You are on page 1of 79

FISIOLOGIA VEGETAL COMPLEMENTAR

(2013/2014)

Assimilation of Mineral Nutrients
Plant Physiology, Fifth Edition (2010), L. Taiz & E. Zeiger
Chapter 12 - Assimilation of Mineral Nutrients
Nutrient assimilation is the often energy-requiring
process by which plant incorporate inorganic nutrients
into the carbon constituents necessary for growth and
development.
Plant Physiology, Fifth Edition (2010), L. Taiz & E. Zeiger
Chapter 12 - Assimilation of Mineral Nutrients
What is nutrient assimilation?!....
The assimilation of sulfate (SO) into amino acid cysteine
via the two pathways foud in plants consumes about 14 ATPs.
Nitrogen in the Environment
Nitrogen in the Environment

When nitrogen is fixed into ammonia (NH) or Nitrate (NO), it passes through several organic
or inorganic forms before it eventually returns to molecular nitrogen (N) (Figure 12.1).
UNSSIMILATED AMMONIUM OR NITRATE MAY BE DANGEROUS


Ammonium, if accumulated to high levels in living tissues, is toxic to both
plants and animals. Ammonium dissipates transmembrane proton
gradients that are required for both photosynthetic and respiratory
electron transport and for sequestring metabolites in the vacuole.


In contraste with ammonium, plants can store high levels of nitrate, or
they can translocate it from tissue to tissue without deleterious effect.
If livestock or humans consume plant material that his high in nitrate,
they may suffer methemoglobinemia, a desease in which the liver
reduces nitrate to nitrite, which combines with hemoglobin and renders
hemoglobin unable to bind oxygen. Humans or other animals may also
convert nitrate into nitrosamines, which are potent carcinogens.
At high concentrations, ammonium (NH) is toxic to living tissues,
but nitrate (NO) can be safely stored and translocated in plant
tissues (Figure 12.2).
Nitrate assimilation

Plant roots actively absorb nitrate, then reduce it to nitrite (NO) in
the cytosol (Figure 12.3).

Nitrate, light, and carbohydrates affect the transcription and
translation of nitrate redutase (Figure 12.4).

Darkness and Mg can inactivate nitrate redutase. Such
inactivation is faster than regulation by reduced synthesis or
degradation of the enzyme.

In chloroplasts or plastids, the enzyme nitrite redutase reduces
nitrite to ammonium (Figure 12.5).

Both roots and shoots assimilate nitrate (Figure 12.6).
Plant roots actively absorb nitrate from the soil solution via several
low- and high-affinity nitrate-proton cotransporters.

Plants eventualy assimilate most of this nitrate into organic nitrogen
compounds. The first step of this process is the reduction of nitrate
to nitrite in the cytosol, a reaction that involves the transfer of two
electrons. The enzyme NITRATE REDUCTASE catalyses this reaction:




Where NAD(P)H indicates NADH (mostly used, namely in green
tissues) or NADPH (some nitrate reductase forms, mostly from
nongreen tissues use NADH or NADPH as donor of electrons).
The nitrate reductases of higher plants are
composed of two identical subunits, each
containing three prosthetic groups: flavin
adenine nucleotide (FAD); heme, and a
molybdenium atom complexed to an organic
molecule called pterin
Plant roots actively absorb nitrate, then reduce it to nitrite (NO)
in the cytosol (Figure 12.3).

Light, carbohydrate levels, and other environmental factors stimulate a protein
phosphatase that dephosphorylates a key serine residue in hinge 1 region of nitrate
reductase (between the molybdenium complex and heme-binding domains; Fig. 12.3)
and thereby activates the enzyme.
Many factors regulate nitrate reductase

Nitrate, light, and carbohydrates affect the transcription and
translation of nitrate redutase (Figure 12.4).

NITRITE REDUCTASE CONVERTS NITRITE TO AMMONIUM

Darkness and Mg can inactivate nitrate redutase. Such
inactivation is faster than regulation by reduced synthesis or
degradation of the enzyme.

In chloroplasts or plastids, the enzyme nitrite redutase reduces
nitrite to ammonium (Figure 12.5).
Both roots and
shoots assimilate
nitrate (Figure 12.6).
Ammonium Assimilation

Plant cells avoid ammonium toxicity by rapidly converting
ammonium into amino acids (Figure 12.7).

