You are on page 1of 62

EFFECTS OF CO2 PLANT RESPIRATION

NAME :- SNEHA BAG

UNIVERSITY ROLL NO. :- S03919BOT004

DEPARTMENT :- BOTANY DEPARTMENT (6TH SEM)

COLLEGE :- MUNICIPAL COLLEGE ROURKELA ,


ROURKELA
ACKNOWLEDGEMENT

I would like to express my sincere gratitude to Dr. Sanatan Pradhan, Head


Of Botany Department to guide me and providing facilities to carryout my
project work.

I am highly indebited to my Amy Jocelyn Kujur, Lecture Botany department


for her valuable guidance which has promoted my efforts in all the stages of
this project work. My thanks and appreciation go to my classmates in
developing my project and to the people who have willingly helped me out
with their abilities

Finally, words are not sufficient to express gratitude my cherished family


member for supporting me without their encouragement and support. I would
have not reached this stage.

2|Page
CERTIFICATE

This is to certify that “Sneha Bag” student of +3 Bot(H) 6th sem bearing Roll
no:S03919BOT004 has successfully completed her Botany project on
“Effects Of CO2 On Plant Respiration” under my supervision. The
Original research work was carried out by her under my supervision in the
academic year 2021-2022. On the basis of the declaration made by her , I
recommended this project report for evaluation.

Signature of HOD, Botany department


Municipal College Rourkela

3|Page
CONTENTS

Sl.no Topics Page no


1 Abstract 5
2 Introduction 6
3 Significance of Respiration 6-7
4 Respiratory Quotient 7-12
5 Types of Respiration 13
6 Mechanism of Respiration 13-17
7 Review of Literature 18-19
8 Direct and Indirect effects of [CO2] on 20
plant respiration rate
9 Direct effect 21
10 Direct effect on enzyme 21-24
11 Direct effect on intact tissue 24-26
12 Are there direct effects of CO2 on 27-28
whole plant
13 Indirect effect 28-29
14 Indirect effect on enzymes 29-32
15 Indirect effect of CO2 on respiration of 33-39
tissue and whole plant
16 Integrated effect of elevated [CO2] on 39-45
plant respiration at the ecosystem level
17 Conclusion 46
18 Reference 47-62

4|Page
ABSTRACT

Elevated levels of atmospheric [CO2] are likely to enhance photosynthesis


and plant growth, which in turn should result in increased specific and whole-
plant respiration rates. However, a large body of literature has shown that
specific respiration rates of plant tissues are often reduced when plants are
exposed to, or grow at, high [CO2] due to direct effects on enzymes and
indirect effects derived from changes in the plant’s chemical composition.

5|Page
INTRODUCTION

Respiration is essential for growth and maintenance of all plant tissues, and
plays an important role in the carbon balance of individual cells, whole-
plants and ecosystems, as well as in the global carbon cycle. Through the
processes of respiration, solar energy conserved during photosynthesis and
stored as chemical energy in organic molecules is released in a regulated
manner for the production of ATP(adenosine triphosphate), the universal
currency of biological energy transformations, and reducing power (e.g.
NADH and NADPH).

NADPH and NADH are coenzymes, which take part in various metabolic
processes. NADPH contains an extra phosphate group. NADH is involved in
cellular respiration, whereas NADPH is involved in photosynthesis. NADPH
and NADH are the reduced forms of NADP+ and NAD+, respectively. These
are two universal coenzymes functioning as hydride(H-) carries at a standard
reduction potential of -320V.

Significance of Respiration
The respiration is an important process because

(i)It releases energy which is consumed in various metabolic processes

essential for plant lite and activates cell division.

(ii) It brings about the formation of other necessary compounds participating

as important cell constituents.

6|Page
(iii) It converts insoluble food into soluble form.

(iv) It liberates carbon dioxide and plays a part actively in maintaining the

balance of carbon cycle in nature.

(v) It converts stored energy (potential energy) into usable form (kinetic

energy).

RESPIRATORY QUOTIENT

Respiratoy Quotient (RQ) may be defined as "the ratio between the volume
of carbon dioxide given out and oxygen taken in simultaneously by a given
weight of the tissue in a given period of time at standard temperature and
pressure.
Volume of CO2 evolved
____________________________
Respiratory Quotient =
Volume of O2 absorbed
The value of RQ depends upon the nature of respiratory substrate or on the
amount of oxygen present in the respiratory substrate and the extent to which
substrate is broken down. Theoretically, the value of RQ should be unity but
deviations in the value from unity are very common because of the deviations
in the oxidation and reduction levels of the respiratory substrates. The value
of RQ also depends upon whether all oxygen which is being absorbed is
being used up in respiration or some of it is being utilised for some other
purposes. Because of these reasons the value of RQ may be unity, more than
unity, less than unity and even zero. The value of RQ provides some idea
about the type of substrate being oxidised and of the degree of aerobic
respiration preceding. The RQ value of different substrates is discussed here
as follows.

7|Page
(1) Respiratory Quotient of Carbohydrates
Carbohydrates are principle respiratory substrates in plants and the
commonest among them are starch, inulin, sucrose, glucose and fructose. Of
these, complex carbohydrates are first hydrolysed to simple or hexose sugars
because as such they can't be utilised as respiratory substrate.

Starch -› polysaccharides -› disaccharides -› monosaccharides or hexoses.

When these carbohydrates such as hexoses are used as respiratory substrate,


the volume of CO, evolved is equal to the volume of O, absorbed as is evident
from the equation given below. The RQ is found to be unity

C6H12O6 + 6O2 6CO2 + 6H2O


(6 VOLUMES) (6 VOLUMES)
RQ=CO2/ O2 = 6CO2/6O2 = 1 or Unit

In low oxygen concentration anaerobic respiration takes place in addition to


aerobic respiration. As a result, the CO, production increases which results
in raising the value of RQ.

During respiration, a part of carbon dioxide may be utilised


nonphotosynthetically by living cells in lengthening the carbon chains of
their organic compounds. Therefore, RQ may appear less than unity in such
cases of carbohydrates. Similarly, RQ may deviate from unity when (1)
respiratory substrate is other than carbohydrate, (2) carbohydrates or other
respiratory substrates are partially oxidised, (3) oxygen absorbed is used up
in a process other than respiration or (4) CO, formed is utilised metabolically
instead of being given out.

8|Page
(2) Respiratory Quotient of Fats
Fats are not commonly found in vegetative parts of the plant but these are
important as storage food in seeds. Nearly 80 per cent angiospermic seeds,
prefer to store fat as main reserve food. At the time of seed germination,
major portion of fat is converted into carbohydrates while the rest is utilised
for respiration.

Fats are poorer in oxygen and the proportion of oxygen to Carbon in fats is
invariably less as compared to carbohydrate, hence they require more oxygen
for complete oxidation. Apart from this, fats are not oxidised directly. They
are first hydrolysed to fatty acid and glycerol. A fraction of oxygen is used
up in this process. Due to these reasons equations given below indicate that
the value of RQ of fats found to be about 0.7, i.e. less than unity.

2 C51M96O6 + 145O2 102CO2 + 98H2O


(Tripalmitia)

CO
____2 2CO
____
RQ = =
O2
___ 2O2
___

However, simple fatty acids such as acetic acid liberate equal amount of CO,
and in such cases RQ is unity.

CH3COOH+2O2 2CO2+2 H2O


(Acetic Acid)

CO
____2 2CO
____ 2
RQ = = = 1 or unity
O2
___ 2O
___ 2

It is quite interesting to note that falShberate more energy than


carbohydrates. A gram of carbhoydrate yields 3.8 kcal, while a similar
amount of fat produces about 9.1 kcal. May this be the reason why fats are
common reserve food materials in seeds.

9|Page
(3) Respiratory Quotient of Proteins and Derivatives
Proteins and amino acid derivatives serve normally as respiratory substrate
only in seeds rich in proteins. The vegetative cells comsume protein only
during Starving conditions. Otherwise, proteins are seldom respired.

Like fats, proteins are also compounds with lesser oxygen as compared to
carbohydrate and the proportion of oxygen to Carbon is invariably low.
Proteins rarely serve as respiratory substrates but when they do serve their
hydrolysis products require more oxygen for complete oxidation as a result
of which, the value of RQ falls to less than unity. The RQ value of proteins
fluctuates around 0.79. During protein oxidation, the value of RQ may
be=1.0(0.99) when ammonia is produced or 0.8 (0.79) when amide formation
occurs. When amides are oxidised the RO value rises above one.

(4) Respiratory Quotient of Succulents


In succulent plants such as Opuntia or members of family Crassulaceae and
in anthocyanin rich leaves, the complete oxidation of carbohydrate does not
occur, as a result of which only intermediate products are formed without the
production of CO2 . At night, when stomata are open in, succulent plans,
oxygen is absorbed and intermediate compounds are formed due to partial
oxidation on account of which there is no evolution of CO2 and the RQ value
is found to be less than one, mostly zero. In day time, when complete
oxidation of intermediate acids occurs the CO2 produced is used up in
photosynthesis with the result that there is no evolution of CO2 and RQ falls
to zero.

10 | P a g e
(5) Respiratory Quotient of Organic Acids
Organic acids are rich in oxygen and the proportion of oxygen to carbon is
very high. When organic acids are used as respiratory substrate lesser oxygen
is needed to be absorbed and more CO, is evolved with the result that RQ
value is found to be more than unity.

Malic acid may also convert into hexose, with following intermediate stages.

These reactions are popularly known as Wood-Werkman reaction.

2COOH. CH2CHOH.COOH 2COOH.CH2CO.COOH+4(H)

2COOH.CH2CO.COOH 2CH3CO.COOH+2CO2

2CH3.CO.COOH+4(H) C6H12O6

(hexose)

If this hexose is partly respired, the RQ may increase to 1.33 or more. Some
times RQ value is found to be 0.2 to 0.3 when the reaction gets associated
with a combination of hexose respiration and organic acid synthesis at the
expense of CO, i.e.

C6H12O6 + 2CO2 2C4H6O5


(malic acid)

11 | P a g e
(6) Respiratory Quotient when Oxygen is Utilised for Other
Metabolic Processes

Apart from respiration some other metabolic processes such as synthesis of


anthocyanins and conversion of fats to carbohydrates also require oxygen. In
such cases amount of CO, evolved does not correspond to the amount of
oxygen absorbed and, therefore, the value of RQ falls below unity. In
Bryophyllum, leaves are capable of utilising CO, (liberated during
respiration) for the synthetizing organic acids in the dark so the RQ value
falls below unity.

(7) Respiratory Quotient of Maturing Fatty Seeds


During the maturation of fatty seeds simple carbohydrates are converted into
fats. In the process oxygen is released but it is used up in respiration. As a
result CO, is released during respiration but oxygen is not absorbed from
outside. In such cases the value of RQ is found to be more than unity. But in
germinating fatty seeds, RQ value falls below unity because of combined
effect of the seed using fat substrate for respiration and also synthetizing
carbohydrates from fats.

(8) Respiratory Quotient of Tissue Respiring in Absence of


Oxygen
In absence of oxygen (anaerobic respiration) in which CO, is evolved without
O2 being absorbed, the RQ value is found to be more than unity.

C6 H12 O6 2C2H5OH+2CO2
RQ=CO2 /O2=2/0=infinity.
12 | P a g e
TYPES OF RESPIRATION
On the basis of availability of oxygen, respiration has been divided into two
categories.

