Sie sind auf Seite 1von 9

Journal of Experimental Botany, Vol. 68, No. 3 pp.

359–367, 2017
doi:10.1093/jxb/erw421  Advance Access publication 31 December 2016

OPINION PAPER

Chloride: essential micronutrient and multifunctional


beneficial ion
John A. Raven1,2,*
1 
Division of Plant Science, University of Dundee at the James Hutton Institute, Invergowrie, Dundee DD2 5DA, UK
2 
School of Plant Biology, University of Western Australia MO84, Stirling Highway. Crawley, WA 6009, Australia

Downloaded from https://academic.oup.com/jxb/article-abstract/68/3/359/2768889 by guest on 24 September 2019


*  Correspondence: j.a.raven@dundee.ac.uk

Received 13 May 2016; Editorial decision 20 October 2016; Accepted 20 October 2016

Editor: Christine Raines, University of Essex

Abstract
Cl− is an essential micronutrient for oxygenic photolithotrophs. About half of global primary productivity is carried out
by oxygenic photolithotrophs exposed to saline waters with Cl− concentrations orders of magnitude higher than that
needed to satisfy the micronutrient requirement. The other half of primary productivity involves terrestrial and fresh-
water glycophytes sometimes in environments containing significantly more Cl− than is needed for the micronutrient
requirement, but less than the toxic Cl– concentration for glycophytes. Intracellular Cl− acts in regulation of cell turgor
and volume, including that of stomatal and pulvinar nastic movements, is a major ion in streptophyte and ulvophy-
cean action potentials, and is involved in ion currents flowing around apices of pollen tubes and Acetabularia cells.
More work is needed on the essentiality of Cl− in these processes, as well as the recent finding that Cl− at 1–5 mol m−3
increases water use efficiency of growth and leaf area in Nicotiana tabacum.

Keywords:  Action potential, active transport, beneficial nutrient, chloride, ion channels, micronutrient, pulvini, stomata, turgor
regulation, water use efficiency.

Introduction
Cl− (chloride) is rightly regarded as an essential micronu- Tolerance of high external Cl− (significantly above the ben-
trient for oxygenic photosynthetic organisms, first noted eficial range) is clearly the case for the organisms contribut-
by Arnon and Whatley (1949) (see below), and is generally ing the almost 50% of global annual net primary production
present in the environment at concentrations well in excess that occurs in saline environments (Field et  al., 1998). This
of what is needed to satisfy the micronutrient requirement. is despite the evidence that not only halophytic embryo-
Franco-Navarro et al. (2016) have added the suggestion that phytes (mainly flowering plants) are secondarily salt tolerant
Cl− acts as a beneficial nutrient when supplied to the glyco- (Flowers et al., 2010), but also the cyanobacteria (Blank and
phyte Nicotiana tabacum at concentrations (1–5 mol m−3) in Sanchez-Baracaldo, 2010) and the Archaeplastida; that is,
excess of those needed to satisfy the micronutrient require- those eukaryotic algae with primary chloroplast endosym-
ments, but insufficient to cause toxicity. Herein the evidence biosis (Blank, 2013). Also considered is the range of other
relating to Cl− as an essential micronutrient in oxygenic pho- functions of Cl− in the plant. These processes include turgor
tolithotrophs is analysed, and suggestions for further tests generation used in growth and nastic (stomata, pulvini) move-
are made. ments, and information transfer over tens of micrometres to

© The Author 2016. Published by Oxford University Press on behalf of the Society for Experimental Biology. All rights reserved.
For permissions, please email: journals.permissions@oup.com
360  |  Raven

a metre by electric currents carried by Cl−. Also considered is (Rognes, 1980) and tonoplast V-type H+ ATPase (Churchill
the role of Cl− in charge and acid–base regulation in relation and Sze, 1984). More controversial is the role of Cl− in acti-
to the form in which N is supplied and the influence on water vating amylase activity. Eyster (1952) suggests that plant as
use efficiency (WUE). Finally, the various functions attrib- well as animal amylases are activated by Cl−; D’Amico et al.
uted to Cl− in plants are considered in relation to the possibil- (2000) point out that, although the known Cl−-binding motif
ity that other solutes can substitute for Cl−. is confined to the amylases of metazoan and some Gram-
negative bacteria, an unrelated Cl−-binding site could be
involved in Cl− stimulation of other amylases.
Chloride as an essential micronutrient in Franco-Navarro et al. (2016) built on the role of Cl− as an
essential micronutrient and provided evidence that it also has
photosynthetic organisms
a role at higher external concentrations as a beneficial macro-
It is very difficult to investigate the essentiality of Cl− by nutrient. Their work compared N. tabacum L. plants grown
culturing algae and plants in the absence of Cl−. The high with three anion treatments and constant cation concentra-
solubility of almost all Cl− salts means they are difficult tions compared with the basal solution (BS) with no addi-

Downloaded from https://academic.oup.com/jxb/article-abstract/68/3/359/2768889 by guest on 24 September 2019