The primary pathway of this conversion involves the sequential actions
of Glutamine Synthetase and Glutamate Synthetase

Nitrogen is incorporated into other amino acids via transamination
reactions involving glutamine and glutamate.

The amino acid asparagine is a key compound for nitrogen transport
and storage.


(A) The GS-GOGAT pathway that forms glutamine and glutamate.
A reduced cofactor is required for the reaction: ferredoxin in
green leaves and NADH in nonphotosynthetic tissue.

Figure 12.7. Structure and pathways of compounds involved in ammonium
metabolism. Ammonium can be assimilated by one of several processes:

( A )
( B )

(B) The GDH pathway that forms glutamate using NADH or
NADPH as a reductant.
ALTERNATIVE PATWAY FOR AMMONIUM ASSIMILATION
Transamination reactions transfer nitrogen
Figure 12.7C.


Asparagine and glutamine link carbon and
nitrogen metabolism (Figure 12.7D
Figure 12.7. Structure and pathways of compounds involved in ammonium
metabolism. Ammonium can be assimilated by one of several processes:
( C )
( D )
( C ) Transfer of the amino
group from glutamate to
oxaloacetate to form
aspartate (catalized by
aspartate aminotransferase)
( D ) Synthesis of asparagine
by transfer of an amino acid
group from glutamine to
aspartate (catalyzed by
asparagine synthetase).
Amino Acid Biosynthesis




The carbon skeleton for amino acids derives from intermediates
of glycoysis and the citric acid cycle ( Figure 12.8).

The carbon skeleton for amino acids derives from intermediates
of glycoysis and the citric acid cycle ( Figure 12.8).
NH + Glutamato Glutamina (GS)
Glutamine + 2-Oxoglutarate 2 Glutamates (GOGAT)
NH + 2-oxoglutarate Glutamate (GDH)
Glutamate + Oxaloacetate Aspartate + 2-Oxoglutarate(Asp-AT)
Glutamine + Aspartate Asparagine + Glutamate (AS)

Biological Nitrogen Fixation

Some Free-living and symbiotic bacteria fix nitrogen
(N) and convert it into ammonium.

Most common type of symbiotic associations (Table 12.2):

- Leguminosae rhizobia
- Woody plant species (actinorhizal plants) Frankia sp.)
- Fern Azolla sp. cyanobacteria Nostoc sp. and Anabaena sp.

Biological nitrogen fixation accounts for most of the ammonium
formed from the atmospheric N (Figure 12.1; Table 12.2).

Nitrogen fixation requires anaerobic conditions


Oxygen being a strong electron acceptor, can damage the
nitrogen enzymes that catalyze the reactions of nitrogen
fixation and irreversibly inactivate them, so nitrogen must be
fixed under anaerobic contitions.


Nitrogen fixing bacteria either functions under natural anaerobic
conditions or creates an internal, local anaerobic environment in
presence of oxygen.
Several types of nitrogen-fixing bacteria form symbiotic associations with higher
plants (Figures 12.9, 12.10; Table 12.3).
Symbiotic nitrogen-fixing prokaryotes function within specialized structures formed
by plant host (Figures 12.9, 12.10).
Heterocysts formation occurs when cyanobacteria
are deprived of NH. These cells lack photosystem II,
the oxygen-producing photosystem of chloroplasts.
Free-living bacteria that are capable of fixing
nitrogen are aerobic, facultative, or anaerobic


Aerobic nitrogen-fixing bacteria, such as Azotobacter are thought to
mantain a low oxygen concentration (microaerobic conditions)
through their high levels of respiration. Others, such Gloeothece,
evolve O during the day and fix nitrogen during the night.

Facultative organisms, which are able to grow under both aerobic,
and anaerobic conditions, generally fix nitrogen only under
anaerobic conditions.

Anaerobic nitrogen-fixing bacteria grows in the absence of oxygen,
thus it is not a problem.

Symbiotic nitrogen fixation occurs in specialized structures

Symbiotic nitrogen-fixing prokaryotes dwell within nodules, the
special organs of plant host that enclose nitrogen-fixing bacteria.
Establishing symbiosis requires na exchange of signals

Symbiosis between legumes and rhizobia is not obligatory for both
organisms.

Under nitrogen-limited conditions, however, the symbionts seek
each other out through na elaborate exchange of signals involving
specific genes in both host and the symbionts.

