.,

in

Photosynthetic

(Japanese Soc.
Plant Physiol., Eds., Univ. Tokyo
Press, Tokyo, 1 963)
2 1 8. Drew, E . A., Smith, D. C., New
Phytologist, 66, 379-88 ( 1967)
2 1 9. Drew, E . A., Smith, D. C., New
Phytologist, 66, 389-400 ( 1967)
220. Whitton, B. A., f. Gen. Micro biol. ,
40, 1-1 1 ( 1965)

221. Bednar, T. W., Holm-Hansen, 0.,


Plant Cell Physiol. ( Tokyo), 5,

297-303 ( 1964)
222. Sagan, L., f. Theoret. BioI., 14, 22574 ( 1967)
223. Hall, W. T., Claus, G., f. Phycol.,
3, 3 7-5 1 ( 19 67)
224. Ris, H., Plaut, W., J. Cell Bioi., 13,
3 83-9 1 ( 1962)
225. Arnon, D. 1., Science, 149, 1 460-69

( 1965)

226. Wildman, S. G., in B io chem istry of


Chloroplasts, II, 295-3 19. ( Good
win, T. W., Ed., Academic Press,
N.Y., 776 pp., 1967)
227. Hall, W. T., Claus, G., J. Cell B ioI.,
19, 5 5 1-63 ( 1963)
228. Echlin, P., Brit. Phycol. Bull., 3,
225-39 ( 1 967)

Вам также может понравиться