to remove from even the most highly purified salts used in tional salts; in all cases, 75 mmol m−3 Cl− was present to satisfy
making up hydroponic culture solutions (Broyer et al., 1954; the micronutrient Cl− requirement. The treatments were ‘CL’
Johnson et  al., 1957; Brownell and Crosland, 1972, 1974). with 5 mol m−3 added KCl, ‘N’ with 5 mol m−3 added KNO3,
Furtheremore, there is atmospheric NaCl input as aerosols and ‘SP’ with 1.25 mol m−3 KH2PO4, 0.625 mol m−3 MgSO4,
produced from the ocean–atmosphere interface by turbulent and 0.625 mol m−3 CaSO4. In all cases, the total cation con-
winds interacting with wave action from the ocean, with 1.8 centrations were 4.62  mol m−3 K+, 2.63  mol m−3 Ca2+, and
Pg Cl− entering the marine boundary layer—the lower 1000 1.63 mol m−3 Mg2+. The CL treatment increased leaf water
m of the troposphere—each year (Winterton, 2000). Brownell content, and decreased leaf osmotic potential (i.e. a higher
and Crossland (1972, 1974) maintained plant growth cham- concentration of solutes in leaf cells), and increased turgor
bers at above atmospheric pressure to minimize inputs of of leaf cells, relative to the other treatments, and increased
atmospheric NaCl in their work on Na+ requirements of (i.e. less negative) leaf water potential. The photosynthetic
plants with photosynthesis using C4 and Crassulacean acid rate and rate of fresh and dry matter production was greatest
metabolism. Johnson et al. (1957) found Cl− concentrations for the N treatment; the stomatal conductance and transpira-
in 11 plant species that were 3.3–44.4 times the Cl− that could tion rate were lowest, and the instantaneous WUE (mol CO2
be accounted for as contamination in the culture solution, assimilated in photosynthesis per mol H2O lost in transpira-
and the rest was assumed to result from atmospheric Cl− tion) and the growth WUE (g FW gain per m3 water lost)
input. Kirkby and Armstrong (1980) found 150–200  mmol were highest, for the CL treatment. This higher WUE with
per kg dry matter in NO3−-grown Ricinus communis despite CL is typical of the higher WUE found in N-, K-, and (some-
no deliberate addition of Cl− to the culture medium. Eyster times) P-replete conditions relative to deficiency of these
(1958) used the critical Cl− concentration (the minimum nutrients (Woodward, 1699; Raven et al., 2004; Raven, 2013).
external concentration yielding the maximum growth rate) in Franco-Navarro et al. (2016) point out that there are several
studies on the green microalga Chlorella, and the findings did data sets on the effects of Cl− addition, below salinization
not permit conclusions on the role of Cl− as a micronutrient concentrations, on plants (Xu et al., 2000) but that these gen-
in photolithotrophic cultures, although this can be shown for erally have not adequately separated the effects of the cation
chemo-organotrophically grown cells. added with Cl− from the effects of Cl− itself (e.g. Braconnier
The considerations in the last paragraph show that it is and d’Auzac, 1990; Braconnier and Bonneau, 1998). An
difficult to demonstrate the micronutrient role of Cl− from exception, cited in Xu et al. (1999), is Christensen et al. (1981)
whole-organism culture studies. Additional data on the sta- who added (NH4)2SO4 in parallel with NH4Cl in the work on
tus of Cl− as a micronutrient come from work on subcellu- Triticum aestivum L, and found increased leaf turgor pressure
lar systems, and especially a role in PSII of photosynthesis and decreased leaf water and osmotic potentials when NH4Cl
(e.g. Arnon and Whatley, 1949; Winterton, 2000). The role rather than (NH4)2SO4 was added to the crop. The findings
in PSII is in facilitating the proton flux from the water oxi- on turgor pressure and osmotic potential are similar to those
dation complex to the thylakoid lumen; in the absence of of Franco-Navarro et al. (2016) on N. tabacum, but the effects
Cl−, H+ transfer is impeded by formation of a salt bridge on water potential are different.
between lys317 of D2 and asp61 of D1 (Guskov et al., 2009; An important aspect of increased WUE by increased Cl−
Pokhrel et al., 2011; Rivalta et al., 2011; Boussac et al., 2012; (CL treatment) is that a decrease in nitrogen use efficiency
Kato et al., 2012; Wang et al., 2014; Zhang et al., 2014). The (NUE: rate of dry matter gain per unit plant N) is predicted
photosynthetic role may be obscured (Terry, 1977) in less as a result of the decreased stomatal conductance resulting in
resolved systems than were used by Rivalta et al. (2011), a lower steady-state intercellular CO2 concentration and cor-
and PSII is partially inhibited when growth-inhibiting exter- responding decreased CO2 saturation of Rubisco carboxylase
nal Cl− concentrations (decoupled from external Na+) are and increased Rubisco oxgenase activity in C3 plants such as
supplied to whole Phaseolus vulgaris L. plants (Tavakkoli N. tabacum. It may be asked why Cl− supply should regulate
et al., 2010). Other roles for Cl− in plants and algae in acti- the interaction between WUE and NUE when more obvious
vating catalysis by proteins involve asparagine synthetase signals are water availability relative to transpiratory water
Chloride: essential micronutrient and multifunctional beneficial ion  |  361

demand, and N availability, respectively. However, Franco- to the ocean and, by evaporation of confined bodies of water,
Navarro et  al. (2016) point out that there is no decrease in forms halite (NaCl) that yields locally high concentrations
the photosynthetic rate in the CL treatment relative to the in sedimentary rocks (Berner and Berner, 1996; Winterton,
SP or BS treatments, and they suggest that this results from 2000). The mean Cl− in rivers is 0.16 mol m−3 from soil and
an increased mesophyll conductance, compensating for the groundwater derived from natural sources; anthropogenic Cl−
decreased stomatal conductance. If this were the case, there brings this to 0.23 mol m−3. The natural sources of Cl− soils in
would be no decrease in NUE in the CL treatment relative river water are, in decreasing order of importance, weathering
to the BS and SP treatments. The possibility of an effect of of halite, sea salt from dry deposition or rain, and volcanic
Cl− on mesophyll conductance adds further complexity to rocks (Berner and Berner, 1996; Winterton, 2000). Cl− is the
attempts to understand the mechanism of genotypic and phe- dominant anion in seawater, and has the longest replacement
notypic variability of mesophyll conductance (Barbour et al., time (87 Ma) of any seawater component (Berner and Berner,
2016; Raven and Beardall, 2016). 1996). The input of Cl− to the ocean today exceeds removal,
Franco-Navarro et al. (2016) propose that Cl− supplied at with anthropogenic Cl− input and an absence of major evap-
concentrations well in excess of the critical concentration, but orative basins removing Cl− (and Na+) as halite (Berner and