(A) Aerobic respiration

It takes place in presence of oxygen and the stored food (respiratory


substrate) gets completely oxidised into carbon dioxide and water, i.e.

C6H12O6 + 6O2 6CO2 + 6H2O + 686 kcal.

This type of respiration is of common occurrence and is found in all plants.

(B) Anaerobic respiration

It takes place in absence of oxygen or when oxygen concentration is less than


one per cent. The stored food is incompletely Oxidised and instead of carbon
dioxide and water certain other compounds are also formed. This type of
respiration is of rare occurrence but common among micro-organisms like
yeasts and can be represented by

C6H12O6 2C2H5OH + 2CO2 + 56 kcal

MECHANISM OF RESPIRATION
Respiratory substrate. The substrates which are broken down in respiration
for the release of energy may be carbohydrates, fats or proteins. Proteins are
used up as respiratory substrate only when carbohydrates and fats are not
available Blackman proposed that respiration in which carbohydrates are

13 | P a g e
used as respiratory substrate are called floating respiration and if proteins are
used, protoplasmic respiration.

Fats are used as respiratory substrates after their hydrolysis to fatty acids and
glycerol by lipase and their subsequent conversion to hexose sugars. Proteins
serve as respiratory substrates after their breakdown to amino acids by
proteolytic enzymes. As regards carbohydrates, not only simple hexose
sugars like glucose and fructose but complex disaccharides particularly
sucrose and polysaccharides such as starch, inulin and hemicelluloses are
also used as respiratory substrates.

During respiration, the complex substrates are broken down into simpler ones
and finally CO, is liberated and water is formed. During oxidation of
respiratory substrate, certain amount of energy is released. Part of this energy
is trapped in the form of energy rich compounds such as ATP while
remaining part is lost in the form of heat. The energy trapped in ATP
molecules can be used in various ways (both for physical and chemical
requirements). Here, the typical example of respiratory substrate, i.e.
carbohydrate has been discussed. Oxidation of fats and proteins are given
separately.

All complex carbohydrates are firstly converted into hexoses (glucose of


fructose) before actually entering into respiratory process. The oxidation of
glucose to CO2 and water consists of two distinguishable phases:

(i) Glycolysis and (ii) Krebs cycle.

In glycolysis glucose is converted into pyruvic acid. Steps of glycolytic


pathway are common to all kinds of respiration, hence may be called
common respiratory metabolism.

A quantitatively important by-product of respiration is CO2 and, therefore,


plant and ecosystem respiration play a major role in the global carbon cycle.

14 | P a g e
Terrestrial ecosystems exchange about 120 Gt of carbon per year with the
atmosphere, through the processes of photosynthesis and respiration
(Schlesinger ,1997). Roughly, half of the CO2 assimilated annually through
photosynthesis is released back to the atmosphere by plant respiration
(Gifford 1994; Amthor, 1995). Because terrestrial biosphere-atmosphere
fluxes of CO2 far outweigh anthropogenic inputs of CO2 to the atmosphere,
a small change in terrestrial respiration could have a significant impact on
the annual increment in atmospheric [CO2] (Amthor, 1997). A large body of
literature has indicated that plant respiration is reduced in plants grown at
high [CO2]. For example, it was estimated that the observed 15–20 %
reduction in plant tissue respiration caused by doubling current atmospheric
[CO2] (Amthor, 1997; Drake et al., 1997; Curtis and Wang, 1998), could
increase the net sink capacity of global ecosystems by 3·4 Gt of carbon per
year (Drake et al., 1999), and thus offset an equivalent amount of carbon
from anthropogenic CO2 emissions. Therefore, not only are gross changes in
respiration important for large-scale carbon balance issues, changes in
specific rates of respiration can have significant impact on basic plant
biology such as growth, biomass partitioning or nutrient uptake (Amthor,
1991; Wullschleger et al., 1994; Drake et al., 1997).

Several reviews of the literature have concluded that elevated concentrations


of carbon dioxide, [CO2], generally reduce rates of respiration, primarily by
a direct effect (c.f. Drake et al., 1999). However, some recent studies have
found little or no direct response of respiration to [CO2] and have suggested
that direct effects of [CO2] on respiration are measurement artefacts
attributable primarily to leaks (Amthor, 2000a; Jahnke, 2001; Tjoelker et al.,
2001a; Bruhn et al., 2002; Jahnke and Krewitt, 2002). While some studies
that have measured leakage and taken it into account have still found effects
of [CO2] on respiration (e.g. Hilbert et al., 1987; Baker et al.,
2000; Hamilton et al., 2001), it is difficult to be sure that CO2 exchange rates
are completely without systematic error. Another approach to resolve the
controversy would be to determine whether [CO2] affects processes directly
dependent on respiration.

15 | P a g e
Photosynthetically active non‐growing leaves generally accumulate
carbohydrates during the day, which are mainly translocated to other parts of
the plant during the night. Translocation, at least in darkness, requires energy,
supplied by respiration, for the active transport of carbohydrates and other
materials into the phloem and sometimes also for inter‐conversions among
forms of carbohydrates (Amthor 2000b). While carbohydrates are the major
type of material translocated in the phloem, a decision was made here to
estimate translocation from changes in total dry mass rather than from
changes in non‐structural carbohydrates, because non‐structural
carbohydrates as usually measured may be a variable fraction of the total dry
mass accumulated and translocated diurnally (e.g. Bunce, 1982).

Earlier work (Bunce, 2002) indicated that translocation from soybean leaves
was affected by [CO2] at night. However, in that work, translocation was
determined from changes in dry mass in relation to photosynthesis and
respiration rate over three day/night cycles. In the present work, the fact that
plants exposed to long days have rapid rates of translocation at night
(e.g. Chatterton and Silvius, 1979) was utilized, which allowed translocation
to be measured with sufficient accuracy over a single 8‐h dark period. The
estimate of translocation was therefore independent of photosynthetic
measurements, and depended only to a minor extent on estimates of
respiration.

The chemical reduction of nitrate to nitrite requires metabolic energy. During


the light period, some of this energy may be supplied photochemically, but
during the dark period it must come from respiration (Amthor, 2000b). In the
present work, the rate of reduction of nitrate in the dark was also examined
for responsiveness to [CO2]. The rate of nitrate reduction was measured by
the rate of disappearance of nitrate. This was necessarily determined with
excised leaves, because in intact leaves the amount of nitrate entering a leaf
via the transpiration stream would be unknown and might vary with [CO 2]
treatment.

16 | P a g e
While responses of respiration to [CO2] are controversial, responses of
respiration to temperature are not disputed. Comparisons were made of the
responses of respiration, translocation and nitrate reduction to [CO2] and to
temperature, to test whether a given change in respiration produced the same
change in translocation or nitrate reduction regardless of whether respiration
was changed by altered [CO2] or temperature. Similar responses would
strongly suggest that the apparent response of respiration to [CO 2] was real.

Scaling the effects of an increase in atmospheric [CO2] on plant respiration


at the biochemical level to the whole ecosystem is difficult for at least two
important reasons: (1) the fine and coarse control points of respiratory
pathways in tissues and whole plants are not well known; and (2) it is unclear
how respiration rates actively or passively adjust to the effects that elevated
[CO2] may have both upstream (e.g. on substrate availability) and
downstream (e.g. on energy demand) of the carbon oxidation pathways. In
addition, accurate measurements of respiratory rates as CO 2 evolution have
been proven difficult with current techniques (Gonzalez-Meler and Siedow,
1999; Hamilton et al., 2001; Jahnke, 2001), presenting yet another limitation
to scaling up process-based measurements of respiration rates to organism,
ecosystem or global levels. In this document, we re-evaluate the theory on
respiration responses to elevated [CO2] in view of recent studies that have
provided new insights on the effects of a short- and long-term change in
atmospheric [CO2] on plant respiration.

17 | P a g e
Review of literature

Much is still unknown about how plants will react to e[CO2] and with nutrient
deficiencies observed in agricultural crops, this will become increasingly
more important to understand. The production of carbohydrates is increased
in plants grown under e[CO2] due to an increase in photosynthesis. Some
carbohydrates are produced in higher quantities than others depending on the
plant, though production of sucrose is reportedly higher compared to
hexoses. The studies discussed provide an insight into how these sugars can
be used to regulate many functions in roots. Most of the information on sugar
signaling discusses the glucose and sucrose pathways. The amount of carbon
partitioned into either of those carbohydrates may be in part determined by
which carbohydrate the plant requires to regulate specific genes, though this
is unknown. Nutrient acquisition appears to be regulated by sugars, as
evidenced by the regulation of expression of various ion transporters as well
as the ability for sugars to affect root growth. Finally, both e[CO2] and sugars
are able to affect the biosynthesis of certain plant hormones, which may
suggest that sugars function as an intermediate in e[CO 2] control of
hormones. From these studies we can begin to think about what changes
might occur in roots of plants grown in future carbon dioxide concentrations.

Respiration of attached leaves increased by a factor of 2·2 for the 10 °C range


in temperature, similar to innumerable other studies (reviewed in Tjolker et
al., 2001b). The increase in respiration with short‐term increases in
temperature has been envisioned as a direct effect of temperature on
biochemical reaction rates in the respiratory pathway, and not the result of
an increase in the demand for respiration by processes utilizing the energy
supplied by respiration (Amthor, 2000b). While it seems possible that
temperature could affect translocation and nitrate reduction by mechanisms
unrelated to respiratory energy supply, it seems less likely that [CO 2] directly
affects both translocation and nitrate reduction. The similarity of the changes
in translocation relative to respiration and of the changes in nitrate reduction
relative to respiration for both the temperature and [CO2] treatments

18 | P a g e
therefore suggests that both the changes in translocation and nitrate
reduction were mediated by the changes in respiration. This suggests that
translocation and nitrate reduction were limited by the energy supplied by
respiration in this case. However, regardless of the mechanisms involved, the
parallel responses of translocation and nitrate reduction for both the
temperature and [CO2] treatments make it unlikely that the response of
respiration to one variable was an artefact while the response to the other
was real, particularly since respiration was measured by the same method
and in the same apparatus in both cases. Recently, Pinelli and Loreto (2003)
measured dark respiration using measurements of 12CO2 emission and also
found that it decreased as [CO2] increased.

While there is considerable information about the amount of respiratory


energy used in the many processes dependent on it during normal growth,
little is known about how the energy is partitioned among the processes
competing for it when the respiration is limited (Amthor, 2000b). In these
results, translocation was relatively more sensitive to the inhibition of
respiration by the low temperature and elevated [CO2] treatments than was
nitrate reduction, which suggests nitrate reduction had higher priority for
respiratory energy supply. While complete cessation of translocation by the
20 °C treatment seems a drastic response, subsequent experiments have
shown that respiration and translocation both recovered to their initial rates
after a few days of acclimation to 20 °C.

