You are on page 1of 4

10.

11 RESPIRATION PROVIDES CARBON SKELETONS FOR


BIOSYNTHESIS
Before leaving the subject of cellular respiration, it is important to note that production of reducing potential and
ATP is not the sole purpose of the respiratory pathways. In addition to energy, the synthesis of nucleic acids,
protein, cellulose, and all other cellular molecules requires carbon skeletons as well. As noted at the beginning
of this chapter, respiration also serves to modify
the carbon skeletons of storage compounds to form these basic building blocks of cell structure. A few of the
more important building blocks that can be formed from intermediates in glycolysis and the citric acid cycle are
represented in Figure 10.14.

FIGURE 10.14 The role of respiration in biosynthesis. Intermediates in glycolysis and the citric acid cycle
are drawn off to serve as building blocks for the synthesis of cellular molecules. Carbon in the cycle is
maintained by the synthesis of oxaloacetate through anaplerotic reactions. This scheme is incomplete
and is intended only to give some indication of the importance of these two schemes in biosynthesis.

The withdrawal of glycolytic and citric acid cycle intermediates for the synthesis of other molecules means, of
course, that not all of the respiratory substrate will be fully oxidized to CO2 and water. The flow of carbon
through respiration no doubt represents a balance between the metabolic demands of the cell for ATP to drive
various energy-consuming functions on the one hand and demands for the reducing equivalents and carbon
skeletons required to build cell structure on the other.
It is also important to note that during periods of active synthesis, diversion of carbon from the citric acid cycle
for synthetic reactions will lead to a significant reduction in the level of oxaloacetate. These synthetic reactions
require not only carbon, but energy in the form of reducing potential and ATP as well. Without some means of
compensating for this loss of oxaloacetate, the cycle will slow down or, in the extreme case, come to a complete
halt and energy production will be impaired. This eventuality is precluded by the action of two cytosolic
enzymes: phosphoenolpyruvate (PEP) carboxylase (see Chapter 15) and malate dehydrogenase.
All plants, not just those with C4 photosynthetic activity (see Chapter 15), have some level of PEP carboxylase
activity that converts phosphoenolpyruvate (PEP) into oxaloacetate:
PEP + HCO-3 → oxaloacetate (10.9)
In this case the PEP is derived from glycolysis.
Although there is some evidence that oxaloacetate may be translocated directly into the mitochondrion, it is
more likely that oxaloacetate is quickly reduced to malate by the action of cytosolic malate dehydrogenase:
oxaloacetate + NADH →malate + NAD+ (10.10)
The malate would then pass into the mitochondrion, via amalate (or dicarboxylate) translocator, where it is
reoxidized to oxaloacetate by the action of a mitochondrial malate dehydrogenase:
malate + NAD+ → oxaloacetate + NADH (10.11)
The replenishment of oxaloacetate in this way is an example of a ‘‘filling-up’’ mechanism or anaplerotic
pathway. Thus carbon from glycolysis is delivered to the citric acid cycle through two separate but equally
important streams: (1) to citrate via pyruvate and acetyl-CoA and (2) from PEP via oxaloacetate and malate to
compensate for carbon ‘‘lost’’ to synthesis. Anaplerotic reactions such as the latter help to ensure that diversion
of carbon for synthesis does not adversely influence the overall carbon balance between energy-generating
catabolic reactions and biosynthetic anabolic reactions.
In addition to the normal enzymes of the citric acid cycle, plant mitochondria tend to have significant levels of
NAD+ -malic enzyme, which catalyzes the oxidative decarboxylation of malate:
malate+NAD+ → pyruvate+CO2+NADH (10.12)
The pyruvate may be further metabolized by pyruvate dehydrogenase to acetyl-CoA and from there enter the
citric acid cycle. Thus the mitochondrial pool of malate may replenish the citric acid intermediates through
either oxaloacetate or pyruvate.
In addition to serving an anaplerotic role, the uptake and oxidation of malate by mitochondria via either malic
enzyme or malate dehydrogenase also provides an alternative pathway for metabolizing malate. This alternative
pathway may be particularly significant in plants such as those of the family Crassulaceae (see Chapter 15) and
others that store significant levels of malate in their vacuoles. Finally, it should be noted that diversion of
pyruvate through oxaloacetate and malate bypasses the pyruvate kinase step in glycolysis (Figure 10.5, reaction
6) and thus reduces the yield of ATP by one. This reduction is, however, offset by gains achieved by reduction
of malate in the cytosol and its subsequent reoxidation in the mitochondrion. This sequence of reactions
effectively shuttles extramitochondrial NADH (generated during glycolysis) into the mitochondrion, where it
can be used to generate three molecules of ATP. This is a gain of one ATP over the two ATP generated via the
NADH-reductase route described earlier for extramitochondrial NADH.
10.12 RESPIRATORY RATE VARIES WITH DEVELOPMENT AND
METABOLIC STATE
The study of respiration at the level of individual organs or the whole plant becomes much more difficult than it
is for the study of individual cells. Whole plant respiration is normally studied by measuring the uptake of
oxygen or the evolution of CO2, but respiration rates obtained this way are highly variable. The balance of O 2
and CO2 exchange is dependent on the substrate being respired and the balance of fermentation, citric acid cycle,
and alternative pathway activities at any point in time. In addition, respiration rates differ between organs,
change with age and developmental state, and are markedly influenced by temperature, oxygen, salts, and other
environmental factors. Nevertheless, the study of respiration at the organ and plant level is a field of active
study. Understanding respiration at this level has important implications for the plant physiologist interested in
growth and development, for the physiological ecologist interested in plant biomass production, and the
agricultural scientist because of its impact on productivity and yield.
As a general rule, respiratory rate is a reflection of metabolic demands. Younger plants, organs, or tissues
respire more rapidly than older plants, organs, or tissues (Figure 10.15). The rapid rate of respiration during
early stages of growth is presumably related to synthetic requirements of rapidly dividing and enlarging cells.
As the plant or organ ages and approaches maturity, growth and its associated metabolic demands decline. Many
organs, especially leaves and some fruits, experience a transient rise in respiration, called a climacteric, that
marks the onset of senescence and the degenerative changes that precede death. Typically the climacteric rise in
O2 consumption is accompanied by a decline in oxidative phosphorylation, indicating that ATP production is no
longer tightly coupled to electron transport. The respiration rate of woody stems and branches, expressed on a
weight or mass basis, also declines as they grow. This is because as the diameter increases, the relative
proportion of nonrespiring woody tissue also increases.