Downloaded from https://academic.oup.com/jxb/article-abstract/68/3/359/2768889 by guest on 24 September 2019


below the concentration at which the growth rate of glyco- Berner, 1996; Winterton, 2000).
phytes is decreased by increased salinity (i.e. 1–5 mol m−3) is Cl− is the main naturally occurring form of the element
best regarded as a ‘beneficial nutrient’, a term usually reserved Cl in the biosphere. However, other inorganic (perchlorate:
for non-essential elements. All of the functions of Cl− (fig. 9 ClO4−) and organic (mainly methyl chloride: CH3Cl) forms
of Franco-Navarro et al., 2016) as an essential micronutrient, occur naturally (see Supplementary Appendix S1 at JXB
or as a beneficial nutrient, in plants (and algae and almost online).
all cyanobacteria) require intracellular Cl−, generally involv- Cl− is usually present at ‘macronutrient’ concentrations
ing active transport (against the electrochemical gradient for in soils (Xu et  al., 1999; White and Broadley, 2001), but
Cl− and so requiring energy input). The Cl– beneficial nutri- there are chloride-deficient soils in terms of plant growth
ent concentration range is in the concentration range of the described for south Sumatra for coconut (Cocos nucifera;
low affinity Cl− influx mechanism: this has either a linear or Braconnier and D’Auzac, 1990; Braconnier and Bonneau,
a hyperbolic (K1/2 4–22  mol m−3) relationship to the exter- 1998), for Columbia for growth of oil palm (Elaeis guineen-
nal Cl− concentration (table 1 of White and Broadley, 2001; sis Jacq.; Ollagnier and Ochs, 1971; Dubos et  al., 2011), for
fig. 9 of Franco-Navarro et al., 2016). The high affinity influx the sandy pampas region of Argentina for the growth of
mechanism in terrestrial glycophytes, charalean freshwater
algae, and a freshwater cyanobacterium has a lower K1/2 of
7–57 mmol Cl− m−3 (Ritchie, 1992b; Beilby and Walker, 1981; Table 1.  Vacuolar (v) and whole vacuolated cell (wvc) Cl−
table 1 of White and Broadley, 2001; fig. 9 of Franco-Navarro concentrations in eukaryotic photosynthetic organisms
et al., 2016).
The exception to the requirement for intracellular Cl− for Organisms External Cl− Internal Cl−
the major role as a micronutrient is in the water-oxidizing Charophyceae aquatic (+wall) 0.1–10 87–179 (v)
and oxygen-producing reactions of PSII in the glycophytic Charophyceae aquatic (+wall) 10–500 174–520
rock-dwelling cyanobacterium Gloeobacter. This is because Chlorophyceae aquatic (+wall) 0.4 30 (v)
Gloeobacter has the two photosystems, with the other mem- Chlorophyceae aquatic (–wall) 400, 1500 283, 325 (v)
brane-associated redox processes of photosynthesis and of Embryophyta Not determined; 670 1–100 (wvc)
respiration, ATP synthetase, and H+-pumping proteorho- Spermatophytes terrestrial (+wall) species of wild collected
dopsin, in the cell membrane, and the Cl−-requiring water- glycophytes
Embryophyta 500 578 (wvc)
oxidizing machinery on the periplasmic (external medium)
Magnoliopsida
side of the membrane (Raven, 2009; Rexroth et  al., 2011).
Submerged halophyte=seagrass
Gloeobacter is the extant cyanobacterium phylogenetically (+wall)
closest to the ancestral cyanobacterium (Blank and Sánchez- Embryophyta 500 600–1620 (wvc)
Baracaldo, 2000). More derived cyanobacteria, with the pho- Magnoliopsida
tosystems in thylakoids, and no PSII in the plasmalemma, Terrestrial and emergent coastal
have active Cl− influx (Ritchie, 1992a, b), but it is not known halophytes(+wall)
if this is the case for Gloeobacter. Presumably such active Rhodophyta 425 74–635 (wvc)
influx is required for the other essential roles of Cl− (fig. 9 of Florideophyceae (+wall)
Franco-Navarro et al., 2016). Ochrophyta 380 44–619 (wvc)
Phaeophyceae (+wall)
Ochrophyta 492 78–445 (wvc)
Bacillariophyceae (+wall)
Availability of chloride Alveolata 492 530 (wvc)
Dinophyta (+wall)
The ultimate source of Cl− in the biosphere is igneous rocks,
with ~200 mg Cl kg–1 (Kuroda and Sandell, 1953). Weathering From Supplementary Appendix S2. 
releases the very soluble Cl− that cycles through water on land Concentrations of Cl− are in mol m−3.
362  |  Raven

Table 2.  External and “cytoplasmic” Cl− concentration (in mol m−3) in cyanobacterial and eukaryotic cells with a minimal fraction of
vacuoles in the cell volume, and the cytosol of vacuolate cells estimated in the specified way and in sieve tubes

Organism Compartment and method External Cl− Internal Cl−


Cyanobacteria Cytosol+thylakoids 0.51 17
Charophyceae Cytosol, Cl− microelectrode 2.1 10
Charales
Embryophyta Cytosol, Cl− microelectrode 1.2 7
Marchantiopsida
Embryophyta Cytosol, Cl− microelectrode 0.4 11
Magnolipiopsida
Chlorophyceae Cytosol, Cl− microelectode 0.6 2.2
Ulvophyceae Cytosol?, compartmental analysis 25.5 194
Ulvophyceae Cytosol?, compartmental analysis 547 231

Downloaded from https://academic.oup.com/jxb/article-abstract/68/3/359/2768889 by guest on 24 September 2019