Plant genes are called nodulin genes (Nod). Rhizobial genes are
called nodulation (nod) genes.

nodA, nodB, nodC found in all rhizobial strains
nodP, nodQ, nodH, nodF, nodF, nodE, nodL host specific nod genes
(differ among the rhizobial species and determine the plant that can be infected)

Only nodD is constitutively expressed and its protein product (NodD) regulates
the transcription of other nod genes.
Nod factors produced by bacteria act as signals for
symbiosis

The nod genes, which nodD activates, code for nodulation proteins,
most of which are involved in the biosynthesis of Nod factors.


Nod factors are lipochitin oligosaccharide signal molecules, all of
which have a chitin -14-linked N-acetyl-D-glucosamine
backbone (varying in lenght from three to six sugars units) and a
fatty acid chain on C-2 position of the non-reducing sugar (Figure
12.11)
The symbiotic relationship is initiated by the migration of the nitrogen-fixing bacteria
toward the roots, which is mediated by chemical attractants secreted by the roots.

Attractants activate the rhizobial NodD protein which then induces the biosynthesis
of Nod factors that act as signals for symbiosis (Figure 12.11).
Host specific nod genes (nodP, nodQ, nodH, nodF, nodF, nodE,
nodL) that vary among rhizobial species are involved in the
modification of the fatty acyl chain or the addition of groups
important in determining host specificity.
Nod factors induce root hair curling, sequestration of rhizobia, cell
wall degradation, and bacterial access to the root hair plasma
membrane, from which an infection thread forms (Figure 12.12).

Filled with proliferating rhizobia, the infection thread elongates
through root tissue in the direction of the developing nodule, which
arises from cortical cells (Figure 12.12).

In response to a signal from the plant, the bacteria in the nodule
stop dividing and differentiate into nitrogen-fixing bacteroids.
Nodule formation involves phytohormones

Two processes infection and nodule organogenesis occur
simultaneously during nodule formation.
Deeper into root cortex, near the xylem, cortical cells
dedifferentiate and start dividing, forming a distinct area
within the cortex, called nodule primordium, from which
the nodule will develop. The nodule primordia form
opposit the protoxylem poles of the root vascular bundle.

The infection thread filled with proliferating rhizobia elongates
through the root hair and cortical cell layers, in the direction of the
nodule primordium.

When infection reaches specialized cells within the nodule, its
fuses with the plasma membrane of the host cell, releasing bacterial
cells that are packaged in a membrane derived from the host cell
plasma membrane.

Branching of the infection thread inside the nodules enables the
infection of may cells.
Under an undetermined signal from plant, bacteria stop
dividing and began to enlarge and to differentiate into
nitrogen-fixing endosymbiotic organeles called bacteroids.
The membrane surrounding the bacteroids is called
peribacteroid membrane.

The nodule as a whole develops such features as a
vascular system which enables the exchange of fixed
nitrogen and nutrients between the bacteria and host
plant). A layer of cells is formed to exclude the O from the
root nodule interior.
The nitrogenase enzyme complex fixes N
The reduction of N to NH is catalyzed by the nitrogenase enzyme
complex (Figure 12.13).

The Fe protein ( 30 to 72 kDa) participates in the conversion of N to NH. The Fe protein is
inactivated by O
The MoFe protein (180 to 235 kDa) is also inactivated by O can reduce numerous
substracts (Table 12.4).
Fixed nitrogen is transported as amides or ureides (Figure 12.14).

The symbiontic nitrogen-fixing prokaryotes release ammonia (NH) that, to avoid
toxicity, must be rapidly converted into organic forms in the root nodules before being
transported to the shoot via xylem (Figure 12.14).
Biological Nitrogen Fixation

Biological nitrogen fixation accounts for most of the ammonium
formed from the atmospheric N (Figure 12.1; Table 12.2).

Several types of nitrogen-fixing bacteria form symbiotic associations
with higher plants (Figures 12.9, 12.10; Table 12.3).

Nitrogen fixation requires anaerobic conditions.

Symbiotic nitrogen-fixing prokaryotes function within specialized
structures formed by plant host (Figures 12.9, 12.10).

The symbiotic relationship is initiated by the migration of the
nitrogen-fixing bacteria toward the roots, which is mediated by
chemical attractants secreted by the roots.
Attractants activate the rhizobial NodD protein which then induces the
biosynthesis of Nod factors that act as signals for symbiosis (Figure 12.11).

Nod factors induce root hair curling, sequestration of rhizobia, cell wall
degradation, and bacterial access to the root hair plasma membrane, from
which an infection thread forms (Figure 12.12).