19 | P a g e
DIRECT AND INDIRECT EFFECTS OF [CO2] ON PLANT
RESPIRATION RATE

Amthor,[ Plant, Cell and Environment 14:13–20.](1991) described two


different effects of [CO2] on plant respiration rates that can be distinguished
experimentally: a direct effect, in which the rate of respiration of
mitochondria or tissues could be rapidly and reversibly reduced following a
rapid increase in [CO2] (Amthor ,[ Plant Physiology 98:1–4.] 1992; Amthor,[
Tree Physiology 20:139–144] 2000a), and an indirect effect in which high
[CO2] changes the rate of respiration in plant tissues compared with the rate
seen for those plants grown in normal ambient [CO2] when both are tested at
a common background [CO2] (Azcón-Bieto [ Plant Physiology-106], 1994).
Most of the studies investigating the effects of a change in [CO2] on
respiration rates have described magnitude (from non-significant to 60 %)
and direction (either stimulation, inhibition or no effect) of these two effects,
with little progress on their underlying mechanisms. Confounding
measurement artefacts are an added complication for establishing the
presence and magnitude of direct and indirect effects of [CO 2] on respiration
(Gonzalez-Meler and Siedow,[ Tree Physiology 19:253–259]
1999; Amthor,[ Tree Physiology20:139–144] 2000a; Jahnke,[ Plant, Cell
and Environment24:1139–1151] 2001; Davey,[Plant Physiology134:520–
527.] 2004).

20 | P a g e
DIRECT EFFECTS

Previously, a rapid short-term doubling of current atmospheric CO 2 levels


was reported to inhibit respiration of mitochondria and plant tissues by 15–
20 % (Amthor,[ American Society of Agronomy Special Publication]
1997; Drake [Annual Review of Plant Physiology and Plant Molecular
Biology 48:609–639.], 1997; Curtis and Wang,[Oecologia 113:299–313]
1998). Although direct effects of [CO2] on respiratory enzymes have been
reported, the magnitude of the direct inhibition of intact tissue respiration by
elevated [CO2] has now been shown to be explained by measurement
artefacts (i.e. Gonzalez-Meler and Siedow,[Tree Physiology 19:253–259.]
1999; Jahnke,[Plant, Cell and Environment 24:1139–1151.] 2001),
diminishing the impact that such reductions in respiration rate may have on
plant growth and the global carbon cycle.

DIRECT EFFECTS ON ENZYMES

Doubling current levels in [CO2] (here we review the effects of only two- or
three-fold increases in atmospheric CO2) have been shown to inhibit the
oxygen uptake of isolated mitochondria and the activity of mitochondrial
enzymes under some conditions (Reuveni.,[Annals of Botany 76:291–
295]1995; Gonzàlez-Meler,[ Global change: effects on coniferous forests
and grasslands. SCOPE, Vol. 56] 1996a, b). Twice current atmospheric
[CO2] reduces the activity of cytochrome c oxidase and succinate
dehydrogenase in isolated mitochondria from cotyledon and roots of Glycine

21 | P a g e
max (Gonzàlez-Meler,[ Plant Physiology 112:1349–1355.] 1996b).
Mitochondrial oxygen uptake may show either no response or up to 15 %
inhibition by a rapid increase in [CO2], depending on the electron donor used
in the assay (Gonzàlez-Meler,[Plant Physiology 112:1349–1355.] 1996b; see
also Fig. 1), indicating that the response of mitochondrial enzymes to high
[CO2] depends on the cell's metabolism (Gonzalez-Meler and Siedow,[ Tree
Physiology 19:253–259.] 1999; Affourtit[Biochimica Biophysica Acta –
Bioenergetics 1504:58–69.], 2001).

FIG. 1.

Effect of doubling the concentration of CO2 on the oxygen uptake of


mitochondria isolated from 5-d-old soybean cotyledons in state 4 conditions
(absence of ADP). Oxidation of NADH and pyruvate was measured in intact
22 | P a g e
mitochondria that were incubated for 10 min with 0 (open bars) or 0·1 (solid
bars) mM dissolved inorganic carbon (DIC) at 25 °C. Experiments were
done using the oxygen isotope technique as described in Ribas-Carbo.[ Plant
Physiology 109:829–837.] (1995) to detect the effects of CO2 on the total
activity of mitochondrial respiration (Vt), the activity of the cytochrome
pathway (vcyt) and the activity of the alternative pathway (valt). Results are
average of four replicates and standard errors.
In scaling the direct effects of high [CO2] on respiratory enzymes to intact
tissues two factors need to be considered: (1) cytochrome oxidase exerts up
to 50 % of the total respiratory control (for definitions, see Kacser and
Burns,[ Biochemical Society Transactions 7:1149–1160.] 1979) at the level
of the mitochondrial electron transport chain (reviewed by Gonzalez-Meler
and Siedow,[ Tree Physiology 19:253–259.] 1999); and (2) the competitive
nature of the cytochrome and alternative pathways of mitochondrial
respiration (Affourtit,[ Biochimica Biophysica Acta – Bioenergetics
1504:58–69.] 2001). The activity of the alternative pathway of respiration
could increase upon doubling of the [CO2], masking a direct CO2 inhibition
of the cytochrome pathway (Fig. 1). The oxygen isotope technique (Ribas-
Carbo,[ Plant Physiology 109:829–837.] 1995) allows for the distinction of
direct effects of [CO2] on oxygen uptake activity by the cytochrome and the
alternative oxidases. Figure 1 shows that mitochondrial electron transport
activity is affected when CO2 (a mild inhibitor) restricts the normal electron
flow through one of the pathways (Cyt pathway – vcyt). Under conditions of
low alternative pathway activity (for details, see Ribas-Carbo,[ Plant
Physiology 109:829–837.] 1995), inhibition of the cytochrome pathway (18
%) by doubling the ambient [CO2] is compensated for by a similar increase
in alternative pathway activity (valt), resulting in no significant reduction in
overall oxygen uptake of the isolated mitochondria. These results show that
enhanced activity of the alternative pathway upon an increase in [CO 2]
provides a mechanism by which the direct [CO2] inhibition of the
cytochrome oxidase may not be seen in the total oxygen uptake under
conditions of high adenylate control (low ADP). Gonzàlez-Meler and

23 | P a g e
Siedow[Tree Physiology 19:253–259] (1999) argued that direct effects of
CO2 on respiration exceeding >10 % inhibition are likely due to factors other
than inhibition of mitochondrial enzymes. Due to confounding measurement
artefacts (Jahnke,[ Plant, Cell and Environment 24:1139–1151.] 2001), rates
of tissue respiration have been reported to decrease >10 % when the
atmospheric [CO2] doubles.

DIRECT EFFECTS ON INTACT TISSUES

Recent evidence indicates that, in studying direct effects of CO2 on


respiratory CO2 efflux of intact tissues, investigators should first be
concerned about measurement artefacts. Amthor[. Advances in carbon
dioxide effects research. American Society of Agronomy Special
Publication](1997) and Gonzalez-Meler and Siedow[Tree Physiology
19:253–259.] (1999) showed that gas CO2 exchange measurement errors
could augment and explain the magnitude of the direct effect. New studies
made in this context have shown that respiration rates are little or not at all
inhibited by a doubling of atmospheric [CO2] (Amthor, 2000a[Tree
Physiology 20:139–144.]; Amthor,[ Journal Experimental Botany 52:2235–
2238.] 2001; Bunce, 2001[Journal Experimental Botany 52:2235–
2238.]; Hamilton,[ Plant, Cell and Environment 24:975–
982.]2001; Tjoelker,[New Phytologist 150:419–424.] 2001; Bruhn,[
Physiologia Plantarum 114:57–64.] 2002; Davey,[ Plant Physiology
134:520–527.] 2004). Indeed, Jahnke[Plant, Cell and Environment 24:1139–
1151.](2001) and Jahnke and Krewitt[Plant, Cell and Environment 25:641–
651.] (2002) confirmed that measurement artefacts due to leakage in CO 2-
exchange systems could be as large as the previously reported direct
inhibitory effects. They also found that leaks through the aerial spaces of
homobaric leaves showed a significant apparent inhibition of CO2 efflux that
was not due to an inhibition of respiration by elevated [CO2]. Therefore,
considerations of the impact direct effects of [CO2] on plant respiration rates
may have on the global carbon cycle were overstated (Gonzalez-
24 | P a g e
Meler,[ SCOPE, Vol. 56. UK: John Wiley & Sons, 161–181.]
1996a; Drake,[ Plant, Cell and Environment 22:649–657.] 1999).
Corrections for instrument leaks can be applied in most cases (Bunce,
2001; Jahnke,[Annals of Botany 87:463–468.] 2001; Pons and Welschen,[
Plant, Cell and Environment 25:1367.] 2002). Applying some corrections for
gas exchange leaks, Bunce[Annals of Botany 87:463–468.] (2001) reported
a significant reduction in respiration after a rapid increase in [CO 2].
Moreover, simultaneously switching the air isotopic composition
from 12CO2 to pure 13CO2 when the [CO2] was changed, showed that the
efflux of 12CO2 (which represents net respiratory CO2 evolution originating
from respiratory substrates minus the CO2 being re-fixed by carboxylases,
see below) was reduced in leaves of C3 plants, but not C4, when the [CO2]
increased (Pinelli and Loreto,[ Journal of Experimental Botany 54:1761–
1769.] 2003). These recent results illustrate that not all reports of direct
inhibition of respiration by elevated [CO2] have yet been reconciled with
each other, but they do not show a clear proof that the direct effect, as defined
by Amthor[Plant, Cell and Environment 14:13–20.] (1991) exits either.
An alternate explanation for the apparent reduction in respiration rates upon
an increase in [CO2] (Bunce,[Annals of Botany 87:463–468.] 2001; Pinelli
and Loreto,[Journal of Experimental Botany 54:1761–1769.] 2003) is an
increase in dark CO2 fixation catalysed by phosphoenolpyruvate carboxylase
(PEPC), resulting in an apparent reduction of CO2 efflux (Amthor,[
American Society of Agronomy Special Publication] 1997; Drake,[Plant,
Cell and Environment 22:649–657.] 1999). However,
in Rumex and Glycine leaves, Amthor.[ Journal Experimental Botany
52:2235–2238.] (2001) found an effect of elevated [CO2] on CO2 efflux in
only a minority of experiments, with no effect of elevated [CO2] on the
O2 uptake of the same leaves, indicating at most a small effect of elevated
[CO2] on PEPC activity or respiration rates. Similar results were found for a
variety of species, involving 600 measurements, in which [CO 2] increases
did not alter the CO2 and O2 leaf exchanges in the dark (Davey,[ Plant
Physiology 134:520–527.] 2004).
It seems that direct effects of [CO2] on mitochondrial enzymes may have no
consequence on the specific respiratory rate of whole tissues. Because

25 | P a g e
mitochondrial respiratory enzymes are generally ‘in excess’ to the levels
required to support normal tissue respiratory activity under most growth
conditions (Gonzalez-Meler and Siedow,[ Tree Physiology 19:253–259.]
1999; Atkin and Tjoelker,[Trends in Plant Science 8:343–351.]
2003; Gonzalez-Meler and Taneva,[ Plant respiration. Dordrecht: Kluwer
Academic Publishers, in press.] 2004), a minor inhibition of enzymatic
activity by [CO2] will not translate to the overall tissue respiration rate. A
small increase in enzyme levels in plants exposed to elevated [CO 2] could be
enough to compensate for the direct effects of CO2 on mitochondrial enzyme
activity, as it has been seen in leaves of some plants grown at high [CO 2]
(i.e. Griffin,[ Proceedings of the National Academy of Science of the USA
98:2473–2478.] 2001).
Another factor to consider is that increases in alternative pathway activity
upon inhibition of cytochrome oxidase by a change in [CO2] will result in
unaltered dark respiratory rates (as illustrated in Fig. 1). However, an
increase in the non-phosphorylating activity of the alternative pathway
reduces the efficiency with which energy from oxidized substrates supports
growth and maintenance of plants (Gonzalez-Meler and Siedow,[ Tree
Physiology 19:253–259.] 1999; Gonzalez-Meler and Taneva,[ Plant
respiration. Dordrecht: Kluwer Academic Publishers, in press.] 2004). Some
studies have reported that the growth of plants exposed to elevated [CO 2]
only during the night-time is altered (Bunce,[ Annals of Botany 75:365–368.]
1995, 2001, 2002; Reuveni,[Annals of Botany 80:539–546.]
1997; Griffin,[Plant, Cell and Environment 22:91–99.] 1999). Such altered
growth patterns have been attributed, in part, to effects of [CO2] on plant
respiration, including increased activity of the non-phosphorylating
alternative pathway. These CO2 effects are considered next.