Enbryophyta, Cytosol?, compartmental analysis 10.2 10–11
Marchantiopsida
Embryophyta Cytosol?, compartmental analysis 0.1 5.7–12.1
Magnoliopsida
Embryophyta Cytosol?, compartmental analysis 1 17–21
Magnoliopsida
Embryophyta Sieve tube sap 0.6–4 0.1–52.1
Magnoliopsida
Embryophyta Sieve tube sap Freshwater 5.4
Halophytic Seawater 28.3
Magnoliopsida
Ochrista Phaeophyceae Sieve tube sap 530 230–310
Laminariales
Embryophyta Cytosol, X-ray microanalysis 0.02 0.5, 19.3
Magnoliopsida Chloroplast, X-ray microanalysis 0.02 0.4, 1.4
Embryophyta Cytosol, X-ray microanalysis 100 91, 176
Magnoliopsida Chloroplast, X-ray microanalysis 100 55, 91
Embryophyta Cytosol, X-ray microanalysis 200 238, 241
Magnoliopsida Chloroplast, X-ray microanalysis 200 87, 271
Chlorophyceae Cytosol, X-ray micoanalysis 400 26 (vacuole 283)
Chlorophyceae Cytosol, X-ray microanlysis 1500 30 (vacuole 325)
Embryophyta Chloroplast, aqueous extraction 0.02 13–20
Magnoliopsida
Embryophyta Chloroplast, aqueous extraction 20 24–36
Magnoliopsida
Embryophyta Chloroplast, aqueous extraction 400 42–64
Magnoliopsida
Chlorophyceae Whole non-vacuolate cell 1.0 1.34
Three freshwater Whole non-vacuolate cell ≤1.0 1.3–2.1
Microalgae
Three marine Whole non-vacuolate cell 100–600 9–35
Microalgae
Chlorophyceae Whole non-vacuolate cell 50–1000 295–497

From Supplementary Appendix S3.

wheat (Triticum aestivum; Diaz-Zorita et  al., 2004), and for stress field environment’ was increased in some years by Cl−
north-western NSW in Australia for growth of durum wheat addition as KCl, with a K+ effect ruled out by adding equimo-
(Schwenke et  al., 2009). Cl− addition to wheat crops in the lar K+ as K2SO4, and decreases in the incidence of corn stalk
Williamette Valley of western Oregon decreased the incidence rot (Heckman, 1995, 1998). Although not increasing yield,
of Gaeumannomyces graminis var. tritici and increased yields Cl− fertilization of pastures (Pehrson et al., 1999; Goff et al.,
(Christensen et al.. 1981). Diaz-Zorita et al. (2004) included 2007) showed that increased Cl− in plant dry matter decreased
a fungicide treatment so fungal disease was shown not to be the inorganic cation:inorganic anion ratio, and the charge-bal-
a factor in their Cl− stimulation of wheat productivity. Grain ancing organic anions decreased the pre-disposition to peri-
yield of Zea mays in New Jersey in the USA under a ‘minimum paturuient hypoglycaemia (milk fever) in cows.
Chloride: essential micronutrient and multifunctional beneficial ion  |  363

Chloride transport anion salt to the root, where catabolism of the organic anions
generates NADPH and/or ATP with excretion of OH− (Raven
What is known of the transport of Cl− across membranes and Smith, 1976; Andrews et  al., 2013) (see Supplementary
of algae and plants (see Supplementary Appendix S2 at Appendix S4 at JXB online). There is also accumulation of
JXB online) is based on physiological/biophysical data and organic anions in NH4+- or N2-grown plants, with H+ efflux and
on molecular genetics of the expression of Cl− transporters influx of some other cation; this is clearly not part of acid–base
and channels at the cell and membrane level. Supplementary regulation in N assimilation, and is more energetically expensive
Appendix S2 shows that the molecular data have been very than accumulation of an external anion such as Cl− (Andrews
helpful in understanding, for example, the role of Cl− chan- et al., 2013) (Supplementary Appendix S4).
nels in the regulation of stomatal aperture. Less helpful in The accumulation of additional organic anions in Ricinus
understanding Cl− transport and its relationship to the ben- communis with NO3− as the N source involves a decrease in
eficial Cl− range in glycophytes is the absence of a molecu- Cl− per mol m−3 of plant water relative to growth on NH4+,
lar genetic explanation for the sound biophysical evidence of with corresponding greater H+ efflux on NH4+ than OH− on
a 2H+:1Cl− symporter at the plasmalemma of glycophytes, NO3− (Allen and Raven, 1987; Tables 3, 4; Supplementary

Downloaded from https://academic.oup.com/jxb/article-abstract/68/3/359/2768889 by guest on 24 September 2019


