Next Article in Journal
Interactions between Mycorrhizal Fungi, Tea Wastes, and Algal Biomass Affecting the Microbial Community, Soil Structure, and Alleviating of Salinity Stress in Corn Yield (Zea mays L.)
Previous Article in Journal
A Cyclic Nucleotide-Gated Channel, HvCNGC2-3, Is Activated by the Co-Presence of Na+ and K+ and Permeable to Na+ and K+ Non-Selectively
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Three Methods of Estimating Mesophyll Conductance Agree Regarding its CO2 Sensitivity in the Rubisco-Limited Ci Range

Adaptive Cropping Systems Lab and PP Systems, USDA-ARS, Haverhill, MA 01913, USA
Plants 2018, 7(3), 62; https://doi.org/10.3390/plants7030062
Submission received: 27 June 2018 / Revised: 24 July 2018 / Accepted: 3 August 2018 / Published: 5 August 2018
(This article belongs to the Special Issue Plant Photosynthetic Gas Exchange: a Current Perspective)

Abstract

:
Whether the mesophyll conductance to CO2 movement (gm) within leaves of C3 plants changes with CO2 concentration remains a matter of debate, particularly at low CO2 concentrations. We tested for changes in gm over the range of sub-stomatal CO2 concentrations (Ci) for which Rubisco activity limited photosynthesis (A) in three plant species grown under the same conditions. Mesophyll conductance was estimated by three independent methods: the oxygen sensitivity of photosynthesis, variable J fluorescence combined with gas exchange, and the curvature of the Rubisco-limited A vs. Ci curve. The latter assay used a new method of rapidly obtaining data points at approximately every 3 μmol mol−1 for Rubisco-limited A vs. Ci curves, allowing separate estimates of curvature over limited Ci ranges. In two species, soybean and sunflower, no change in gm with Ci was detected using any of the three methods of estimating gm. In common bean measured under the same conditions as the other species, all three methods indicated large decreases in gm with increasing Ci. Therefore, change in gm with Ci in the Rubsico-limited region of A vs. Ci curves depended on the species, but not on the method of estimating gm.

1. Introduction

The importance of mesophyll conductance to CO2 movement (gm) within leaves of C3 species in limiting rates of photosynthesis (A) has become increasingly apparent [1,2,3]. Several basically different methods of estimating gm have been developed, including on-line carbon isotope discrimination [4,5], two different types of chlorophyll fluorescence measurements combined with CO2 fixation rates [6], methods based on the curvature of initial A vs. Ci curves [7], and calculation based on the oxygen sensitivity of photosynthesis [8]. All of these methods are based on discrepancies between sub-stomatal CO2 (Ci) values estimated by gas exchange and estimates of CO2 at the site of Rubisco inside the chloroplast (Cc) based on the biochemical C3 photosynthesis model of Farquhar, von Caemmerer and Berry [9]. Several variations on some of the fluorescence-based methods have also been developed. Singh and Reddy [10] compared several methods and some of their variations in soybean leaves, and found little disagreement among the methods compared. Killi and Haworth [11] reported similar gm values obtained from curvature of A vs. Ci curves and from fluorescence measurements combined with gas exchange. However, those comparisons did not deal with possible changes in gm with CO2 concentration. Variation in gm with measurement CO2 concentration has been reported in several cases [5,12,13,14,15,16], but was not found in others [4,6,16]. However, many of these measurements were limited to Ci values > 200 μmol mol−1 because lower Ci values increase errors in the estimate [6]. However, because photosynthesis would be most sensitive to variation in gm in the Rubisco-limited region of low Ci, this work focused on comparing methods in this region of response curves. Some modelling efforts suggested that some apparent variation in gm with CO2 concentration could be artefacts caused by photorespiration [17]. However, Flexas et al. [12] and Mizokami et al. [18] reported changes in gm with CO2 even at low O2, which eliminates photorespiration in these particular cases. Furthermore, an effect of CO2 concentration on aquaporins has been reported [19], potentially providing a mechanism for CO2 impacts on gm [18].
In this paper, three different methods of estimating gm were used to test for changes in gm with Ci over the Rubisco-limited range of Ci in three different plant species. One of the methods used a new method of rapidly generating many densely positioned data points on A vs. Ci curves of single leaves.