Filled with proliferating rhizobia, the infection thread elongates through
root tissue in the direction of the developing nodule, which arises from
cortical cells (Figure 12.12).

In response to a signal from the plant, the bacteria in the nodule stop
dividing and differentiate into nitrogen-fixing bacteroids.

The reduction of N to NH is catalyzed by the nitrogenase enzyme
complex (Figure 12.13).

Fixed nitrogen is transported as amides or ureides (Figure 12.14).
Sulfur Assimilation




Most assimilated sulfur derives from sulfate (SO) absorved from
the soil solution, but plants can also metabolize gaseous sulfur dioxide
(SO) entering via stomata.

Synthesis of sulfur-containing organic compounds begins with the
reduction of sulfate to the amino acid cysteine (Figure 12.15).

Sulfate is assimilated in leaves and exported as glutathione via the
phloem to growing sites.

Sulfur is one of the most versatile elements in living organisms.



- As nitrogen, sulfur shares multiple stable oxidation states.

- Clusters sulfur-iron are involved in electron transport systems .

- Sulfur participates in some catalytic sites of enzymes and co-
enzymes such as urease and coenzyme A.

- Dissulfide bridges in proteins play important structural and
regulatory roles.

- Some secondary metabolites contains sulfur (eg. Rhizobial Nod
factors, alliin (antiseptic) in garlic, sulforaphane (anticarcinogenic)
in brocoli.
Sulfur is absorbed in plants, mostly in the form of sulfate



SO is absorbed via an H- SO symporter
(The natural origin of SO is the weathering of rock material)


SO (from the polution) can enter the in the gas form though the
stomata of the plant and be metabolized. However, extensive
exposure (8h) to SO (0.3 ppm) causes extensive tissue damage
because formation of HSO.

In the gas phase, SO reacts with HO
-
and O
2
to form sulfur trioxide
(SO
3
). SO
3
dissolves in water to become sulfuric acid (H
2
O
4
).
1st The first step in the synthesis of sulfur-containing organic
compounds is the reduction of sulfate to the aminoacid
cysteine (Figure 12.15).
Sulfate is very stable and thus needs to be activated before any
subsequent reactions may proceed. Activation begins with the
reaction between sulfate and ATP to form 5-adenylylsulfate
(which is sometimes referred to as adenosine-5-phosphosulfate
and thus is abreviated APS) and pyrophosphate (PPi) (Figure
12.15)
ATP sulfurylase
The enzyme that catalyzes the reaction 12.11, ATP sulfurylase,
has two forms: The major one is found in plastids, and a minor
one is found in cytoplasm. The activation reaction is
energetically unfavorable. To drive this reaction forward, the
products APS and PPi must be converted immediately to other
compounds. PPi is hydrolyzed to inorganic phosphate (Pi) by
inorganic pyrophosphatase according to the following reaction:
inorganic pyrophosphatase
Figure 12.15 Structure and pathways of compounds involved in sulfur assimilation. The enzyme ATP
sulfurylase cleaves pyrophosohate from ATP and replaces it with sulfate. Sulfide is produced from APS
through reactions involving reduction by glutathione and ferredoxin. The sulfide reacts with O-
acetylserine to form cysteine. Fd, ferredoxin; GSA, glutathione, reduced; GSSG, glutathione , oxidized.
Synthesis of sulfur-containing organic compounds begins with the
reduction of sulfate to the amino acid cysteine (Figure 12.15).

eg. Choline,
brassinosteroids,
flavonoids, gallic
acid glucoside,
glucosinolates,
peptides,
polyssacharides,
The other product, APS., is rapidly reduced or phosphorylated.

Reduction is the dominant pathway.

The reduction of APS is a multistep process that occurs exclusively
in the plastids. First, APS reductase (after APS sulfotransferase)
transfers two electrons, aparently from reduced glutathione (GSH),
to produce sulfite (SO):
Where GSSG stands for oxidized glutathione (The SH in GSH
and SS in GSSG stand for S-H and S-S bonds respectively).
APS sulfotransferase APS reductase
2nd Sulfite reductase transfers six electons from ferredoxin
(Fdred) to produce sulfide (S)