and the physiological role of a cation–chloride co-trans- Appendix S4). These results are for plants growth on 4.1 mol
porter (CCC) transporter identified by molecular genetics Cl− m−3 (i.e. within the ‘beneficial’ range of Franco-Navarro
(Supplementary Appendix S2). et al., 2016), and Table 5 shows that WUE is higher on NO3−
than on NH4+. This is presumably superimposed on any
higher WUE resulting from Cl− concentration in the ‘benefi-
Functions of Cl− other than as a cial’ range. A less marked effect is found for Phaseolus vulgaris
grow on the two N sources when grown with 0.8 mol Cl− m−3
micronutrient (i.e. just below that ‘beneficial’ range) (Allen et  al., 1988;
Phylogeny of the osmotic role of Cl− Table 5; Supplementary Appendix S4). Franco-Navarro et al.
(2016)) only used NO3− as the N source in showing increased
This is considered in Supplementary Appendices S2 and S3, WUE, with ‘beneficial’ Cl− concentrations relative to basal,
showing that there is a significant phylogenetic and environ- micronutrient, Cl− concentrations.
mental (aclimatory and adaptive) variation in the intracellu-
lar Cl− concentration, together with the range of functions
that turgor, in part maintained by Cl−, fulfils (Tables 1 and 2). Osmolytes driving pressure flow in the phloem
Cl− only contributes 1.5% of the total osmolarity of phloem
sap in the glycophyte R.  communis (Smith and Milburn,
Influence of nitrogen source on the use of Cl− or
1980). Cl− contributes a much larger fraction of osmo-
organic anions as an osmoticum
larity to sieve tube sap of the brown marine macroalga
Franco-Navarro et  al. (2016) only used NO3− as the N Macrocystis pyrifera (see Supplementary Appendix S3 at
source for their work on Nicotiana, although globally JXB online).
many photosynthetic organisms have NH4+ or organic N
as their main N source, and there are diazotrophs catalys- Nastic movements: stomata and pulvini
ing N2 fixation in both natural and agro-ecosystems (see
Supplementary Appendix S4 at JXB online). Raven and The opening phase in stomata and the energy-dependent
Smith (1976) and Raven (1986) examined published data, phase of pulvinar movements involve increased turgor mainly
and concepts, related to acid–base regulation in vascu- driven by the accumulation of ions, with variable contribu-
lar plants as a function of N uptake and assimilation as tions from Cl− and organic anions such as malate, some-
NH4− or NO3−, extended by Raven et  al. (1990) to diazo- times with contributions from NO3− (see Supplementary
trophy. The N assimilation processes and other synthetic Appendix S5 at JXB online). In some cases, at least for
processes produce ~1.3 excess H+ per NH4+-N, ~0.3 excess stomata, Cl− is the only anion involved (Supplementary
H+ per N2- or NH3-N, and 0.7 excess OH− per NO3−-N Appendix S5). Aside from the work of Franco-Navarro
(Raven and Smith, 1976; Raven, 1986; Raven et  al., 1990) et al. (2016) there seem to be no direct studies of the effect
(Supplementary Appendix S4). of Cl− in the beneficial range rather than at lower supplied
There are energetic advantages of NO3− assimilation using Cl− concentrations.
photosynthetic reductant in the shoot rather than by respiration
in the root; the resulting excess OH− is neutralized by organic Action and variation potentials and circulating currents
acid synthesis with accumulation of the organic anion with in Ulvophyceae and Streptophyta
the cation accompanying NO3− up the xylem, although there
are energetic costs of this synthesis that partly or wholly offset Action potentials are essential for the functioning of the
the energetic advantages of photochemical NO3− assimilation carnivorous flowering plant Dionaea, and are involved in
(Raven and Smith, 1976; Andrews et  al., 2013). The costs of the the transmission of other stimuli with effects distant
organic acid synthesis during NO3− assimilation in illuminated from the stimulus (see Supplementary Appendix S5 at JXB
shoots can be largely recouped by transport of the organic online). Circulating currents carried by Cl− within, and in
364  |  Raven

the aqueous medium around, a large cell or a multicellu- Cl− and active water transport?
lar structure, with internal and external currents linked
by transmembrane ion currents, are found in a number of A further active transporter is the electroneutral CCC known
Ulvophyceae and Streoptophyta, where they occur in grow- from genomic studies in some algae and plants (Teakle and
ing pollen tubes (Supplementary Appendix S5). The evi- Tyerman, 2010; Raven and Doblin, 2014; Wegner, 2014;
dence of a unique role for Cl− is often lacking, and even Henderson et  al., 2015; see Boyd and Gradmann, 2002).
where this is available it is not generally clear whether (in These transporters transport two Cl− with two singly charged
glycophytes) a Cl− supply in the beneficial range is needed cations, typically one K+ and one Na+ with ~100 H2O, so they
(Supplementary Appendix S5). have been suggested to be components of active water trans-
port mechanisms in xylem refilling (Wegner, 2014) and in
algae in buoyancy generation in marine diatoms and volume
regulation in effectively wall-less, flagellate, freshwater algal
Table 3.  Mol negative charge m−3 plant water for Ricinus
cells (Raven and Doblin, 2014). However, the intracellular,
communis grown on NH4+ or NO3− as N source
and among cell types, location of CCC in Arabidopsis thali-

Downloaded from https://academic.oup.com/jxb/article-abstract/68/3/359/2768889 by guest on 24 September 2019


ana and Vitis vinifera makes functional assignment difficult
Anion NH4+ as N source NO3− as N source (Henderson et al., 2015).
Cl− 93 64
NO3− 0 37
SO42− 36 20 Influence on glycophytes of external Cl−
P− 24 35
-COO− 56 71
above the benefical range
Total 209 227 The effects of inhibitory NaCl concentrations on glycophytes
have been widely investigated on crop plants in the context of
Assuming all P (organic and inorganic) has one mol negative charge the effects of salinization of agricultural land (Flowers et al.,
per mol P.
Much of the P is in macromolecules (minimal contribution to mol 2015; Munns and Gillihan, 2015). However, separation of the
soluble negative charge) and phospholipids (no contribution to mol inhibitory effects of Cl− rather than Na+ has received much
soluble negative charge). less attention (Flowers et al., 2015); one of the few exceptions
Some of the -COO− is in macromolecules (proteins, cell wall
polysaccharides) with, respectively, little and no contribution to the mol is the experiments on Vicia faba and Hordeum vulgare with
soluble negative charge. Cl− using conterions other than Na+ (Tavakkoli et al., 2010,

Table 4.  Mol excess H+ generated in Ricinus communis growth on NH4+ and measured H+ efflux, and mol excess OH− generated in
growth on NO3− and measured OH− efflux, all in mol m−3 plant water

Reaction NH4+ as N source NO3− as N source


+ −3
Conversion of N source into neutral organic N 387 mol H m plant water 213 mol OH− m−3 plant water
Production of macromolecular and low molecular mass carboxylic acids 56 mol H+ m−3 plant water 71 mol H+ m−3 plant water
Conversion of SO42− into neutral organic S 24 mol OH− m−3 plant water 16 mol OH− m−3 plant water
Total intracellular production of H+/OH− 419 mol H+ m−3 plant water 158 mol H+ m−3 plant water
Measured H+/OH− efflux from roots to the bathing medium 381 mol H+ m−3 plant water 131 mol OH− m−3 plant water

From table 2 of Allen and Raven (1987).