2. Results

At 2% O2, values of A at each Ci corrected for the O2 sensitivity of water vapor and carbon dioxide analysis by the LiCor 6400 system software were in close agreement with values of A at the same Ci measured with the CIRAS-3 system without correction for O2 concentration (Figure 1).
Examples of multipoint A vs. Ci curves obtained with the CIRAS-3 CO2 ramping technique are presented in Figure 2. The data define reasonably smooth A vs. Ci curves, and agree well with steady-state A vs. Ci curves. There were enough data points generated rapidly from single leaves to use the Sharkey et al. [20] curve fitting program for at least three separate sections of the Rubisco-limited part of the curve. The analysis program generally requires at least five data points in the Rubisco-limited part of the curve.
In testing the sensitivity of the variable J fluorescence method of estimating gm [6] to values of photorespiration at high Ci, photorespiration was either estimated to be zero at the external CO2 concentration of 1800 μmol mol−1 or to be 2.5% of photosynthesis (see Section 4). The values of gm in the Ci range of 100 to 200 μmol mol−1 estimated from combined gas exchange and fluorescence varied by only a few percent depending on the assumed values of photorespiration. At higher Ci values, those differences in gm estimates would have been more substantial.
Although there was considerable leaf to leaf variation in gm, for both sunflower (Helianthus annuus) and soybean (Glycine max), there was no evidence of a change in gm with Ci using any of the three independent methods of estimating gm (Figure 3), as tested using repeated measures analysis of variance. In contrast with those species, in bean (Phaseolus vulgaris), all three methods of estimating gm indicated a significant decrease in gm with Ci over the 117 to 183 μmol mol−1 range of Ci (Figure 4).

3. Discussion

All three of the methods of estimating gm used here are based on a biochemical model of C3 photosynthesis. However, the sensitivity of each of the three methods to model parameter values differs substantially. For example, the O2 sensitivity method was more sensitive to errors in the CO2 compensation point than to respiration or to Michaelis constants for CO2 or O2 [8], while the variable J fluorescence method was quite sensitive to errors in J, especially when Ci was less than about 80 mol mol−1 [6]. Sensitivity analyses are presented in the original references to the methods. Singh and Reddy [10] compared several variations of fluorescence-based methods with the O2 sensitivity method, and with the estimate based on the curvature of the initial slope of A vs. Ci, and found only minor differences in gm estimates among methods. However, that comparison used soybeans, where gm does not change with Ci.
Many prior measurements of responses of gm to Ci using variable J fluorescence were only considered to produce reliable results at Ci values of about 200 μmol mol−1 and higher, (e.g., [12,13,14]) based on the criterion of Harley et al. [6]. Any changes in gm with Ci in the region of A vs. Ci curves in which Rubisco no longer limits A are less important to A, since A becomes much less sensitive to CO2 availability at high Ci than it is in the Rubisco-limited range [21]. The method of estimating gm based on O2 sensitivity of A is most sensitive in the Rubisco-limited range of A vs. Ci curves [8]. The method based on the curvature of A vs. Ci curves was developed for the Rubisco-limited region [7]. The observations presented here also provided a test of the reliability of the variable J fluorescence method of estimating gm at low Ci in these species. Unstressed leaves of C3 species often operate at Ci values at the upper end of the Rubisco-limited region of A vs. Ci curves [21], with generally lower Ci values in stressed leaves. This makes estimates of gm in the Rubisco-limited region especially relevant.
All three methods of estimating gm used here involve measurements of gas exchange at 21% O2, hence are potentially affected by photorespiration and CO2 transfer among cellular compartments, and are therefore potentially subject to the errors described by Tholen et al. [17]. The constant gm reported here for soybean and sunflower indicate that those potential errors in gm do not always occur even at low measurements Ci, while previous measurements of gm at low O2 indicate that photorespiration is not necessary to observe decreases in gm with increasing Ci [12,18]. Thus, it remains unclear how important the potential artefactual decreases in gm with increasing Ci identified by Tholen et al. [17] may be in general, although they did affect the estimates of gm in soybean and sunflower in this experiment.
The lack of substantial change in gm with Ci in soybean and a decrease in gm with increasing Ci in bean has been previously reported for different cultivars of those species, using only the oxygen sensitivity method [15]. The two additional methods used here qualitatively agreed with those results. While significant intraspecific variation in gm values have been found in soybean, and other species c.f. [22], the variable prior results for sunflower are difficult to reconcile. Vrabl et al. [14] reported strongly decreasing gm with Ci in sunflower both for control leaves and for leaves treated with abscisic acid, while Schaufele et al. [23] found no variation in gm with Ci for unstressed plants, but also found that application of abscisic acid resulted in large decreases in gm with increasing Ci. Qiu et al. [24] also found a large effect of abscisic acid on gm in raspberry. Our results for this sunflower cultivar were similar to those of Schaufele et al. [23], with no change in gm with Ci for unstressed plants over the limited, low Ci range tested here. Several papers Page: 6 (e.g., [12,25,26,27]) have suggested that an influence of aquaporins on CO2 transport and gm might provide a mechanism for changes in gm with Ci, such as observed here in bean. It seems possible that differences among species as to whether gm changes with Ci might be related to the predominance of physical diffusion processes in gm in some species and a larger contribution of metabolic factors in other species. The bean vs. soybean contrast in gm sensitivity to Ci could be a useful experimental system to understand species differences in sensitivity.
The new method described here of ramping CO2 to rapidly obtain A vs. Ci curves could also be useful in many other situations, such as comparisons among genotypes or treatments in photosynthetic parameters. In our experiments, and perhaps in other applications of CO2 ramping, it was prudent to let CO2 increase until no further increase in photosynthesis was observed, even though this took a little extra time. During the ramping, it is very difficult to estimate immediately what the Ci value is at any point in time, so allowing CO2 to saturate photosynthesis ensures that sufficiently high Ci has been achieved. The rate of CO2 increase during ramping can be selected by the user. The rate used here was chosen so that stomatal conductance did not change significantly during the ramping for these species, simplifying post-processing of the data.