26 | P a g e
ARE THERE DIRECT EFFECTS OF CO2 ON WHOLE
PLANTS?

In the past, in some studies, plants have been grown at [CO 2] elevated only
at night-time to assess the long-term consequences of the previously
considered direct effects of CO2 on respiration rates. Although most of these
studies have found that growth patterns were altered in plants exposed to high
night-time [CO2], these effects cannot be attributed to direct effects of [CO 2]
on respiration for reasons explained above. Bunce[Annals of Botany 75:365–
368.] (1995) observed that the biomass and leaf area ratio of Glycine
max increased and that photosynthetic rates decreased in plants exposed to
high night-time [CO2] relative to plants grown at normal ambient [CO 2].
Similarly, Griffin[ Plant, Cell and Environment 22:91–99] (1999) found
that Glycine max exposed to high night-time [CO2] had lower leaf respiration
rates and greater biomass than plants grown at ambient or elevated
[CO2]. Reuveni[Annals of Botany 80:539–546] (1997) speculated that
increases in the biomass of Lemna gibba grown at high night-time [CO 2]
relative to control plants, was due to a reduction in alternative pathway
respiration (although the alternative pathway activity was not measured).
Reduction in alternative pathway activity might more fully couple respiration
rates with growth and maintenance, enhancing growth. In contrast, Ziska and
Bunce[Australian Journal of Plant Physiology 26:71–77] (1999) showed that
elevation of night-time [CO2] reduced biomass only in two of the four
C4 species studied. In a later study, Bunce[Annals of Botany 90:399–
403](2002) described that carbohydrate translocation was reduced within 2
d of exposing plants to elevated night-time [CO2] when compared with plants
grown at ambient conditions day and night. These results suggest that
elevated [CO2] may have other uncharacterized direct effects on plant
physiology that can have the consequence of reducing energy demand for
carbohydrate translocation, hence reducing the rate of leaf respiration. If so,
these new types of effects cannot be catalogued as direct effects of [CO 2] on

27 | P a g e
respiration but as indirect effects, making more research in this area
necessary.

In summary, even though rapid changes in [CO2] can inhibit the activity of
some mitochondrial enzymes directly, previously reported direct effects of
[CO2] on tissue respiration are likely to be due to measurement artefacts.
Therefore, direct effects of [CO2] on specific respiration rates (although not
necessarily on respiration physiology) should be dismissed as having a major
impact on the amount of anthropogenic carbon that vegetation could retain.
The role of the alternative pathway in direct respiratory responses to elevated
[CO2], although unresolved, would have little impact, if any, on the general
conclusion that direct effects of [CO2] on plant respiration are not to be
considered in plant growth or carbon cycle models. Other effects of long-
term elevation of night-time [CO2] can alter plant growth characteristics, and
indirectly affect both respiration rates and the relative activity of the
cytochrome and alternative pathways. Therefore these effects cannot be
referred to as direct effects but rather as indirect effects of [CO2] on
respiration rates.

INDIRECT EFFECTS

Indirect CO2 effects represent changes in tissue respiration in response to


plant growth at elevated atmospheric [CO2] (Amthor,[ Advances in carbon
dioxide effects research. American Society of Agronomy Special
Publication ] 1997), because of changes in tissue composition
(Drake,[Annual Review of Plant Physiology and Plant Molecular Biology
48:609–639.] 1997). Other indirect effects of [CO2] on plant respiration
include changes in growth or response to environmental stress, as well as
changes in the respiratory demand for energy relative to that of plants grown

28 | P a g e
at ambient [CO2] (Bunce,[ Physiologia Plantarum 90:427–430.]
1994; Amthor,[Annals of Botany 86:1–20.] 2000b). Such indirect effects of
plant growth at elevated [CO2] can be detected as a reduction in
CO2 emission (or O2 consumption) from plant tissues when measured at
ambient [CO2] (Gonzalez-Meler,[effects on coniferous forests and
grasslands. SCOPE, Vol. 56. UK: John Wiley & Sons, 161–181] 1996a).

INDIRECT EFFECTS ON ENZYMES: IS THERE


ACCLIMATION OF RESPIRATION TO ELEVATED [CO 2]?

Acclimation of respiration to high [CO2] can be defined as the down- or up-


regulation of the respiratory machinery (i.e. amount of respiratory enzymes,
number of mitochondria) irrespective of changes in specific respiration rates
when plants are grown at elevated [CO2] (Drake,[ Plant, Cell and
Environment 22:649–657.] 1999). A change in the leaf respiratory machinery
of plants grown at high [CO2] can be expected because of (a) increases in
carbohydrate content, (b) reduced photorespiratory activity and (c)
reductions in leaf protein content (15 % on average), when compared with
plants grown at current ambient [CO2] (Drake,[ Annual Review of Plant
Physiology and Plant Molecular Biology 48:609–639] 1997). These three
processes can have different and sometimes opposite effects on levels of
respiratory enzymes and specific rates of respiration.
As atmospheric [CO2] rises, increased photosynthesis results in higher
cellular carbohydrate concentrations (Drake,[ Annual Review of Plant
Physiology and Plant Molecular Biology 48:609–639.] 1997; Curtis and
Wang,[ Oecologia 113:299–313] 1998). Increased carbohydrates can
stimulate the specific activity of respiration, due to the greater availability of
respiratory substrates (Azcón-Bieto and Osmond,[ Plant Physiology 71:574–
581] 1983) and higher energy demand for phloem loading of carbohydrates
(Bouma,[ Journal of Experimental Botany 46:1185–1194.]

29 | P a g e
1995; Körner,[Plant, Cell and Environment 18:595–600] 1995; Amthor,[
Annals of Botany 86:1–20] 2000b). Additionally, increased tissue
carbohydrate levels (as in the case of plants grown at high [CO2]) could result
in an increase of transcript levels of cytochrome oxidase (Felitti and
Gonzalez,[ Annals of Botany 86:1–20] 1998; Curi,[Plant Physiology and
Biochemistry 41:689–693] 2003) and cytochrome pathway activity
(Gonzalez-Meler,[Plant, Cell and Environment 24:205–215] 2001). In
contrast, a reduction in photorespiratory activity in plants grown at high
[CO2] will reduce the need for the mitochondrial compartment, which may
result in a reduction in mitochondrial proteins and functions (Amthor,[
American Society of Agronomy Special Publication ] 1997; Drake,[ Plant,
Cell and Environment 22:649–657] l999; Bloom,[ Proceedings of the
National Academy of Sciences of the USA 99:1730–1735] 2002). Not
surprisingly, the available studies show that levels and activity of respiratory
machinery can either increase or decrease in leaves of plants grown at high
[CO2] with no concomitant effects on respiration rates (Fig. 2).
FIG 2

30 | P a g e
Relationship between the relative response of leaf respiration and the leaf
respiratory machinery of plants grown at ambient and elevated [CO2]. Data
extracted from Azcon-Bieto[Plant Physiology 106:1163–1168]
(1994), George[Annales des Sciences Forestieres 53:469–474]
(1996), Gonzalez-Meler[: effects on coniferous forests and grasslands.
SCOPE, Vol. 56. UK: John Wiley & Sons, 161–181] (1996a), Griffin
[ Proceedings of the National Academy of Science of the USA 98:2473–
2478] (2001, 2004), Hrubec[Plant Physiology 79:684–689]
(1985), Tissue[Tree Physiology 22:1157–1166] (2002), Wang
[Environmental and Experimental Botany 51:21–31] (2004) and M. A.
Gonzalez-Meler (unpubl. res.). *Respiratory machinery refers to changes in
either number of mitochondria, or soluble or membrane enzyme activity in
plant extracts from plants grown at either ambient or elevated [CO2] in field
and laboratory experiments. Only studies that have provided both respiration
rates and respiratory machinery from the same tissues or from different
tissues from the same plants are considered. Rates of [CO2] evolution used
are only those measured for plants grown under ambient or elevated [CO2]
at ambient [CO2] conditions.
For instance, indirect effects of elevated [CO2] on respiration of leaves
of Lindera benzoin and stems of Scirpus olneyi were correlated with a
reduction in maximum activity of cytochrome oxidase (Azcon-Bieto,[ Plant
Physiology 106:1163–1168] 1994). However, a reduction of respiratory
enzyme activity was not seen in rapidly growing tissues exposed to elevated
[CO2] (Perez-Trejo,[ Plant Physiology 67:514–517] 1981; Hrubeck,[Plant
Physiology 79:684–689] 1985). No acclimation to elevated [CO2] (i.e.
reduction in cytochrome oxidase) has been observed in leaves of C 4 plants
(Azcon-Bieto,[ Plant Physiology 106:1163–1168] 1994) or roots of C3 plants
(Gonzalez-Meler,[SCOPE, Vol. 56. UK: John Wiley & Sons, 161–181]
1996a). In contrast, the number of mitochondria is consistently shown to
increase in response to plant growth at elevated [CO2] (Griffin,[Proceedings
of the National Academy of Science of the USA 98:2473–2478]
2001; Wang,[ Environmental and Experimental Botany 51:21–31] 2004).

31 | P a g e
Overall changes associated with increases or decreases in the respiratory
machinery in leaves of plants grown at elevated [CO 2] (i.e. mitochondrial
counts, cytochrome oxidase maximum activity) seem to be independent of
responses of the respiration rates to elevated [CO 2] (Fig. 2; r2 = 0·, n.s.).
Therefore, changes in the mitochondrial machinery of plants grown in high
[CO2] are associated with altered mitochondrial function and biogenesis and
are not necessarily related to energy production. Anaplerotic functions of
mitochondrial activity provide reductant and carbon skeletons for the
production of primary and secondary metabolites used in growth and
maintenance processes, as well as other processes such as nitrogen reduction
and assimilation (Gonzalez-Meler,[ SCOPE, Vol. 56. UK: John Wiley &
Sons, 161–181] 1996a; Amthor,[ Advances in carbon dioxide effects
research] 1997; Bloom,[Proceedings of the National Academy of Sciences
of the USA 99:1730–1735] 2002; Davey,[ Plant Physiology 134:520–527]
2004). Vanoosten,[ Plant Physiology and Biochemistry 30:541–547]
(1992) reported increases in fumarase and malic enzyme activities in leaves
of Picea abies grown in high [CO2]. These increases contrasted with
unchanged or decreased activity of PEPC or glycolitic enzymes. Increase in
enzyme activity of the tricarboxylic acid cycle with no increase in respiration
rates may support the export of carbon skeletons and reducing power (mainly
as malate) from the mitochondria to the cytosol for biosynthesis. This
uncoupling between an increase in mitochondrial numbers and no increase
in respiratory activity of plants grown at high [CO2] (Fig. 2) is further
illustrated by the study of Tissue[Tree Physiology 22:1157–1166] (2002) in
which a more than two-fold increase in the leaf mitochondrial counts did not
alter the maximum activity of cytochrome oxidase in the same tissues. With
the exception of two studies (i.e. Azcon-Bieto[Plant Physiology 106:1163–
1168], 1994; Tissue,[ Tree Physiology 22:1157–1166] 2002), there are no
reports on the response of membrane-associated mitochondrial enzymes to
elevated [CO2] in leaves or roots of plants. More research is needed in this
area.