Table 5.  Cl− concentration in the growth medium and in the plant, and water lost in transpiration, during growth on different N sources,
and water lost in transpiration during growth, for Ricinus communis (table 2 of Allen and Raven, 1987) and Phaseolus vulgaris (table 4
of Allen et al., 1988) g water lost in transpiration is calculated from g water uptake during growth on the specified N source minus the g
water in the plants and minus the water used in metabolism during photosynthetic growth (18 g water per mol C)

Plant, N source Mol Cl− m−3 water in growth medium Mol Cl− m−3 water in the plant mg dry matter gain g−1 water lost
Ricinus communis 4.1 93 1.67
NH4+
Ricinus communis 4.1 64 2.33
NO3−
Phaseolus vulgaris 0.8 73 2.27
NH4+
Phaseolus vulgaris 0.8 7 2.50
NO3−
Phaseolus vulgaris 0.8 37 1.79
N2
Chloride: essential micronutrient and multifunctional beneficial ion  |  365

2011). Geilfus and Mühling (2013) showed for V.  faba that Appendix S2. Phylogeny of Cl– transport and the osmotic
Cl− (not Na+) at 70  mol m−3 in the culture medium causes role of chloride from whole-organism and vacuolar analyses.
transient alkalinization of the leaf apoplast, modulated by Appendix S3. Cl– concentration in the total cytoplasm and
the plasmalemma H+ ATPase. Cl− at 70–100 mol m3 or more compartments of the cytoplasm.
inhibits a number of enzymes (Ritchie and Larkum, 1984), Appendix S4. Influence of nitrogen source on the use of
and especially those with anionic substrates (Gimmler et al., chloride rather than organic anions as an osmoticum in ter-
1984), and Cl− is equally, or even more, inhibitory to in vitro restrial embryophytes
protein synthesis than Na+ (Flowers and Dalmond, 1992), Appendix S5. The roles of Cl– in nastic responses, action
More work is needed in which the inhibitory effects of Cl− and variation potentials, circulating ionic currents, response
are distinguished from those of Na+ to decreased external osmolarity, and active water transport.

Acknowledgements
Conclusions and recommendations for
Discussions with Susan Allen, Mitchell Andrews, Mary Beilby, Mary

Downloaded from https://academic.oup.com/jxb/article-abstract/68/3/359/2768889 by guest on 24 September 2019


further work Bisson, Tim Colmer, Martina Doblin, Linda Handley, Hans Lambers, Enid
MacRobbie, Janet Sprent, F.  Andrew Smith, N.  Alan Walker, and Philip
The role of Cl− as an essential micronutrient for oxygenic White are gratefully acknowledged. The University of Dundee is a registered
photosynthetic organisms is widely accepted, largely on the Scottish charity, No. 015096
basis of structural and functional studies on isolated PSII
complexes. It would be useful to extend this detailed under-
standing to other known (V-type H+ ATPase, asparagine References
synthetase) and possible (amylase) Cl−-stimulated enzymes, Allen S, Raven JA. 1987. Intracellular pH regulation in Ricinus communis
as well as to extend attempts to demonstrate essentiality of grown with ammonium of nitrate as N source: the role of long distance
transport. Journal of Experimental Botany 38, 580–596.
Cl− for the growth of oxygenic photolithotrophs.
Allen S, Raven JA, Sprent JI. 1988. The role of long-distance transport
The high concentrations of Cl− generally found in algal in intracellular pH regulation in Phaseolus vulgaris grown with ammonium
and plant cells, and particularly in their vacuoles, when grown or nitrate as nitrogen source, or nodulated. Journal of Experimental Botany
with Cl− concentrations more than sufficient to satisfy micro- 39, 513–528.
nutrient concentrations have been known for many decades, Andrews M, Raven JA, Lea PJ. 2013. Do plants need nitrate? The
mechanisms by which nitrogen form affects plants, Annals of Applied
as has the major role it plays in turgor generation and in Biology 163, 174–199.
cell expansion. It is also known that these roles of Cl− vary Arnon DI, Whatley FR. 1949. Is chloride a cofactor in photosynthesis?
phylogenetically and with the N source. Less attention has Science 110, 554–556.
been paid to how the organisms respond to manipulations of Barbour MM, Evans JR, Simonin KA, von Caemmerer S. 2016.
the Cl− content. The work of Franco-Navarro et  al. (2016) Online CO2 and H2O oxygen isotope fractionation allows estimation
of mesophyll conductance in C4 plants, and reveals that mesophyll
increases our knowledge of the outcome of such manipula- conductance decreases as leaves age in both C4 and C3 plants. New
tions, and proposes that Cl− at concentrations in excess of Phytologist 210, 875–889.
the concentration needed to satisfy the micronutrient require- Beilby MJ, Walker NAW. 1981. Chloride transport in Chara. I. Kinetics