The resultant sulfide then reacts with O-acetylserine (OAS) to
form cystein and acetate The OAS that reacts with S is
formed in a reaction catalyzed by serine acetyltransferase:
The reaction that produces cysteine and acetate is catalyzed by
OAS(thio)lyase:
serine acetyltransferase
OAS(thio)lyase
Figure 12.15 Structure and pathways of compounds involved in sulfur assimilation. The enzyme ATP
sulfurylase cleaves pyrophosohate from ATP and replaces it with sulfate. Sulfide is produced from APS
through reactions involving reduction by glutathione and ferredoxin. The sulfide reacts with O-
acetylserine to form cysteine. Fd, ferredoxin; GSA, glutathione, reduced; GSSG, glutathione , oxidized.
Synthesis of sulfur-containing organic compounds begins with the
reduction of sulfate to the amino acid cysteine (Figure 12.15).

eg. Choline,
brassinosteroids,
flavonoids, gallic
acid glucoside,
glucosinolates,
peptides,
polyssacharides,

The phosphorilation of APS, localized in the cytosol, is
the alternative pathway. First, APS kinase catalizes a
reaction of APS with ATP to form 3-phosphoadenosine-
5phosphosulfate (PAPS):





Sulfotransferases then may transfer the sulphate group
from PAPS to various compounds, including choline,
brassinosteroids, flavonoids, gallic acid glucoside,
glucosinolates, peptides, and polysaccharides.
Sulfur assimilation occurs mostly in leaves

Sulfate is assimilated in leaves and exported as
glutathione via the phloem to growing sites.

The reduction of sulfate to cysteine changes the
oxidation number of sulfur from +6 to -2, thus
entailing the transfer of 8 electrons. Glutathione,
ferredoxin, NAD(P)H, or O-acetylserine may serve
as electron donors at various steps of the pathway.
Leaves are generally much more active than roots in sulfur
assimilation. The photosynthesis provides reduced ferredoxin and
photorespiration generates serine, that may stimulate the
production of O-acetylserine.

Methionine, the other sulfur-containing amino acid, found in
proteins, is synthesized in plastids from cysteine.

Phosphate Assimilation

Roots absorve phosphate (HPO) from the soil solution, via an H-
HPO symporter.

The HPO assimilation occurs with the formation of ATP, the main
entrance in the asssimilatory pathways, being added to the second
phosphate group in ADP.

From ATP, the phosphate group may be transferred to many
different carbon compounds in plant cells, including sugar
phosphates, phospholipids, and nucleotides.

In mitochondria, the energy for ATP synthesis derives from the
oxidation of NADH by oxidative phosphorilation. In chloroplastes the
synthesis of ATP is driven by light dependent phosphorilation. In
cytosol, reactions as glycolysis also assimilate phosphate.

Cation assimilation


Plants assimilate:

(a) macronutrient cations such as potassium, magnesium,
and calcium;
(b) Micronutrient cations such as, copper, iron,
manganese, cobalt, sodium, and zinc.

All of them form complexes with organic compounds by
noncovalent bounds of two types:
(a) coordination bonds; (b) electrostatic bonds.
Cation Assimilation

Polyvalent cations form coordination bonds with carbon molecules
(Figure 12.16).

Monovalent cations form electrostatic bonds with the carboxylate
groups (Figure 12.17).

To absorb sufficient amounts of insoluble ferric iron (Fe) from the
soil solution, several mechanisms are used (Figure 12.18).

Once in leaves, iron undergoes na important assimilatory reaction
(Figure 12.19).

To limit the free radical damage that free iron can cause, plant cells
may store surplus iron as phytoferritin.
Coordination bonds typically form between carbon molecules and
polyvalente cations (copper, zinc, iron, and magnesium and often
calcium).


In coordination complexes several oxygen or nitrogen atoms of a
carbon compound donate unshared electrons to form a bond with
a cation nutrient and consequently neutralize its positive charge.


As examples, figure 12.16 shows the complexes between:
copper and tartaric acid (A); magnesium and chlorophyll a (B),
Calcium and polygalacturonic acid (C).
Polyvalent cations form coordination bonds with carbon
molecules (Figure 12.16).

In general, cations such Mg and Ca are assimilated by the
formation of both coordination complexes and electrostatic bonds
with amino acids, phospholipids, and other negatively charged
molecules.
Electrostatic bonds form because of the attraction of a positively
charged cation for a negatively charged group on a carbon
compound (eg. COO).

Monovalent cations, like K, can form electrostatic bonds with the
carboxylic groups of many organic acids (figure 12.17A)
Much of K remains free, however, in the cytosol and vacuole and
functions in osmotic regulation and enzyme activation

Divalent cations such Ca form electrostatic bonds with
carboxylic groups of polygalacturonic acid, and pectates (figure
12.17B).
Monovalent cations form electrostatic bonds with the carboxylate
groups (Figure 12.17).