32 | P a g e
INDIRECT EFFECTS OF [CO2] ON RESPIRATION OF
TISSUES AND WHOLE PLANTS

It has been historically accepted that leaf respiration is reduced as a


consequence of plant growth at elevated [CO2] (El Kohen,[ Academie des
sciences, serie III 312:477–481] 1991; Isdo and Kimball,[Plant Physiology
99:341–343] 1992; Wullschleger,[ Tree Physiology 10:21–31]
1992a; Amthor,[ American Society of Agronomy Special Publication]
1997; Drake,[ Annual Review of Plant Physiology and Plant Molecular
Biology 48:609–639] 1997; Curtis and Wang,[ Oecologia 113:299–313]
1998; Norby,[ Plant, Cell and Environment 22:683–714] 1999). Poorter
[Australian Journal of Botany 40:501–513](1992) showed that leaf
respiration was reduced, on average, by 14 % when expressed on a leaf mass
basis, but increased by 16 % on a leaf area basis. Curtis and Wang[Oecologia
113:299–313] (1998) compiled respiratory data for woody plants, and
observed that growth at elevated [CO2] resulted in an 18 % inhibition of leaf
respiration (mass basis). In this section we will concentrate on the effects of
elevated [CO2] on respiration of above-ground tissues. It is worth noting,
however, that root respiration rates in the field are not altered by growth at
elevated [CO2] during most of the plant's growing season (Johnson[Plant and
Soil 165:129–138] 1994; Matamala and Schlesinger,[Global Change
Biology 6:967–979] 2000). Total ecosystem root respiration may increase if
root mass and/or turnover increases (Hungate,[ Nature 388:576–579]
1997; Hamilton[Oecologia 131:250–260], 2002; George,[ New Phytologist
160:511–522] 2003; Matamala,[ Science 302:1385–1387] 2003).

Many previous reviews and meta-analyses (see above) have compared


respiratory rates of plants grown and measured at the [CO 2] at which plants
were grown. Therefore, these respiration measurements are also susceptible

33 | P a g e
by the gas exchange artefacts described by Jahnke and co-workers.
Respiratory O2 uptake measured in closed systems is not susceptible to the
measurement artefacts described above. The O2 uptake of green tissues of
plants grown at high [CO2] could be increased, reduced or unaltered when
compared with plants grown at ambient [CO2] (Table 1). Davey[Plant
Physiology 134:520–527](2004) also showed that the O2 uptake of leaves of
plants grown at high [CO2] was slightly increased relative to control plants.
Gas exchange leaks (Jahnke[Plant, Cell and Environment 24:1139–1151],
2001) should not be a factor in determining the rates of CO2 emission rates
when rates are measured and compared at ambient [CO2] for plants grown at
ambient and elevated [CO2]. A re-analysis of the literature focused on leaf
respiratory responses (on a leaf mass basis) of plants grown at ambient and
elevated [CO2] and measured at an ambient [CO2], suggests that specific leaf
respiration rates will be unaltered or even increased in plants grown at
elevated [CO2] (Table 2). Therefore, the generally accepted onclusion that
respiration rates of plants grown at elevated [CO2] is reduced relative to
plants grown at ambient [CO2] should be re-evaluated (Amthor,[Tree
Physiology 20:139–144] 2000a; Davey [Plant Physiology 134:520–
527]2004; Gonzalez-Meler and Taneva[ Plant respiration. Dordrecht:
Kluwer Academic Publishers, in press] 2004). However, there is significant
variability in the leaf respiratory response to growth at elevated [CO2] when
compared with plants grown at ambient conditions, ranging from 40 %
inhibition (Azcon-Bieto[Plant Physiology 106:1163–1168] 1994) to 50 %
stimulation (Williams[SPB Academic Publishing, 547–551] 1992).
Considerations on the physiological basis by which the acclimation response
of respiration to elevated [CO2] varies, is considered next.

TABLE 1.
Indirect effects of [CO2] on the dark O2 uptake rates (on a leaf area basis) at
25 °C of green tissues from Scirpus olneyi (C3), Spartina
patens (C4), Distchilis spicata (C4), Lindera benzoin (C3), Cinna
arundinacea (C3), Quercus geminata (C3), Quercus myrtifolia (C3)

34 | P a g e
and Serenoa repens (C3) plants grown at either normal ambient (A) or normal
ambient + 340 mL L−1 atmospheric CO2 (E), inside open-top chambers in the
field
Species Dark O2 uptakea Dry weight per unit areaa

Saltmarsh

Scirpus olneyi 0·79** 1·00

Spartina patens 0·94 0·99

Distchilis spicata 1·22* 1·00

Woodland understorey

Lindera benzoin 0·77** 0·96

Cinna arundinacea 1·13 1·05

Scrub oak–saw palmetto

Quercus geminata 1·31* 1·11*

Quercus myrtifolia 0·82** 1·24**

Serenoa repens 0·96 1·02

a E/A.Samples were collected at sunrise during the early summer of 1993.


Oxygen uptake was measured at ambient conditions in a liquid phase oxygen
electrode as described in Azcon-Bieto[Plant Physiology 106:1163–1168]
(1994).

35 | P a g e
Significant differences (Student's t-test) between respiratory rates of plants
grown at ambient and elevated [CO2] at P < 0·1 (*) and P < 0·05 (**) are
indicated.

TABLE 2.
Indirect effects of the elevation in atmospheric [CO2] on the dark CO2 efflux
of leaves on a dry mass basis based on literature surveys

No. of
Reference E/A species Observations

Amthor 0·96 21 26 studies, crop and herbaceous and


(1997) woody wild species

Davey et al. 1·07 7 1 study, crop and herbaceous and


(2004) woody wild species

Drake et al. 0·95 17 15 studies, crop and herbaceous wild


(1997) species

Elevated-over-ambient (E/A) refers to the ratio of rate of leaf dark respiration


of plants grown in elevated CO2 to the rate of plants grown in current ambient
CO2, when measured at a common CO2 concentration. Under these
conditions, effects of gas exchange leaks should not be affecting the
comparison of rates of respiration from plants grown at ambient and elevated
CO2.
In boreal species, reduction of respiration of plants grown at high [CO 2] was
related to changes in tissue N and carbohydrate concentration
(Tjoelker[Plant, Cell and Environment 22:767–778] 1999). Tissue N
concentration often decreases in plants grown at elevated [CO 2]

36 | P a g e
(Drake[Annual Review of Plant Physiology and Plant Molecular Biology
48:609–639] 1997; Curtis and Wang[Oecologia 113:299–313] 1998). It is
expected that respiration rate would be lower in tissues having lower [N],
because the respiratory cost associated with protein turnover and
maintenance is a large portion of dark respiration (Bouma[Physiologia
Plantarum 92:585–594] 1994). Hence, the metabolic cost (i.e. respiratory
energy demand) for construction and maintenance of tissues with high
protein concentration is greater than the cost for the maintenance of the same
tissue with low [N] (assuming there are no changes in rates of protein
turnover between plants grown at ambient and elevated [CO2])
(Amthor[Respiration and crop productivity. New York: Springer Verlag.]
1989; Drake[Plant, Cell and Environment 22:649–657] 1999). This idea was
confirmed for leaves of Quercus alba seedlings grown in open-top chambers
in the field, where respiration was 21–56 % lower under elevated [CO2] than
under ambient [CO2] (Wullschleger and Norby [Canadian Journal of Forest
Research 22:1717–1721] 1992). The growth respiration component of these
leaves was reduced by 31 % and the maintenance component by 45 %. These
effects were attributed to the reduced cost of maintaining tissues having
lower [N]. Similar results were obtained in leaves and stems of plants of other
species (Wullschleger [Tree Physiology 10:21–31] 1992a, b [New
Phytologist 121:515–523], 1997; Amthor,[New Phytologist 128:443–450]
1994; Carey[Tree Physiology 16:125–130] 1996; Griffin[Plant, Cell and
Environment 19:729–738] 1996b; Dvorak and Oplustilova[ Dordrecht:
Klurer Academic Publishers, 47–51.] 1997; Will and
Ceulemans[Physiologia Plantarum 100:933–939] 1997), with some
exceptions (e.g. Wullschleger[Physiologia Plantarum 93:47–54] 1995).

The growth component of respiration (Amthor[Annals of Botany 86:1–20]


2000b) could also be reduced in plants grown at elevated [CO 2] as a result of
altered tissue chemistry (Griffin[Oecologia 95:575–580] 1993). Based on the
chemical composition of tissues, Poorter [Plant, Cell and Environment
20:472–482] (1997) found that elevated [CO2] could reduce growth
37 | P a g e
construction costs by 10–20 %. Griffin[Oecologia 95:575–580]
(1993, 1996a) also observed reductions in construction costs of Pinus
taeda seedlings grown at elevated [CO2]. Hamilton [Plant, Cell and
Environment 24:975–982] (2001) reported that elevated [CO2] slightly
reduced construction costs of leaves of mature trees (including P. taeda) at
the top of the canopy, but not at the bottom of the canopy. Such a small
reduction can be explained by reductions in tissue [N], as observed in leaves
exposed to high [CO2] at the top of the canopy. Changes in construction costs
did not result in a decrease in the leaf respiration rates of P. taeda trees
exposed to elevated [CO2] (Hamilton [Plant, Cell and Environment 24:975–
982] 2001).

The lack of long-term effects of increased [CO2] on specific plant respiration


rates could also be due to a lower involvement of the alternative pathway
(Gonzalez-Meler and Siedow[Tree Physiology 19:253–259] 1999; Griffin
[Plant, Cell and Environment 22:91–99] 1999). Respiration through the
alternative pathway bypasses two of the three sites of proton translocation;
so the free energy released is lost as heat, and is unavailable for the synthesis
of ATP. Respiration associated with this pathway will not support growth
and maintenance processes of tissues as efficiently as respiration through the
cytochrome path. On the other hand, the activity of the alternative pathway
of respiration could decrease upon doubling [CO2], masking increases in the
activity of the cytochrome pathway and making respiration more efficient,
as it was the case in understorey trees grown under elevated [CO 2]
(Gonzalez-Meler and Taneva[Plant respiration. Dordrecht: Kluwer
Academic Publishers, in press.] 2004).