ment but below those causing Cl− toxicity is a beneficial and current voltage curves for probable proton symport. Journal of
Experimental Botany 136, 42–54.
nutrient, thus becoming the first essential micronutrient to
Berner EK, Berner RA. 1996. Global environment: water, air and
be suggested also to be a beneficial nutrient. The essential- geochemical cycles. Princeton, NJ: Princeton University Press.
ity of Cl− for nastic (stomata, puvini) responses, circulating Blank CE. 2013. Origin and early evolution of photosynthetic eukaryotes
currents, and action and variation potentials is difficult to in freshwater environments: reinterpreting Palaeozoic palaeobiology and
establish. This is because Cl− is a micronutrient that must biogeochemical processes in light of trait evolution. Journal of Phycology
49, 1040–1055.
be present at low concentrations in the growth medium, and
Blank CE, Sánchez-Baracaldo P. 2010. Timing of the morphological
there is a relatively higher affinity Cl− uptake system in many and ecological innovations in the cyanobacteria—a key to understanding
organisms. However, progress is possible using experiments the rise in atmospheric oxygen. Geobiology 8, 1–12.
maniputating internal Cl− concentrations with measurements Boussac A, Ishida N, Sugiura M, Rappoport F. 2012. Probing the
of the distribution of Cl− between and within cells, combined role of chloride in Photosystem II from Thermosynechococcus elongatus
by exchanging chloride for iodide. Biochimica et Biophysica Acta 1817,
with measurements of nastic (stomata, puvini) responses, cir- 802–810.
culating currents, and action and variation potentials. Boyd CM, Gradmann D. 2002. Impact of osmolytes on buoyancy of
Finally, further data are needed on how the inhibitory marine phytoplankton. Marine Biology 141, 605–618.
effects of high concentrations on Cl− are distinguished Braconnier S, Bonneau X. 1998. Effects of chlorine deficiency in the field as
from those of Na+ on plant growth and on in vitro enzyme a leaf gas exchanges in the PB121 coconut hybrid. Agronomie 68, 563–572.
activities. Braconnier S, d’Auzac F. 1990. Chloride and stomatal conductance in
coconut. Plant Physiology and Biochemistry 28, 105–112.
Brownell PF, Crossland CJ. 1972. The requirement for sodium as
micronutrient by species having the C4 dicarboxylic photosynthetic
Supplementary data pathway. Plant Physiology 49, 794–797.
Brownell PF, Crossland CJ. 1974. Growth responses to sodium by
Supplementary data are available at JXB online. Bryophyllum tubiflorum under conditions inducing crassulacean acid
Appendix S1. Natural occurrence of Cl species other than Cl–. metabolism. Plant Physiology 54, 416–417.
366  |  Raven
Broyer TC, Carlton AB, Johnson CM, Stout PR. 1954. Chlorine—a Kuroda PK, Sandell EB. 1953. Chlorine in igneous rocks. Some aspects
micronutrient element for higher plants. Plant Physiology 29, 526–532. of the geochemistry of chlorine. Geological Society of American Bulletin
Christensen UW, Taylor RG, Jackson TL, Mitchell BL. 1981. Chloride 64, 879–896.
effects on water potentials and yield of winter wheat infected with take-all Munns R, Gilliham M. 2015. Salinity tolerance of crops—what is the
root rot. Agronomy Journal 73, 1053–1058. cost? New Phytologist 208, 668–673.
Churchill KA, Sze H. 1984. Anion-sensitive H+ pumping ATPase of oat Ollagnier M, Ochs R. 1971. Le chlore, nouvel element essential dans la
roots. Direct effect of Cl−, NO3− and a disulfonic stilbene. Plant Physiology nutrition du palmier á huile. Oléagineux 36, 409–421.
76, 490–497. Pehrson B, Svensson S, Gruvaens I, Virkki M. 1999. The influence of
D’Amico S, Gerday C, Feller G. 2000. Structural similarities and acidic diets on the acid–base balance of dry cows and effect of fertilization
evolutionary relationships in chloride-dependent α-amylases. Gene 253, on the mineral content of grass. Journal of Dairy Science 82, 1310–1316.
95–105. Pokhrel R, McConnell IL, Brudwig GW. 2011. Chloride regulation of
Diaz-Zorita M, Duarte GA, Barroco M. 2004. Effect of chloride enzyme turnover: application to the role of chloride in photosystem II.
fertilization on wheat (Triticum vulgare L.) productivity in the sandy pampas Biochemistry 50, 2725–2734.
region, Argentina. Agronomy Journal 96, 839–844. Raven JA. 1986. Biochemical disposal of excess H+ in growing plants?
Dubos B, Alarcón WH, López JE, Ollivier J. 2011. Potassium uptake New Phytologist 104, 175–206.
and storage in oil palm organs: the role of chlorine and the influence of Raven JA. 2009. Functional evolution of photochemical energy
soil characteristics in the Magdalena Valley, Columbia. Nutrient Cycling in transformations in oxygen-producing organisms. Functional Plant Biology