Roots modify the rhizosphere to acquire iron
Iron is important in ion-sulfur proteins and as a catalist in
enzyme-mediated redox reactions, such as those of the
nitrogen metabolism.


In soil, iron is found, primarily as ferric iron Fe in oxides
such as Fe(OH), Fe(OH), and Fe(OH). At neutral pH,
ferric iron is highly insoluble
Roots have developed several mechanisms that increase the
iron solubility and availability (figure 12.18).


Such mechanisms include:

soil acidification, which increases the solubility of ferric ion,

reduction of ferric iron Fe to the more soluble ferrous
form (Fe);

release of compounds (chelators such as malic acid, citric
acid, phenolics etc.) that form stable, soluble complexes with
iron.

To absorb sufficient amounts of insoluble ferric iron (Fe) from the soil
solution, several mechanisms are used (Figure 12.18).


The extrusion of H by the roots during the assimilation of cations,
particularly NH acidify the soil around it.

Besides H, root plants release also organic acids, such as malic acid and
citric acid (chelators) that enhance the iron and phosphate availability.

Iron deficiency stimulates the extrusion of H by roots and the activity
of the iron-chelate redutase an enzyme that reduces Fe to Fe, with
cytosolic NADH or NADPH serving as the electron donor (figure 12.18A).

Grasses produce a special classe of chelators, called siderophores
(special amino acids that are not found in proteins, such as mucigenic
acid) whose production and release by the roots increase under iron
deficiency in the soil.

Root cells of grasses have Fe -siderophores transport systems in their
plasma membranes that bring the chelat into the cytoplasm.
Iron forms complexes with carbon and phosphate
After absorbing the iron-chelate, root cells oxidize the iron to Fe
and translocate it to the leaves as electrostatic complexes, such as
Fe-citrate or Fe-nicotinamine.

In leaves, iron undergoes an important assimilatory reaction,
catalyzed by the enzyme ferrochelatase, through which it is
inserted into the porfyrin precursor of heme groups found
cytocromes located in chloroplasts and mitocondria (Figure 12.19)

Free iron may interact with oxygen to form highly damaging
hydroxyl radicals OH. Plant cells limit such damage by storing
surplus iron in an iron complex called phytoferritin.

How the iron is released from phytoferritin is not yet known but
this complex may be very important for functionalized food for
delivery iron and thereby may address dietary anemia problems.
Once in leaves, iron undergoes na important assimilatory reaction
(Figure 12.19).

To limit the free radical damage that free iron can cause, plant
cells may store surplus iron as phytoferritin.
Oxygen assimilation
90% of 0 is assimilated by respiration

Another major pathway for 0 assimilation into organic
compounds involves 0 incorporation from water

A small proportion of 0 can be directly assimilated into
organic compounds in the process of oxygen fixation via
enzymes called oxygenases.

The ribulose-1,5-bisphosphate carboxylase/oxygenase
(rubisco) is the most proeminent oxygenase. It
incorporate 0 into organic compound and releases
energy.
The Energetics of Nutrient Assimilation

Nutrient assimilation involves large amounts of energy to convert
stable low energy-inorganic compounds into high-energy organic
compounds.

For exemple, ten electrons are needed to reduce nitrate to nitrite and
then in amonium.

Many of these assimilatory reactions occur in the stroma of the
chloroplast, where they have ready access to powerful reducing
agents, such as NADPH, thioredoxin, and ferredoxin.

Energy requiring nutrient assimilation is coupled to photosynthetic
electron transport (Photoassimilation), which generates powerful
reducing agents (Figure 12.20). Photoassimilation and the Calvin-
Benson cycle occur in the same compartment
Energy requiring nutrient assimilation is coupled to photosynthetic electron
transport, which generates powerful reducing agents (Figure 12.20).

Photoassimilation operates only when photosynthetic electron transport generates
reductant in excess of the needs of the Calvin-Benson cycle (Figure 12.21).

Photoassimilation and the Calvin-Benson cycle occur in the same compartment, but only
when photosynthetic electron transport generates reductant in excess of the needs of the
Calvin-Benson cycle for exemple, under conditions of high light and low CO. High levels of
CO inhibit nitrate assimilation in the shoots of C plants.
END

You might also like