4. Materials and Methods

Soybean (Glycine max L. Merr., cultivar Holt), sunflower (Helianthus annuus L., cultivar Mammoth Gray Stripe) and common bean (Phaseolus vulgaris L., cultivar Red Hawk) were grown together in an indoor controlled environment chamber. The chamber air temperature was 25 °C, the dew point temperature was 18 °C, the photosynthetic photon flux density was 1000 μmol m−2 s−1 for 12 h per day from a mixture of metal halide and high-pressure sodium lamps, and CO2 concentration was controlled to 420 ± 20 μmol mol−1 for 24 h per day. Plants were grown one per pot in 15 cm diameter pots filled with vermiculite and flushed daily with a complete nutrient solution. Each measurement of the response of mesophyll conductance response to Ci was made on a different leaf from a different plant.
Most leaf gas exchange measurements were made with a CIRAS-3 photosynthesis system (PP Systems, Amesbury, MA, USA). The leaf cuvette had a 2.5 cm2 window, and light was provided by red, green and blue light-emitting diodes, set for 38% red, 37% green, and 25% blue, as the closest approximation to sunlight. For estimation of gm based on fluorescence, the instrument was equipped for simultaneously measuring chlorophyll fluorescence. Measurements of gm based on the oxygen sensitivity of photosynthesis [8] were made both with the CIRAS-3 and with a LiCor 6400 XT photosynthesis system (LiCor, Inc., Lincoln, NE, USA) in order to test the sensitivity of CO2 and H2O analysis to O2 in the CIRAS-3. The LiCor system had a cuvette window of 6 cm2 area, and used the LiCor red and blue light emitting dioxide light unit. The O2 sensitivity of CO2 and H2O analysis in the LiCor 6400 XT is known, and corrections to outputs based on O2 are built into the instrument operating system. The sensitivity to O2 of the CIRAS-3 system was not known.
Three independent methods of estimating gm were used. The curvature of the A vs. Ci curve in the Rubisco-limited region was used to estimate gm, using the gas exchange calculation utility of Sharkey et al. [20]. In order for this method to be applied for different Rubisco-limited Ci ranges, a new system was developed for rapidly collecting very dense data points. The CIRAS-3 system utility “stored differential-balance” was used to store the change in sensitivity of CO2 and H2O to background CO2 and H2O for the anticipated ranges of each variable. These values are quite stable over time (days), in the CIRAS-3. The CIRAS-3 system utility program which controls linear increases in reference CO2 was set to increase the reference CO2 from 100 μmol mol−1 at a rate of 233 μmol mol−1 min−1 following an initial 2 min period of constant concentration of 100 μmol mol−1. Data were stored approximately every 2 s during the increase in reference CO2. First, this CO2 ramping program was run with an empty cuvette. Then it was run with leaves in the cuvette. Leaf temperature was controlled at 25 ℃, the water vapor pressure deficit (VPD) was between 1.0 and 1.5 kPa, and the photosynthetic photon flux density (PPFD) was 1500 μmol m-2 s−1. Measurements on leaves were terminated when the “apparent” A values no longer increased with reference CO2. For each time step in the data files, the “A” value for the empty cuvette was subtracted from the value when the leaf was present. This is similar to the approach of Stinziano et al. [28], although their method of obtaining corrected values of A was more complex, as necessary with the instrument used. From the corrected A values, Ci was re-calculated in the usual way from stomatal and boundary layer conductances, A, and external CO2 [21], using the corrected values of A. Stomatal conductance never changed substantially during the CO2 ramping procedures. As a test of this new method of developing A vs. Ci curves using rapid CO2 ramping, A vs. Ci curves obtained with ramped CO2 were compared with curves on the same leaves obtained with the same