Despite earlier reports, the responses of plant respiration to growth under


high [CO2] are variable and perhaps species specific, although the overall
trend may be a moderate increase in respiration rates of leaves of plants
grown under elevated [CO2] relative to the ambient ones (Tables 1 and 2).
Ultimately, altered specific respiration rates of tissues will depend on the net

38 | P a g e
balance between the demand for ATP from maintenance (including phloem
loading and unloading) and growth processes and on carbon allocation
patterns between sinks and source tissues. The fact that the mitochondrial
machinery has been shown to increase in leaves of plants grown at elevated
[CO2] suggests, however, a larger participation of mitochondria in functions
other than oxidative phosphorylation. Finally, indirect effects of [CO 2] on
tissue respiration rates can be augmented or offset by changes in plant size
and/or changes in carbon allocation between plant parts.

INTEGRATED EFFECTS OF ELEVATED [CO2] ON PLANT


RESPIRATION AT THE ECOSYSYTEM LEVEL

Terrestrial ecosystems exchange about 120 Gt of carbon per year with the
atmosphere, through the processes of photosynthesis (leading to gross
primary production, GPP) and ecosystem respiration (Re)
(Schlesinger[Biogeochemistry: an analysis of global change, 2nd edn] 1997).
The difference between GPP and Re determines net ecosystem productivity
(NEP), the net amount of carbon retained or released by a given ecosystem.
Currently, the net exchange of C between the terrestrial biosphere and the
atmosphere is estimated to result in a global terrestrial sink of about 2 Gt of
carbon per year (Gifford[Australian Journal of Plant Physiology 21:1–15]
1994; Schimel[Global Change Biology 1:77–91] 1995; Steffen [Science
280:1393–1394] 1998). Unfortunately, the effects of [CO2] on plant and
heterotrophic respiration at the ecosystem level are not well understood,
despite their potential to control ecosystem carbon budgets
(Ryan[ Ecological Applications 1:157–167] 1991; Giardina and
Ryan[Nature 404:858–861] 2000; Valentini [Nature 404:861–865] 2000).

39 | P a g e
Annually, terrestrial plant respiration releases 40–60 % of the total carbon
fixed during photosynthesis (Gifford[Australian Journal of Plant Physiology
21:1–15] 1994; Amthor[Global Change Biology 1:243–274] 1995)
representing about half of the annual input of CO2 to the atmosphere from
terrestrial ecosystems (Schlesinger [Biogeochemistry: an analysis of global
change, 2nd edn] 1997). Therefore the magnitude of terrestrial plant
respiration and its responses to [CO2] are important factors governing the
intrinsic capacity of ecosystems to store carbon. Plant respiration responses
to high [CO2] may stem from several mechanisms (in absence of direct
effects on respiration rates): (a) indirect effects; (b) changes in total plant
biomass; and (c) changes in plant carbon allocation. If it is confirmed that
the response of overall terrestrial plant respiration to an increase in
atmospheric [CO2] is small (Tables 1 and 2), then changes in global plant
respiration should be proportional to changes in biomass. However,
experimental evidence suggests otherwise, or that the response of plant
respiration at the ecosystem level to elevated [CO2] may be more a function
of carbon allocation patterns rather than just of increases in plant size (Drake
[Plant and Soil 187:111–118] 1996; Hamilton [Oecologia 131:250–260]
2002).
Attempts to scale [CO2] effects on mitochondrial or tissue respiration to the
ecosystem level are problematic because, unlike photosynthesis, little is
known about applicable scaling methods for plant respiration (Gifford
[Functional Plant Biology 30:171–86] 2003). Current attempts to build
respiratory carbon budgets at the canopy level require knowledge of
maintenance and growth respiration (see above), as well as tissue respiratory
responses to light, temperature and [CO2] (Amthor [Annals of Botany 86:1–
20] 2000b; Gifford [Functional Plant Biology 30:171–86] 2003; Turnbull
[Functional Ecology 17:101–114] 2003) or have been based on theoretical
respiration-to-photosynthesis ratios (i.e. Norby [Ecological Applications
12:1261–1266] 2002). The respiration-to-photosynthesis ratio was not
affected by elevated [CO2] in soybean plants (Ziska and Bunce [Global
Change Biology 4:637–643] 1998) but was reduced in pine forests stands

40 | P a g e
(Hamilton [Oecologia 131:250–260] 2002) when compared with ambient
controls. The limited available data on the effects of elevating the [CO2] on
intact field ecosystems on plant and total ecosystem respiration are
summarized in Tables 3 and 4.

TABLE 3.
Contribution of CO2-induced reduction in ecosystem respiration (Re) to total
net ecosystem productivity in a C3-dominated salt-marsh ecosystem exposed
to elevated [CO2] since 1988

Year NCEa Rea NEPa % NEP gain from change in Re

1994 31 −57 66 33

1995 30 −36 63 21

a
% CO2 stimulation.
Relative change in CO2 stimulation was calculated form the annual canopy
flux at elevated plots over that of the ambient ones (n = 5) on a ground area
basis.
Modified from Drake [Plant and Soil 187:111–118] (1996) and M. A.
Gonzalez-Meler (unpubl. res.).

TABLE 4.
The effect of elevated [CO2] on ecosystem-level plant gas exchange of
forested ecosystems
[CO2] increase (μL L−1) GPPa Raa NPPa Reference

200 1·18 0·94 1·27 Hamilton et al. (2002)

41 | P a g e
[CO2] increase (μL L−1) GPPa Raa NPPa Reference

200 1·15 1·09 1·24 Schaefer et al. (2003)

150 1·21 1·21 1·21 Norby et al. (2002)

400 1·36 1·41 1·24 R. J. Trueman (unpubl. res.)

a
E/A.
The ratios of elevated (E) over ambient (A) rates of gross primary
productivity (GPP), stand-level plant respiration (Ra) and net primary
productivity (NPP) were extracted from Hamilton [Oecologia 131:250–260]
(2002), Norby [Ecological Applications 12:1261–1266] (2002), Schaefer
[Global Change Biology 9:1378–1400] (2003) and R. J. Trueman and M. A.
Gonzalez-Meler (unpubl. res.).

The contribution of CO2-induced changes in ecosystem respiration to annual


NEP of an intact salt marsh exposed to elevated [CO 2] in open-top chambers
is shown in Table 3. In this ecosystem, elevation of atmospheric CO2 by 350
μL L−1 over ambient levels consistently reduced night-time ecosystem
respiration in C3 and C4 community stands (Drake [Plant and Soil 187:111–
118] 1996). Net ecosystem productivity increased in stands exposed to
elevated [CO2] because the decrease in respiration was accompanied by
increases in GPP. The reduction in annual respiration in this ecosystem
represented from one-fifth to one-third of the difference in NEP between
ambient and elevated open-top chambers (Table 3). Ecosystem-level plant
respiration in forest stands, however, may not be reduced by elevated [CO 2]
as in the case of the salt marsh (Table 4).

42 | P a g e
In forests exposed to elevated [CO2] using free air CO2 enrichment (FACE),
GPP was stimulated by 15–36 % (Table 4) over ambient stands, resulting in
consistent increases in net primary productivity (NPP). Increases in GPP and
NPP seemed to be accompanied by variable responses on whole-plant
respiration, which could be increased by as much as 20 % (Table 4). For
instance, Hamilton [Oecologia 131:250–260] (2002) found that elevated
[CO2] increased forest NPP by 27 % without any increase in total plant
ecosystem respiration. This result can be explained by a possible decrease in
specific respiration rates at high [CO2] accompanied by larger biomass of
plants grown at elevated [CO2]. In a Popolus deltoides plantation at
Biosphere 2 centre, elevated [CO2] increased stand-level plant respiration by
40 % when compared with ambient (Table 4). The total increase in stand-
level plant respiration was larger than the CO2 stimulation of NPP,
suggesting that, in this case, specific respiration rates and plant size both
contributed to the increase in plant respiration of trees grown in high [CO 2]
(R. J. Trueman and M. A. Gonzalez-Meler, unpubl. res.).

Some studies suggest that increases in ecosystem-level plant respiration in


ecosystems exposed to elevated [CO2] mainly occur in below-ground plant
tissues (Hungate [Nature 388:576–579] 1997; Lin [Plant and Soil 229:259–
270] 2001; Hamilton [Oecologia 131:250–260] 2002), which in turn may
stimulate soil respiration rates (Zak [New Phytologist 147:201–222]
2000; Pendall [Global Biogeochemical Cycles 17:1046] 2003). This
requires a substantial proportion of the additional C assimilated by plants
growing at elevated [CO2] to be allocated to roots for growth and turnover
(Johnson [Plant and Soil 165:129–138] 1994; Hungate [Nature 388:576–
579] 1997; King [Oecologia 128:237–250] 2001; Matamala [Science
302:1385–1387] 2003). Greater plant C allocation below ground can
theoretically increase the contribution of root respiration to total soil
respiration because of greater root biomass relative to ambient [CO 2].
However, Hamilton [Oecologia 131:250–260] (2002) reported increases in
total root respiration in a forest exposed to elevated [CO2], where root mass,
turnover (Matamala [Oecologia 131:250–260] 2003) or respiration rates
(Matamala and Schlesinger [Global Change Biology 6:967–979]

43 | P a g e
2000; George [New Phytologist 160:511–522] 2003) were not affected by
the CO2 treatment. The apparent disparity may be due to the different
methodologies employed in these studies and in the contribution of
rhizosphere activity to soil CO2 efflux, emphasizing the need for more
coordinated research in building ecosystem carbon budgets using multiple
approaches, especially below ground.

In conclusion and contrary to what was previously thought, specific


respiration rates are generally not reduced when plants are grown at elevated
[CO2]. This is because direct effects of [CO2] on respiratory enzymes are
inconsequential to specific respiration rates of tissues. A re-analysis of the
literature comparing respiration of leaves of plants grown at ambient [CO 2]
with leaves of plants grown at elevated [CO2] when rates are measured at the
same [CO2], indicate that leaf respiration on average will not be greatly
changed by increasing atmospheric [CO2]. Increases in growth rates appear
to be compensated for by changes in tissue chemistry that affect growth and
maintenance respiration. If specific rates of respiration are not affected by
growth at elevated [CO2], respiration from the terrestrial vegetation in a high-
CO2 world should be proportional to changes in plant biomass. However,
whole ecosystem studies show that canopy respiration does not increase
proportionally to increases in biomass when natural ecosystems are exposed
to elevated atmospheric [CO2]. Field studies also suggest that a larger
proportion of plant respiration takes place in the root system under elevated
[CO2] conditions. Fundamental information is still lacking on how
respiration and the processes supported by it are physiologically controlled,
thereby preventing sound interpretations of what seem to be species-specific
responses of respiration to elevated [CO2]. Therefore the role of plant
respiration in augmenting the sink capacity of terrestrial ecosystems is still
uncertain.

This review is part of a lecture given at the International Congress of Plant


Mitochondria held in Perth in July 2002, partly sponsored by Annals of
44 | P a g e
Botany. M.A.G.-M. specially thanks Jeff Amthor for his work in this area,
his thoughts, discussions and specific comments in parts of this manuscript.
Authors also wish to thank Joaquim Azcon-Bieto, Bert Drake, Evan
DeLucia, Steve Long, Miquel Ribas-Carbo and Jim Siedow for their
discussion on the effects of elevated [CO2] on respiration of organelles,
plants and ecosystems over the years. We acknowledge financial support by
US Department of Agriculture (M.A.G.-M.), UIC fellowship (L.T.), and
Sigma Xi and provost awards to L.T. and R.J.T.