Downloaded from https://academic.oup.com/jxb/article-abstract/68/3/359/2768889 by guest on 24 September 2019


Agroecosystems 89, 219–227. 36, 505–515.
Eyster HC. 1952. Optimum chloride concentration for malt diastase as Raven JA. 2013. The evolution of autotrophy in relation to phosphorus
function of temperature. Plant Physiology 28, 140–142. requirement. Journal of Experimental Botany 64, 2023–4046.
Eyster C. 1958. Chloride effects on the growth of Chlorella pyrenoidosa. Raven JA, Beardall J. 2016. The ins and outs of CO2. Journal of
Nature 181, 1141–1142. Experimental Botany 65, 1415–1424.
Field CB, Behrenfeld MJ, Randerson JT, Falkowski P. 1998. Primary Raven JA, Doblin MA. 2014. Active water transport in unicellular algae:
production of the biosphere integrating terrestrial and marine components. where, why and how? Journal of Experimental Botany 65, 6279–6292.
Science 281, 153–161.
Raven JA, Franco AA, de Jesus EL, Jacob-Neto J. 1990. H+ extrusion
Flowers TJ, Dalman D. 1992. Protein-synthesis in halophytes—the and organic acid synthesis in N2-fixing symbioses involving vascular plants.
effects of potassium, sodium and magnesium in vitro. Plant and Soil 146, New Phytologist 114, 369–389.
153–161.
Raven JA, Handley LL, Wollenweber B. 2004. Plant nutrition and water
Flowers TJ, Galal HK, Bromham L. 2010. Evolution of halophytes: use efficiency. In: Bacon HA, ed. Water use efficiency in plant biology.
multiple origins of salt tolerance in land plants. Functional Plant Biology 37, Oxford: Blackwell Publishing, 171–197
604–612.
Raven JA, Smith FA. 1976. Nitrogen assimilation and transport in
Flowers TJ, Munns R, Colmer TD. 2015. Sodium chloride toxicity and
vascular land plants in relation to intracellular pH regulation. New
the cellular basis of salt tolerance in halophytes. Annals of Botany 115,
Phytologist 76, 415–431.
419–431.
Rexroth S, Mullineaux GD, Ellinger D, Sendtko E., Rögner M,
Franco-Navarro JD, Brumós J, Rosales MA, Cubero-Font P, Talón
Koenig F. 2011. The plasma membrane of the cyanobacterium
M, Colmenero-Flores JM. 2016. Chloride regulates leaf cell size and
Gloeobacter violaceous contains segregated bioenergetics domains. The
water relations in tobacco plants. Journal of Experimental Botany 67,
Plant Cell 23, 2379–2390.
873–891.
Geilfus C-M, Mühling K-H. 2013. Ratiometric monitoring of transient Ritchie RJ. 1992a. The cyanobacterium Synechococcus R-2 (Anacystis
apoplastic alkalinizations in the leaf apoplast of living Vivia faba plants: nidulans, S. leopolensis) PCC 7942 has a sodium-dependent chloride
chloride primes and PM-H+-ATPase shapes NaCl-induced system transporter. Plant, Cell and Environment 15, 163–177.
alkalizations. New Phytologist 197, 1117–1129. Ritchie RJ. 1992b. Kinetics of chloride transport in the cyanobacterium
Gimmler H, Kaaden R, Kirchner U, Weyand A. 1984. The chloride Synechococcus R-2 (Anacystis nidulans, S. leopolensis) PCC 7942. Plant,
sensitivity of Dunaliella parva enzymes. Zeitschrift fűr Pflanzenphysiologie Cell and Environment 15, 179–184.
114, 131–150. Ritchie RJ, Larkum AWD. 1984. Chloride transport in Enteromorpha
Goff JP, Brummer EC, Henning SJ, Dooreenbos RK, Horst RL. 2007. intestinalis (L.) Link. New Phytologist 97, 319–345.
The application of ammonium chloride and calcium chloride to alfalfa Rivalta I, Amin M, Luber S, et al. 2011. Structure–functional role of
cation–anion content and yield. Journal of Dairy Science 90, 5159–5164. chloride in photosystem II. Biochemistry 50, 6312–6315.
Guskov A, Kern J, Gasbdulkhakov A, Broser M, Zouni M, Saenger Rognes SE. 1980. Anion regulation of lupin asparagine synthetase:
W. 2009. Cyanobacterial photosystem II at 2.9 Å resolution and role of chloride activation of the glutamine-utilizing reaction. Phytochemistry 19,
quinones, lipid, channels and chloride. Nature Structural and Molecular 2287–2293.
Biology 16, 334–342.
Schwenke G, Simpfendorfer S, Giblett GJ. 2009. Chloride deficiency
Heckman JR. 1995. Corn responses to chloride in maximum yield in wheat. In: 2009 GRDC Advisor Update: Dubbo 24th and 25th February.
research. Agronomy Journal 87, 415–419. 17–19. Goondiwindi 3rd and 4th March 2009. 138–141.
Heckman JR. 1998. Corn stalk rot suppression and grain yield response
Smith JAC, Milburn TA. 1980, Osmoregulation and the control of
to chloride. Journal of Plant Nutrition 87, 415–419.
phloem sap composition of Ricinus communis L. Planta 148, 28–34.
Henderson SW, Wege S, Qiu J, Blackmore DH, Walker AR, Tyerman
SD, Walker RR, Gillihan M. 2015. Grapevine and Arabidopsis cation– Tavakkoli E, Rengasamy P, McDonald GK. 2010. High concentrations
chloride cotransporters localize to the Golgi and trans-Golgi network and of Na+ and Cl− ions in soil solution have simultaneous detrimental effects
indirectly influence long-distance ion transport and plant salt tolerance. on growth of faba beans under salinity stress. Journal of Experimental
Plant Physiology 169, 2215–2229. Botany 61, 4449–4459.
Johnson CM, Stout PR, Broyer TC, Carlton AB. 1957. Comparative Tavakkoli E, Fatehu F, Coventry S, Rengasamy P, McDonald GK.
chlorine requirements of different species. Plant and Soil 8, 337–353. 2011. Additive effects of Na+ and Cl− ions on barley under salinity stress.
Journal of Experimental Botany 62, 2189–2203.
Kato Y, Shibamoto T, Yamamoto S, Watanabe T, Ishida N, Sugiura
M, Rappoport F, Boussac A. 2012. Influence of the PsbA1/PsbA3, Ca2+/ Teakle NL, Tyerman SD. 2010. Mechanisms of Cl− transport contributing
Sr2+ and Cl−/Br− exchanges on the redox potential of the primary quinone to salt tolerance. Plant, Cell and Environment 33, 566–589.
QA in Photosystem II from Thermosynechoccus elongatus as revealed by Terry N. 1977. Photosynthesis, growth and the role of chloride. Plant
spectroelectrochemistry. Biochimica et Biophysica Acta 1817, 1998–2004. Physiology 60, 69–75.
Kirkby EA, Armstrong MJ. 1980. Nitrate uptake by roots as regulated Wang L, Zhang C, Zhao J. 2014. Location and function of the high-
by nitrate assimilation of the shoot of castor bean plants. Plant Physiology affinity chloride in the oxygen-evolving complex. Implications from
65, 6–14. comparing studies on Cl−/Br−/I− substituted photosystem II using two
Chloride: essential micronutrient and multifunctional beneficial ion  |  367
different methods. Journal of Photochemistry and Photobiology B: Biology Woodward J. 1699. Some thoughts and experiments concerning
138, 249–255. vegetation. Philosophical Transactions of the Royal Society 21, 193–227.
Wegner LH. 2014. Root pressure and beyond: energetically uphill water Xu G, Magen H, Tarchitky J, Kafkafi U. 1999. Advances in chloride
transport into xylem vessels? Journal of Experimental Botany 65, 381–393. nutrition of plants, Advances in Agronomy 68, 97–150.
White PJ, Broadley MR. 2001. Chloride in soils and its uptake and Zhang Z, Coats KL, Chen Z, Hubin TJ, Yin G. 2014. Influence of
movement within plants: a review. Annals of Botany 88, 967–988. calcium(II) and chloride on the oxidative reactivity of a manganese(II)
Winterton N. 2000. Chlorine: the only green element—towards wider complex of a cross-bridged cyclen ligand. Inorganic Chemistry 53,
acceptance of its natural cycles. Green Chemistry 2, 173–225. 11973–11946.

Downloaded from https://academic.oup.com/jxb/article-abstract/68/3/359/2768889 by guest on 24 September 2019

Das könnte Ihnen auch gefallen