instrument using traditional steady-state gas exchange measurements at several steps of external CO2 under the same conditions of PPFD, leaf temperature, and VPD as used in the CO2 ramping. For tests of changes in gm with Ci, A vs. Ci curves obtained by CO2 ramping were arbitrarily separated into four successive sections, from Cis of 100 to 133, 134 to 166, and from Cis of 167 to 200, and all Cis above 200 μmol mol−1. The three lower parts of the low Ci curves were separately combined with the fourth, upper Ci part. The Sharkey et al. utility [20] was then used to estimate gm separately for the three lower Ci sections with uniform upper Ci data, assuming limitation by the maximum carboxylation capacity of Rubisco (VCmax) for the three lower Ci ranges, and triose phosphate utilization (TPU) limitation at the highest Ci. The division between the VCmax and electron transport (J)-limited regions was done by minimizing the error term. The utility program values for temperature dependencies of parameters were used, and the utility was used to estimate respiration. Three leaves from different plants of each species were used to develop these tests of variation of gm with Ci.
The second method used was the “variable J” method combining leaf gas exchange and fluorescence [6]. Steady state A vs. Ci curves were obtained, using sequential external CO2 concentrations of 400, 100, 150, 200, 250, 300, 400, 500, 600, 800, 1000, and 1800 μmol mol−1. Leaves were measured at the same PPFD, leaf temperature (T) and VPD as previously described. Light was supplied by red, green and blue light-emitting diodes, set at 38% red, 37% green, and 25% blue as a close approximation to sunlight. Leaf absorption was assumed to be 0.84, and the fraction absorbed by photosystem II was assumed to be 0.5. At each CO2 level, Phi PSII and J were obtained using Multi-PulseTM fluorescence measurements [29]. The measurement at the highest CO2 (“non-photorespiratory” conditions) was used to determine the proportionality between J and fluorescence yield [6]. Because photorespiration at the highest CO2 concentration could theoretically have still been 2 to 3% of photosynthesis [30], additional estimates of gm were made using adjusted values of the proportionality between J and fluorescence yield assuming photorespiration was 2.5% of photosynthesis. These measurements were made on three leaves from different plants of each species. Values of gm for each leaf were obtained in the same three Ci regions as used in the previous method.
The third method used, the oxygen sensitivity of photosynthesis method of estimating gm [8] was implemented by developing A vs. Ci curves of the same leaf both at 21 and 2% O2. Leaf temperature, VPD, and PPFD were the same as in the prior methods. The A vs. Ci curves at 2% and 21% O2 were used to calculate gm at Ci values ranging from below 100 to about 200 μmol mol−1, by solving for gm values compatible with A vs. Ci values at both O2 concentrations [10]. Values of respiration in the light were determined by extrapolating A vs. Ci curves at 2% O2 to 0 Ci. The calculation utility of Singh and Reddy [10] uses the same other biochemical parameter values as in the Sharkey et al. utility [20]. Values of A vs. Ci at 2% O2 estimated using the LiCor 6400 XT instrument with correction for O2 were compared with A vs. Ci values estimated using the CIRAS-3 instrument without correction for O2 concentration. The comparisons between instruments were made using opposite sides of the same leaves for two individual plants each of sunflower and soybean. The response of gm to Ci using the O2 sensitivity method with the CIRAS-3 system was then determined for three leaves from different plants of each species, and summarized for the same three ranges of Ci as used in the curvature method.
Repeated measures analysis of variance was used separately for each method for each species to test whether gm changed with the Ci range.