45 | P a g e
CONCLUSION

Contrary to what was previously thought, specific respiration rates are


generally not reduced when plants are grown at elevated [CO2]. However,
whole ecosystem studies show that canopy respiration does not increase
proportionally to increases in biomass in response to elevated [CO2],
although a larger proportion of respiration takes place in the root system.
Fundamental information is still lacking on how respiration and the processes
supported by it are physiologically controlled, thereby preventing sound
interpretations of what seem to be species-specific responses of respiration
to elevated [CO2]. Therefore the role of plant respiration in augmenting the
sink capacity of terrestrial ecosystems is still uncertain.

46 | P a g e
REFERENCE

1. Affourtit C, Krab K, Moore AL. 2001. Control of plant


mitochondrial respiration. Biochimica Biophysica Acta-
Bioenergetics1504:58–69.

2. Amthor JS. 1989.Respiration and crop productivity. New York:


Springer Verlag.

3. Amthor JS.1991. Respiration in a future, higher CO2 world. Plant,


Cell and Environment 14:13–20.

4. Amthor JS. 1995. Terrestrial higher-plant response to increasing


atmospheric [CO2] in relation to the global carbon cycle. Global
Change Biology 1:243–274.

5. Amthor JS. 1997. Plant respiratory responses to elevated CO2 partial


pressure. In: Allen LH, Kirkham MB, Olszyk DM, Whitman CE,
eds. Advances in carbon dioxide effects research. American Society of
Agronomy Special Publication (proceedings of 1993 ASA Symposium,
Cincinnati, OH), Madison, WI: ASA, CSSA and SSSA, 35–77.

47 | P a g e
6. Amthor JS. 2000. Direct effect of elevated CO2 on nocturnal in
situ leaf respiration in nine temperate deciduous trees species is
small. Tree Physiology 20:139–144.

7. Amthor JS. 2000. The McCree–de Wit–Penning de Vries–Thornley


respiration paradigms: 30 years later. Annals of Botany 86:1–20.

8. Amthor JS, Koch G, Boom AJ. 1992 . CO2 inhibits respiration in


leaves of Rumex crispus L. Plant Physiology 98:1–4.

9. Amthor JS, Koch GW, Willms JR, Layzell DB. 2001. Leaf
O2 uptake in the dark is independent of coincident CO2 partial
pressure. Journal Experimental Botany 52:2235–2238.

10. Amthor JS, Mitchell RJ, Runion GB, Rogers HH, Prior SA,
Wood CW 1994 . Energy content, construction cost and phytomass
accumulation of Glycine max (L.) Merr. and Sorghum bicolor (L.)
Moench grown in elevated CO2 in the field. New Phytologist 128:443–
450.

11. Atkin OK, Tjoelker MG. 2003. Thermal acclimation and the
dynamic response of plant respiration to temperature. Trends in Plant
Science 8:343–351.

12. Azcón-Bieto J, Osmond CB. 1983 . Relationship between


photosynthesis and respiration. The effect of carbohydrate status on the
rate of CO2 production by respiration in darkened and illuminated
wheat leaves. Plant Physiology 71:574–581.

48 | P a g e
13. Azcón-Bieto J, Gonzàlez-Meler MA, Doherty W, Drake
BG. 1994 . Acclimation of respiratory O2 uptake in green tissues of
field grown native species after long-term exposure to elevated
atmospheric CO2. Plant Physiology 106:1163–1168.

14. Bloom AJ, Smart DR, Nguyen DT, Searles PS. 2002
. Nitrogen assimilation and growth of wheat under elevated carbon
dioxide. Proceedings of the National Academy of Sciences of the USA
99:1730–1735.

15. Bouma TJ, De Viser R, Janseen JHJA, De Kick MJ, Van


Leeuwen PH, Lambers H. 1994 . Respiratory energy requirements
and rate of protein turnover in vivo determined by the use of an
inhibitor of protein synthesis and a probe to assess its
effect. Physiologia Plantarum 92:585–594.

16. Bouma TJ, De Viser R, Van Leeuwen PH, De Kick MJ,


Lambers H. 1995 . The respiratory energy requirements involved in
nocturnal carbohydrate export from starch-storing mature source leaves
and their contribution to leaf dark respiration. Journal of Experimental
Botany 46:1185–1194.

17. Bruhn D, Mikkelsen TN, Atkin OK. 2002 . Does the direct
effect of atmospheric CO2 concentration on leaf respiration vary with
temperature? Responses in two species of Plantago that differ in
relative growth rate. Physiologia Plantarum 114:57–64.

49 | P a g e
18. Bunce JA. 1994 . Responses of respiration to increasing
atmospheric carbon dioxide concentrations. Physiologia Plantarum
90:427–430.

19. Bunce JA. 1995 . Effects of elevated carbon dioxide


concentration in the dark on the growth of soybean seedlings. Annals
of Botany 75:365–368.

20. Bunce JA. 2001 . Effects of prolonged darkness on the sensitivity


of leaf respiration to carbon dioxide concentration in C3 and C4
species. Annals of Botany 87:463–468.

21. Bunce JA.2002. Carbon dioxide concentration at night affects


translocation from soybean leaves. Annals of Botany 90:399–403.

22. Carey EV, DeLucia EH, Ball JT. 1996 . Stem maintenance and
construction respiration in Pinus ponderosa grown in different
concentrations of atmospheric CO2. Tree Physiology 16:125–130.

23. Curi GC, Welchen E, Chan RL, Gonzalez DH. 2003 . Nuclear
and mitochondrial genes encoding cytochrome c oxidase subunits
respond differently to the same metabolic factors. Plant Physiology
and Biochemistry 41:689–693.

24. Curtis PS, Wang X. 1998 . A meta-analysis of elevated


CO2 effects on woody plant mass, form, and physiology. Oecologia
113:299–313.

50 | P a g e
25. Davey PA, Hunt S, Hymus GJ, DeLucia EH, Drake BG,
Karnosky DF, Long SP. 2004 . Respiratory oxygen uptake is not
decreased by an instantaneous elevation of [CO2], but is increased with
long-term growth in the field at elevated [CO2]. Plant Physiology
134:520–527.

26. Drake BG, Azcón-Bieto J, Berry JA, Bunce J, Dijkstra P,


Farrar J, Koch GW, Gifford R, Gonzàlez-Meler MA, Lambers
H, et al.1999 . Does elevated CO2 inhibit plant mitochondrial
respiration in green plants? Plant, Cell and Environment 22:649–657.

27. Drake BG, Gonzalez-Meler MA, Long SP. 1997 . More


efficient plants: a consequence of rising atmospheric CO2? Annual
Review of Plant Physiology and Plant Molecular Biology 48:609–639.

28. Drake BG, Meuhe M, Peresta G, Gonzàlez-Meler MA,


Matamala R. 1996. Acclimation of photosynthesis, respiration and
ecosystem carbon flux of a wetland on Chesapeake Bay, Maryland, to
elevated atmospheric CO2 concentration. Plant and Soil 187:111–118.

29. Dvorak V, Oplustilova M. 1997. Respiration of woody tissues


of Norway spruce in elevated CO2 concentrations. In: Mohren GMJ,
Kramer K, Sabate S, eds. Impacts of global change on tree physiology
and forest ecosystems. Dordrecht: Klurer Academic Publishers, 47–51.

30. El Kohen A, Pontailler Y, Mousseau M. 1991 . Effect d'un


doublement du CO2 atmosphérique sur la respiration à l'obscurité des
parties aeriennes de jeunes chataigniers (Castanea
51 | P a g e
sativa Mill.). Compte rendue de l'Academie des sciences, serie III
312:477–481.

31. Felitti SA, Gonzalez DH. 1998 . Carbohydrates modulate the


expression of the sunflower cytochrome c gene at the mRNA
level. Planta 206:410–415.

32. George V, Gerant D, Dizengremel P. 1996 . Photosynthesis,


Rubisco activity and mitochondrial malate oxidation in pedunculate
oak (Quercus robur L.) seedlings grown under present and elevated
atmospheric CO2 concentrations. Annales des Sciences Forestieres
53:469–474.

33. George K, Norby RJ, Hamilton JG, DeLucia EH. 2003 . Fine-
root respiration in a loblolly pine and sweetgum forest growing in
elevated CO2. New Phytologist 160:511–522.

34. Giardina CP, Ryan MG. 2000 . Evidence that decomposition


rates of organic carbon in mineral soil do not vary with
temperature. Nature 404:858–861.

35. Gifford RM. 1994. The global carbon cycle: a view point on the
missing sink. Australian Journal of Plant Physiology 21:1–15.

36. Gifford RM. 2003 . Plant respiration in productivity models:


conceptualization, representation and issues for global terrestrial
carbon-cycle research. Functional Plant Biology 30:171–86.

52 | P a g e
37. Gonzàlez-Meler MA, Siedow JN. 1999 . Inhibition of
respiratory enzymes by elevated CO2: does it matter at the intact tissue
and whole plant levels? Tree Physiology 19:253–259.

38. Gonzàlez-Meler MA, Taneva L. 2004. Integrated effects of


atmospheric CO2 concentration on plant and ecosystem respiration. In:
Lambers H, Ribas-Carbo M, eds. Plant respiration. Dordrecht: Kluwer
Academic Publishers, in press.

39. Gonzàlez-Meler MA, Drake BG, Azcón-Bieto J. 1996 . Rising


atmospheric carbon dioxide and plant respiration. In: Breymeyer AI,
Hall DO, Melillo JM, Ågren GI, eds. Global change: effects on
coniferous forests and grasslands. SCOPE, Vol. 56. UK: John Wiley &
Sons, 161–181.

40. Gonzàlez-Meler MA, Giles L, RB Thomas, Siedow JN. 2001


. Metabolic regulation of leaf respiration and alternative pathway
activity in response to phosphate supply. Plant, Cell and Environment
24:205–215.

41. Gonzàlez-Meler MA, Ribas-Carbo M, Siedow JN, Drake


BG. 1996 . Direct inhibition of plant mitochondrial respiration by
elevated CO2. Plant Physiology 112:1349–1355.

42. Griffin KL, Anderson OR, Gastrich MD, Lewis JD, Lin G,
Schuster W, Seemann JR, Tissue DT, Turnbull M, Whitehead
D. 2001 . Plant growth in elevated CO2 alters mitochondrial number

53 | P a g e
and chloroplast fine structure. Proceedings of the National Academy of
Science of the USA 98:2473–2478.

43. Griffin KL, Anderson OR, Tissue DT, Turnbull MH,


Whitehead D. 2004. Variations in dark respiration and mitochondrial
numbers within needles of Pinus radiata grown in ambient or elevated
CO2 partial pressure. Tree Physiology 24:347–353.

44. Griffin KL, Ball JT, Strain BR. 1996 . Direct and indirect
effects of elevated CO2 on whole-shoot respiration in ponderosa pine
seedlings. Tree Physiology 16:33–41.

45. Griffin KL, Sims DA, Seemann JR. 1999 . Altered night-time
CO2 concentration affects growth, physiology and biochemistry of
soybean. Plant, Cell and Environment 22:91–99.

46. Griffin KL, Thomas RB, Strain BR.1993 . Effects of nitrogen


supply and elevated carbon dioxide on construction cost in leaves
of Pinus taeda (L.) seedlings. Oecologia 95:575–580

47. Griffin KL, Winner WE, Strain BR. 1996 . Construction cost
of loblolly and ponderosa pine leaves grown with varying carbon and
nitrogen availability. Plant, Cell and Environment 19:729–738.