FIGURE 10.15 Respiratory rate as a function of age. This type of curve applies generally to most
herbaceous plants, tissues, and organs. The magnitude of the climacteric will vary—some organs exhibit
little or no climacteric rise.
Carbon lost to the plant due to respiration can represent a significant proportion of the available carbon. Actual
respiration rates for plant tissues range in the extreme between barely detectable (0.005 µmol CO2 gW−1 d h−1) in
dormant seeds to 1000 µmol CO2 gW−1 d h−1 or more in the spadix of skunk cabbage during the respiratory crisis.
More typically, rates for vegetative tissues range from 10 to 200 (mol CO2 gW−1 d h−1 (Table 10.2). This may
represent a considerable fraction of the carbon assimilated by photosynthesis during a 24-hour period. (Recall
that photosynthesis occurs only during daylight hours, but respiration, especially of roots and similar tissues, is
ongoing 24 hours a day.) On the average, 30 to 60 percent of daily photoassimilate is lost as respiratory CO2. In
tropical rainforest species, probably because of accelerated enzymic activities at higher temperatures, this loss
may exceed 70 percent. Of the total daily
TABLE 10.2 Approximate specific dark respiration rates at 20 ◦C for crop species, deciduous
foliage, and conifers.
10.13 RESPIRATION RATES RESPOND TO ENVIRONMENTAL
CONDITIONS
10.13.1 LIGHT
The effects of light on mitochondrial respiration have been the subject of considerable debate for some time.
Traditionally, photosynthesis investigators and crop physiologists have tacitly assumed that respiration
continues in the light at a rate comparable to that in the dark. The true rate of photosynthesis is therefore taken
as equal to the apparent rate (measured as CO2 uptake) plus the rate of respiration (CO 2 evolved) in the dark.
However, attempts to study respiration in green leaves have led to alternative and conflicting conclusions. These
range from complete inhibition of mitochondrial activities, to partial operation of the citric acid cycle, or to
stimulation of respiration by light. The problem lies in the difficulty of measuring respiration during a period
when gas exchange is dominated by the overwhelming flux of CO2 and O2 due to photosynthesis, the recycling
of CO2 within the leaf, and the exchange of metabolites by chloroplasts and mitochondria.
Light effects on respiration during a subsequent dark period have been demonstrated. For example, dark
respiratory rates in leaves adapted to full sun (sun leaves) are generally higher than those of leaves of the same
species adapted to shade (shade leaves) (Table 10.2). As well, the rate is consistently higher in mature leaves of
shade-intolerant species than in shade-tolerant species. Indeed, a reduced respiratory rate appears to be a fairly
consistent response to low irradiance. This is probably related to lower growth rates also observed in shade-
grown plants, but it is not known which is cause and which is effect. It is projected that low respiratory and
growth rates may confer a survival advantage under conditions of deep shade. The basis for light regulation of
respiratory rate is unknown, although some have suggested that low respiratory rates under conditions of low
irradiance may reflect availability of substrate. For example, the respiration rate 1 to 2 hours following a period
of active photosynthesis is higher than after a long dark period.