5. Conclusions

The results presented here indicate that the variable J fluorescence method of estimating gm may be valid at lower Ci values than often assumed because it agreed with two other methods, which are especially suited to measurements of gm in the Rubisco-limited region of A vs. Ci curves. Whether gm varies with Ci in the Rubisco-limited range depends upon the species, but three different methods of estimating gm were in agreement regarding whether gm changed with Ci or was constant.
* Mention of specific brands of instruments does not constitute endorsement of those brands of instruments by the USDA to the exclusion of others which may be suitable.

Funding

This research received no external funding.

Acknowledgments

I thank Andrew Lintz, PP Systems, for informing me of the capability of the CIRAS-3 system to do high speed CO2 ramping and for developing a preliminary script for the instrument to accomplish it.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Flexas, J.; Ribas-Carbo, M.; Diaz-Espejo, A.; Galmes, J.; Medrano, H. Mesophyll conductance to CO2: Current knowledge and future prospects. Plant Cell Environ. 2008, 31, 602–621. [Google Scholar] [CrossRef] [PubMed]
  2. Warren, C.R. Stand aside stomata, another actor deserves centre stage: The forgotten role of the internal conductance to CO2 transfer. J. Exp. Bot. 2008, 59, 1475–1487. [Google Scholar] [CrossRef] [PubMed]
  3. Von Caemmerer, S.; Evans, J.R. Temperature responses of mesophyll conductance differ greatly between species. Plant Cell Environ. 2015, 38, 629–637. [Google Scholar] [CrossRef] [PubMed]
  4. Tazoe, Y.; von Caemmerer, S.; Badger, M.R.; Evans, J.R. Light and CO2 do not affect the internal conductance to CO2 diffusion in wheat leaves. J. Exp. Bot. 2009, 60, 2291–2301. [Google Scholar] [CrossRef] [PubMed]
  5. Tazoe, Y.; von Caemmerer, S.; Estavillo, G.M.; Evans, J.R. Using tunable diode laser spectroscopy to measure carbon isotope discrimination and mesophyll conductance to CO2 diffusion dynamically at different CO2 concentrations. Plant Cell Environ. 2011, 34, 580–591. [Google Scholar] [CrossRef] [PubMed]
  6. Harley, P.C.; Loreto, F.; Di Marco, D.; Sharkey, T.D. Theoretical considerations when estimating the mesophyll conductance to CO2 flux by analysis of the response of photosynthesis to CO2. Plant Physiol. 1992, 98, 1429–1436. [Google Scholar] [CrossRef] [PubMed]
  7. Ethier, G.H.; Livingstone, N.J. On the need to incorporate sensitivity to CO2 transfer conductance into the Farquhar-von Caemmerer-Berry leaf photosynthesis model. Plant Cell Environ. 2004, 27, 137–153. [Google Scholar] [CrossRef]
  8. Bunce, J. Use of the response of photosynthesis to oxygen to estimate mesophyll conductance to carbon dioxide in water-stressed soybean leaves. Plant Cell Environ. 2009, 32, 875–881. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Farquhar, G.D.; von Caemmerer, S.; Berry, J.A. A biochemical model of photosynthetic CO2 assimilation in leaves of C3 species. Planta 1980, 149, 78–90. [Google Scholar] [CrossRef] [PubMed]
  10. Singh, S.K.; Reddy, V.R. Methods of mesophyll conductance estimation: Its impact on key biochemical parameters and photosynthetic limitations in phosphorus-stressed soybean across CO2. Physiol. Plant 2016, 157, 234–254. [Google Scholar] [CrossRef] [PubMed]
  11. Killi, D.; Haworth, M. Diffusive and metabolic constraints to photosynthesis in quinoa during drought and salt stress. Plants 2017, 6, 49. [Google Scholar] [CrossRef] [PubMed]
  12. Flexas, J.; Diaz-Espejo, A.; Galmes, J.; Kaldenhoff, R.; Medrano, H.; Ribas-Carbo, M. Rapid variations of mesophyll conductance in response to changes in CO2 concentrations around leaves. Plant Cell Environ. 2007, 30, 1284–1298. [Google Scholar] [CrossRef] [PubMed]