48. Hamilton JG, DeLucia EH, George K, Naidu S, Finzi AC,


Schlesinger WH. 2002 . Forest carbon balance under CO2. Oecologia
131:250–260.

54 | P a g e
49. Hamilton JG, Thomas RB, Delucia EH. 2001 . Direct and
indirect effects of elevated CO2 on leaf respiration in a forest
ecosystem. Plant, Cell and Environment 24:975–982.

50. Hrubec TC, Robinson JM, Donaldson RP. 1985 . Effects of


CO2 enrichment and carbohydrate content on the dark respiration of
soybeans. Plant Physiology 79:684–689.

51. Hungate BA, Holland EA, Jackson RB, Chapin III FS,
Mooney HA, Field CB.1997 . The fate of carbon in grasslands under
carbon dioxide enrichment. Nature 388:576–579.

52. Isdo SB, Kimball BA. 1992 . Effects of atmospheric


CO2 enrichment on photosynthesis, respiration and growth of sour
orange trees. Plant Physiology 99:341–343.

53. Jahnke S. 2001 . Atmospheric CO2 concentration does not


directly affect leaf respiration in bean or poplar. Plant, Cell and
Environment 24:1139–1151.

54. Jahnke S, Krewitt M. 2002 . Atmospheric CO2 concentration


may directly affect leaf respiration measurement in tobacco, but not
respiration itself. Plant, Cell and Environment 25:641–651.

55. Johnson DW, Geisinger D, Walker R, Newman J, Vose JM,


Elliott KJ, Ball T. 1994 . Soil pCO2, soil respiration, and root activity
in CO2-fumigated and nitrogen-fertilized ponderosa pine. Plant and
Soil 165:129–138.

55 | P a g e
56. Kacser H, Burns JA. 1979 . Molecular democracy: who shares
control? Biochemical Society Transactions 7:1149–1160.

57. King JS, Pregitzer KS, Zak DR, Sober J, Isebrands JG,
Dickson RE, Hendrey GR, Karnosky DF. 2001 . Fine-root biomass
and fluxes of soil carbon in young stands of paper birch and trembling
aspen as affected by elevated atmospheric CO2 and tropospheric
O3. Oecologia 128:237–250.

58. Körner CH, Pelaez-Riedl S, Van Bel AJE. 1995


. CO2 responsiveness of plants: a possible link to phloem
loading. Plant, Cell and Environment 18:595–600.

59. Lin G, Rygiewicz PT, Ehleringer JR, Johnson MG, Tingey


DT. 2001 . Time-dependent responses of soil CO2 efflux components
to elevated atmospheric [CO2] and temperature in experimental forest
mesocosms. Plant and Soil 229:259–270.

60. Matamala R, Schlesinger WH. 2000 . Effects of elevated


atmospheric CO2 on fine-root production and activity in an intact
temperate forest ecosystem. Global Change Biology 6:967–979.

61. Matamala R, Gonzalez-Meler MA, Jastrow JD, Norby RJ,


Schlesinger WH. 2003. Impacts of fine root turnover on forest NPP
and soil C sequestration potential. Science 302:1385–1387.

56 | P a g e
62. Norby RJ, Hanson PJ, O'Neill EG, Tschaplinski TJ, Weltzin
JF, Hansen RA, Cheng WX, Wullschleger SD, Gunderson CA,
Edwards NT, et al. 2002 . Net primary productivity of a CO2-enriched
deciduous forest and the implications for carbon storage. Ecological
Applications 12:1261–1266.

63. Norby RJ, Wullschleger SD, Gunderson CA, Johnson DW,


Ceulemans R. 1999 . Tree responses to rising CO2 in field
experiments: implications for the future forest. Plant, Cell and
Environment 22:683–714.

64. Pendall E, Del Grosso S, King JY, LeCain DR, Milchunas


DG, Morgan JA, Mosier AR, Ojima DS, Parton WA, Tans PP, et
al. 2003. Elevated atmospheric CO2 effects and soil water feedbacks
on soil respiration components in a Colorado grassland. Global
Biogeochemical Cycles 17:1046.

65. Perez-Trejo MS. 1981. Mobilization of respiratory metabolism


in potato tubers by carbon dioxide. Plant Physiology 67:514–517.

66. Pinelli P, Loreto F. 2003 . (CO2)-C-12 emission from different


metabolic pathways measured in illuminated and darkened C-3 and C-
4 leaves at low, atmospheric and elevated CO2 concentration. Journal
of Experimental Botany 54:1761–1769.

67. Pons TL, Welschen RAM. 2002 . Overestimation of respiration


rates in commercially available clamp-on leaf chambers.

57 | P a g e
Complications with measurement of net photosynthesis. Plant, Cell
and Environment 25:1367.

68. Poorter H, Gifford RM, Kriedemann PE, Wong SC. 1992 . A


quantitative analysis of dark respiration and carbon content as factors
in the growth response of plants to elevated CO2. Australian Journal of
Botany 40:501-513.

69. Poorter H, VanBerkel Y, Baxter R, DenHertog J, Dijkstra P,


Gifford RM, Griffin KL, Roumet C, Roy J, Wong SC. 1997 . The
effect of elevated CO2 on the chemical composition and construction
costs of leaves of 27 C-3 species. Plant, Cell and Environment 20:472–
482.

70. Reuveni J, Gale J, Mayer AM. 1995. High ambient carbon


dioxide does not affect respiration by suppressing the alternative
cyanide-resistant respiration. Annals of Botany 76:291–295.

71. Reuveni J, Gale J, Zeroni M. 1997 . Differentiating day from


night effects of high ambient [CO2] on the gas exchange and growth
of Xanthium strumarium L. exposed to salinity stress. Annals of
Botany 80:539–546.

72. Ribas-Carbo M, Berry JA, Yakir D, Giles L, Robinson SA,


Lennon AM, Siedow JN. 1995. Electron partitioning between the
cytochrome and alternative pathways in plant mitochondria. Plant
Physiology 109:829–837.

58 | P a g e
73. Ryan MG. 1991. Effects of climate change on plant
respiration. Ecological Applications 1:157–167.

74. Schafer, KVR, Oren R, Ellsworth DS, Lai C, Herricks JD,


Finzi AC, Richter DD, Katul GG. 2003 . Exposure to an enriched
CO2 atmosphere alters carbon assimilation and allocation in a pine
forest ecosystem. Global Change Biology 9:1378–1400.

75. Schimel DS. 1995 . Terrestrial ecosystems and the carbon


cycle. Global Change Biology 1:77–91.

76. Schlesinger WH. 1997 .Biogeochemistry: an analysis of global


change, 2nd edn. San Diego, CA: Academic Press.

77. Steffen W, Noble I, Canadell J, Apps M, Schulze ED, Jarvis


PG, Baldocchi D, Ciais P, Cramer W, Ehleringer J, et al. 1998. The
terrestrial carbon cycle: implications for the Kyoto protocol. Science
280:1393–1394.

78. Tissue DT, Lewis JD, Wullschleger SD, Amthor JS, Griffin
KL, Anderson R. 2002. Leaf respiration at different canopy positions
in sweetgum (Liquidambar styraciflua) grown in ambient and elevated
concentrations of carbon dioxide in the field. Tree Physiology
22:1157–1166.

59 | P a g e
79. Tjoelker MG, Oleksyn J, Lee TD, Reich PB. 2001 . Direct
inhibition of leaf dark respiration by elevated CO2 is minor in 12
grassland species. New Phytologist 150:419–424.

80. Tjoelker MG, Reich PB, Oleksyn J. 1999 . Changes in leaf


nitrogen and carbohydrates underlie temperature and CO2 acclimation
of dark respiration in five boreal species. Plant, Cell and Environment
22:767–778.

81. Turnbull MH, Whitehead D, Tissue DT, Schuster WSF,


Brown KJ, Griffin KL. 2003. Scaling foliar respiration in two
contrasting forest canopies. Functional Ecology 17:101–114.

82. Valentini R, Matteucci G, Dolman AJ, Schulze ED, Rebmann


C, Moors EJ, Granier A, Gross P, Jensen NO, Pilegaard K, et
al. 2000 . Respiration as the main determinant of carbon balance in
European forests. Nature 404:861–865.

83. Vanoosten JJ, Afif D, Dizengremel P. 1992 . Long-term effects


of a CO2 enriched atmosphere on enzymes of the primary carbon
metabolism of spruce trees. Plant Physiology and Biochemistry
30:541–547.

84. Wang XZ, Anderson OR, Griffin KL. 2004 . Chloroplast


numbers, mitochondrion numbers and carbon assimilation physiology
of Nicotiana sylvestris as affected by
CO2 concentration. Environmental and Experimental Botany 51:21–
31.

60 | P a g e
85. Will RE, Ceulemans R. 1997. Effects of elevated
CO2 concentrations on photosynthesis, respiration and carbohydrate
status of coppice Populus hybrids. Physiologia Plantarum 100:933–
939.

86. Williams ML, Jones DG, Baxter R, Farrar JF. 1992 . The
effect of enhanced concentrations of atmospheric CO2 on leaf
respiration. In: Lambers H, van der Plas LHW, eds. Molecular,
biochemical and physiological aspects of plant respiration. The Hague:
SPB Academic Publishing, 547–551.

87. Wullschleger SD, Norby RJ. 1992. Respiratory cost of leaf


growth and maintenance in white oak saplings exposed to atmospheric
CO2 enrichment. Canadian Journal of Forest Research 22:1717–1721.

88. Wullschleger SD, Norby RJ, Gunderson CA. 1992 . Growth


and maintenance respiration in leaves of Liriodendron tulipifera L.
exposed to long-term carbon dioxide enrichment in the field. New
Phytologist 121:515–523.

89. Wullschleger SD, Norby RJ, Hanson PJ. 1995. Growth and
maintenance respiration in stems of Quercus alba after four years of
CO2 enrichment. Physiologia Plantarum 93:47–54.

90. Wullschleger SD, Norby RJ, Hendrix DL 1992 . Carbon


exchange rates, chlorophyll content, and carbohydrate status of two

61 | P a g e
forest tree species exposed to carbon dioxide enrichment. Tree
Physiology 10:21–31.

91. Wullschleger SD, Norby RJ, Love JC, Runck


C. 1997. Energetic costs of tissue construction in yellow-poplar and
white oak trees exposed to long-term CO2 enrichment. Annals of
Botany 80:289–297.

92. Wullschleger SD, Ziska LH, Bunce JA. 1994. Respiratory


response of higher plants to atmospheric CO2 enrichment. Physiologia
Plantarum 90:221–229.

93. Zak DR, Pregitzer KS, King JS, Holmes WE. 2000 . Elevated
atmospheric CO2, fine roots and the response of soil microorganisms:
a review and hypothesis. New Phytologist 147:201–222.

94. Ziska LH, Bunce JA. 1998. The influence of increasing growth
temperature and CO2 concentration on the ratio of respiration to
photosynthesis in soybean seedlings. Global Change Biology 4:637–
643.

95. Ziska, LH, Bunce JA. 1999 . Effects of elevated carbon dioxide
concentration at night on the growth and gas exchange of selected
C4 species. Australian Journal of Plant Physiology 26:71–77.

62 | P a g e

You might also like