Other experiments, however, have shown that dark respiratory rates do not correlate with CO2 supply during the
previous light period. This seems to suggest a more direct effect on respiration. The pyruvate dehydrogenase
complex (PDH) of the mitochondrion can exist either in an active, nonphosphorylated form or an inactive,
phosphorylated form (Figure 10.16). It has been shown that the photorespiratory-generated NH+4 (Chapter 8)
stimulates the phosphorylation of the mitochondrial pyruvate dehydrogenase complex, thereby inhibiting the
rate of respiration in the light by decreasing the rate at which acetyl-CoA is generated for the CAC. Light
regulation of respiration remains a controversial issue in plant physiology.

FIGURE 10.16 Regulation of mitochondrial pyruvate dehydrogenase complex. The pyruvate dehydrogenase
complex (PDH) exists in two forms: an active, nonphosphorylated form and an inactive, phosphorylated
form. The reversible interconversion of the active and inactive forms of this enzyme is the result of the
activity of the PDH kinase, which consumes ATP to phosphorylate PDH, and the activity of the phospho-
PDH phosphatase, which dephosphorylates PDH. Ammonium ion (NH 4 ) stimulates the PDH kinase and
+

hence inactivates the PDH complex.


10.13.2 TEMPERATURE
One of the most commonly applied quantitative measures used to describe the effect of temperature on aprocess
is the temperature coefficient, or Q10, given by the expression:
Q10 = rate at(t + 10)◦C (10.13)
rate at t◦C

At temperatures between 5◦C and about 25◦C or 30◦C, respiration rises exponentially with temperature and the
Q10 value is approximately 2.0 in many but not all plants (see Chapter 14). Within this temperature range, a
doubling of rate for every 10◦C rise in temperature is typical of enzymic reactions. At temperatures above 30 ◦C,
the Q10 in most plants begins to fall off as substrate availability becomes limiting. In particular, the solubility of
O2 declines as temperatures increase and the diffusion rate (with a Q10 close to 1) does not increase sufficiently
to compensate. As temperatures approach 50◦C to 60◦C, thermal denaturation of respiratory enzymes and
damage to membranes bring respiration to a halt.
Some investigators have observed differences in the rate of respiration in tropical, temperate, and arctic species
at different temperatures. For example, the respiration rate of leaves of tropical plants at 30 ◦C is about the same
as that of arctic species at 10◦C. The temperature coefficient (Q10) for respiration is the same in both cases and
there is no evidence of intrinsic differences in the biochemistry of respiration. It is likely that the differences
reflect differences in temperature optima for growth of arctic and tropical species—optima determined by
factors other than respiration—and the consequent metabolic demand for ATP. A detailed discussion of the
effects of temperature stress and acclimation on respiration will be focused in Chapters 13 and 14.
10.13.3 OXYGEN AVAILABILITY
As the terminal electron acceptor, oxygen availability is obviously an important factor in determining respiration
rate. The oxygen content of the atmosphere is relatively stable at about 21 percent O2. The equilibrium
concentration of oxygen in air-saturated water, including the cytosol, is approximately 250 M. However,
cytochrome c oxidase has a very high affinity for oxygen with a Km (see Chapter 8, Box 8.1) less than 1 M.
Under normal circumstances, oxygen is rarely a limiting factor.
There are some situations, however, where oxygen availability may become a significant factor. One is in bulky
tissues with low surface-to-volume ratios, such as potato tubers and similar storage tissues, where the diffusion
of oxygen may be slow enough to restrict respiration. This may not be a serious problem, however. A significant
volume—as much as 40 percent—of roots and similar tissues may be occupied by intercellular air spaces that
aid in the rapid distribution of O2 absorbed from the soil or, in some cases, from the aerial portions of the plant.
Plants are most likely to experience oxygen deficits during periods of flooding, when air in the large pore spaces
of the soil is displaced by water, thereby decreasing the oxygen supply to the roots. For similar reasons, plants
grown in hydroponic culture must be aerated to maintain adequate oxygen levels in the vicinity of the roots
(Chapter 3).

You might also like