  13. Hassiotou, F.; Ludwig, M.; Renton, M.; Veneklaasds, E.J.; Evans, J.R. Influence of leaf dry mass per area, CO2, and irradiance on mesophyll conductance in sclerophylls. J. Exp. Bot. 2009, 60, 2303–2314. [Google Scholar] [CrossRef] [PubMed]
  14. Vrabl, D.; Vaskova, M.; Hronkova, M.; Flexas, J.; Santrucek, J. Mesophyll conductance to CO2 transport estimated by two independent methods: Effect of variable CO2 concentration and abscisic acid. J. Exp. Bot. 2009, 60, 2315–2323. [Google Scholar] [CrossRef] [PubMed]
  15. Bunce, J.A. Variable responses of mesophyll conductance to substomatal carbon dioxide concentration in common bean and soybean. Photosynthetica 2010, 48, 507–512. [Google Scholar] [CrossRef]
  16. Bunce, J. Using new gas exchange methods to estimate mesophyll conductance and non-stomatal inhibition of photosynthesis caused by water deficits. HortScience 2012, 47, 687–690. [Google Scholar]
  17. Tholen, D.; Ethier, G.; Genty, B.; Pepin, S.; Zhu, X.-G. Variable mesophyll conductance revisited: Theoretical background and experimental implications. Plant Cell Environ. 2012, 35, 2087–2103. [Google Scholar] [CrossRef] [PubMed]
  18. Mizokami, Y.; Noguchi, K.; Kojima, M.; Sakakibara, H.; Terashima, I. Mesophyll conductance decreases in the wild type but not in an ABA-deficient mutant (aba1) of Nicotiana plumbaginifolia under drought conditions. Plant Cell Environ. 2015, 38, 388–398. [Google Scholar] [CrossRef] [PubMed]
  19. Mori, I.C.; Rhee, J.; Shibasaka, M.; Sasano, S.; Kaneko, R.; Horie, T.; Dasuhara, M. CO2 transport by PIP2 aquaporins of barley. Plant Cell Physiol. 2014, 55, 251–257. [Google Scholar] [CrossRef] [PubMed]
  20. Sharkey, T.D.; Bernacchi, C.J.; Farquhar, G.D.; Singsaas, E.L. Fitting photosynthetic carbon dioxide response curves for C3 leaves. Plant Cell Environ. 2007, 30, 1035–1040. [Google Scholar] [CrossRef] [PubMed]
  21. Farquhar, G.C.; Sharkey, T.D. Stomatal conductance and photosynthesis. Annu. Rev. Plant Physiol. 1982, 33, 317–345. [Google Scholar] [CrossRef]
  22. Bunce, J. Variation among soybean cultivars in mesophyll conductance and leaf water use efficiency. Plants 2016, 5, 44. [Google Scholar] [CrossRef] [PubMed]
  23. Schaufele, R.; Santrucek, J.; Schnyder, H. Dynamic changes of canopy-scale mesophyll conductance to CO2 diffusion of sunflower as affected by CO2 concentration and abscisic acid. Plant Cell Environ. 2011, 34, 127–136. [Google Scholar] [CrossRef] [PubMed]
  24. Qiu, C.; Ethier, G.; Pepin, S.; Dube, P.; Desjardins, Y.; Gosselin, A. Persistent negative temperature response of mesophyll conductance in red raspberry (Rubus idaeus L.) leaves under both high and low vapour pressure deficits: A role for abscisic acid? Plant Cell Environ. 2017, 40, 1940–1959. [Google Scholar] [CrossRef] [PubMed]
  25. Hanba, Y.T.; Shibasaka, M.; Hayashi, Y.; Hayakawa, T.; Kasamo, K.; Terashima, I.; Datsuhara, M. Overexpression of the varley aquaporin HvPIP2; 1 increases internal CO2 conductance and CO2 assimilation in the leaves of transgenic rice plants. Plant Cell Physiol. 2004, 45, 521–529. [Google Scholar] [CrossRef] [PubMed]
  26. Miyazawa, S.-I.; Yoshimura, S.; Shinzaki, Y.; Maeshima, M.; Miyake, C. Deactivation of aquaporins decreases internal conductance to CO2 diffusion in tobacco leaves grown under long-term drought. Funct. Plant Biol. 2008, 35, 553–564. [Google Scholar] [CrossRef]
  27. Xiong, D.; Liu, X.; Liu, L.; Douthe, C.; Li, Y.; Peng, S.; Huang, J. Rapid responses of mesophyll conductance to changes of CO2 concentration, temperature and irradiance are affected by N supplements in rice. Plant Cell Environ. 2015, 38, 2541–2550. [Google Scholar] [CrossRef] [PubMed]
  28. Stinziano, J.R.; Morgan, P.B.; Lynch, D.J.; Saathoff, A.J.; McDermitt, D.K.; Hanson, D.T. The rapid A-Ci responses: Photosynthesis in the phenomic era. Plant Cell Environ. 2017, 40, 1256–1262. [Google Scholar] [CrossRef] [PubMed]
  29. Earl, H.; Ennahli, S. Estimating photosynthetic electron transport via chlorophyll fluorometry without Photosynthem II light saturation. Photosynth. Res. 2004, 82, 177–186. [Google Scholar] [CrossRef] [PubMed]
  30. Sharkey, T.D. Estimation of the rate of photorespiration in leaves. Physiol. Plant 1988, 73, 147–152. [Google Scholar] [CrossRef]
Figure 1. CO2 assimilation rates (A) of sunflower (Helianthus annuus) leaves at 2% O2 in N2 measured with a LiCor 6400 portable photosynthesis system, using the system corrections for O2 concentration, and measured with a CIRAS-3 portable photosynthesis system at the same sub-stomatal CO2 concentrations, without correction for O2 concentration. The line is the 1:1 line.
Figure 1. CO2 assimilation rates (A) of sunflower (Helianthus annuus) leaves at 2% O2 in N2 measured with a LiCor 6400 portable photosynthesis system, using the system corrections for O2 concentration, and measured with a CIRAS-3 portable photosynthesis system at the same sub-stomatal CO2 concentrations, without correction for O2 concentration. The line is the 1:1 line.
Plants 07 00062 g001
Figure 2. CO2 assimilation rate (A) over a range of sub-stomatal CO2 concentrations (Ci) in single bean (Phaseolus vulgaris), soybean (Glycine max), and sunflower (Helianthus annuus) leaves. Open symbols: steady-state data points. Closed symbols: data points obtained from ramped CO2 on the same leaves. See text for details of methods.
Figure 2. CO2 assimilation rate (A) over a range of sub-stomatal CO2 concentrations (Ci) in single bean (Phaseolus vulgaris), soybean (Glycine max), and sunflower (Helianthus annuus) leaves. Open symbols: steady-state data points. Closed symbols: data points obtained from ramped CO2 on the same leaves. See text for details of methods.
Plants 07 00062 g002
Figure 3. Mesophyll conductance to CO2 (gm) as a function of sub-stomatal CO2 concentration (Ci) in soybean and sunflower. Mesophyll conductance was measured using three methods: the oxygen sensitivity of photosynthesis, variable J fluorescence, and the curvature of the Rubisco-limited A vs. Ci curve. See text for details.
Figure 3. Mesophyll conductance to CO2 (gm) as a function of sub-stomatal CO2 concentration (Ci) in soybean and sunflower. Mesophyll conductance was measured using three methods: the oxygen sensitivity of photosynthesis, variable J fluorescence, and the curvature of the Rubisco-limited A vs. Ci curve. See text for details.
Plants 07 00062 g003
Figure 4. Mesophyll conductance to CO2 (gm) as a function of sub-stomatal CO2 concentration (Ci) in common bean. Mesophyll conductance was measured using three methods: the oxygen sensitivity of photosynthesis, variable J fluorescence, and the curvature of the Rubisco-limited A vs. Ci curve. See text for details.
Figure 4. Mesophyll conductance to CO2 (gm) as a function of sub-stomatal CO2 concentration (Ci) in common bean. Mesophyll conductance was measured using three methods: the oxygen sensitivity of photosynthesis, variable J fluorescence, and the curvature of the Rubisco-limited A vs. Ci curve. See text for details.
Plants 07 00062 g004

Share and Cite

MDPI and ACS Style

Bunce, J. Three Methods of Estimating Mesophyll Conductance Agree Regarding its CO2 Sensitivity in the Rubisco-Limited Ci Range. Plants 2018, 7, 62. https://doi.org/10.3390/plants7030062

AMA Style

Bunce J. Three Methods of Estimating Mesophyll Conductance Agree Regarding its CO2 Sensitivity in the Rubisco-Limited Ci Range. Plants. 2018; 7(3):62. https://doi.org/10.3390/plants7030062

Chicago/Turabian Style

Bunce, James. 2018. "Three Methods of Estimating Mesophyll Conductance Agree Regarding its CO2 Sensitivity in the Rubisco-Limited Ci Range" Plants 7, no. 3: 62. https://doi.org/10.3390/plants7030062

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop